You are on page 1of 837

DEVELOPMENTS IN SEDIMENTOLOGY 44

Clays, Muds, and Shales


FURTHER TITLES IN THIS SERIES
VOLUMES 1- 1 1, 13- 15 and 2 1-24 are out of print

12 R.C.G. BATHURST
CARBONATE SEDIMENTS AND THEIR DIAGENESIS
13 H.H. RIEKE I//and G. V. CHILINGARIAN
COMPACTION OF ARGILLACEOUS SEDIMENTS
17 M .D. PICARD and L.R. HIGH Jr.
SEDIMENTARY STRUCTURES OF EPHEMERAL STREAMS
18 G. V. CHILINGARIANand K.H. WOLF, Edirors
COMPACTION OF COARSE-GRAINED SEDIMENTS
19 W . SCHWARZACHER
SEDIMENTATION MODELS AND QUANTITATIVE STRATIGRAPHY
2 0 M.R. WALTER, Editor
STROMATOLITES
2 5 G. LARSENand G. V. CHILINGAR. Editors
DIAGENESIS IN SEDIMENTS AND SEDIMENTARY ROCKS
2 6 T. SUDO and S. SHIMODA, Editors
CLAYS AND CLAY MINERALS OF JAPAN
2 7 M .M . MORTLAND and V.C. FARMER, Editors
INTERNATIONAL CLAY CONFERENCE 1978
2 8 A . NISSENBAUM, Editor
HYPERSALINE BRINES AND EVAPORITIC ENVIRONMENTS
2 9 P. TURNER
CONTINENTAL RED BEDS
30 J.R.L. ALLEN
SEDIMENTARY STRUCTURES
3 1 T. SUDO, S. SHIMODA, H. YOTSUMOTO and S. AITA
ELECTRON MICROGRAPHS OF CLAY MINERALS
3 2 C.A. NITTROUER, Editor
SEDIMENTARY DYNAMICS OF CONTINENTAL SHELVES
33 G.N. BATURIN
PHOSPHORITES ON THE SEA FLOOR
3 4 J.J. FRlPIAJ,Editor
ADVANCED TECHNIQUES FOR CLAY MINERAL ANALYSIS
3 5 H. VAN OLPHENand F. VENIAL€, Editors
INTERNATIONAL CLAY CONFERENCE 198 1
3 6 A . IIJIMA, J.R. HEINand R. SIEVER, Editors
SILICEOUS DEPOSITS IN THE PACIFIC REGION
3 7 A . SINGERand E. GALAN, Editors
PALY GORSKITE-SEPIOLITE:OCCURRENCES, GENESIS AND USES
38 M.E. BROOKFIELDand T.S. AHLBRANDT, EditorsAOLlAN SEDIMENTS AND PROCESSES
3 9 B.GRfENWOODand R.A. DAVIS Jr., Edirors
HYDRODYNAMICS AND SEDIMENTATION IN WAVE-DOMINATED COASTAL ENVIRONMENTS
4 0 5. VELDE
CLAY MINERALS - A PHYSICO-CHEMICAL EXPLANATION OF THEIR OCCURRENCE
4 1 G.V. CHlLlNGARIANand K.H. WOLF, Edttors
DIAGENESIS, I
4 2 L.J. DOYLEand H.H. ROBERTS, Editors
CARBONATE-CLASTIC TRANSITIONS
4 3 G. V. CHILINGARIA N and K .H. WOLF, Ed/tors
DIAGENISIS, II
4 4 C.E. WEAVER
CLAYS, MUDS, AND SHALES
45 G.S. ODIN, Editor
GREEN MARINE CLAYS
4 6 C.H. MOORE
CARBONATE DlAGENESlS AND POROSITY
DEVELOPMENTSIN SEDIMENTOLOGY 44

Clays, Muds, and Shales


Charles E. Weaver
School of Geophysical Sciences, Georgia Institute of Technology, Atlanta,
GA 30332, U.S.A.

ELSEV IER
Amsterdam - Oxford - New York - Tokyo 1989
ELSEVIER SCIENCE PUBLISHERS B.V.
Sara Burgerhartstraat 25
P.O. Box 2 1 1, 1000 AE Amsterdam, The Netherlands

Distributors for the United States and Canada:

ELSEVIER SCIENCE PUBLISHING COMPANY INC


655, Avenue of the Americas
New York, NY 10010, U S A .

L i b r a r y o f Congress C a t a l o g i n g - i n - P u b l i c a t i o n Data

W e a v e r , C h a r l e s E. ( C h a r l e s E d w a r d )
Clays. muds, and s h a l e s / C h a r l e s E. Weaver.
p. cm. -- ( D e v e l o p m e n t s in s e d i m e n t o l o g y ; 44)
Includes bibliographical references.
I S B N 0-444-87381-3 ( U . S . )
1. C l a y . 2 . Mud. 3. Shale. 4. S e d i m e n t a t i o n and d e p o s i t i o n .
I. T i t l e . 11. S e r i e s .
QE471.3.W43 1989
552'. 5--d~20 89-2361 1
CIP

ISBN 0-444-87381-3

0Elsevier Science Publishers B.V., 1989

All rights reserved. No part of this publication may be reproduced, stored in a retrieval system or
transmitted in any form or by any means, electronic, mechanical, photocopying, recording or
Otherwise, without the prior written permission of the publisher, Elsevier Science Publishers B.V./
Physical Sciences & Engineering Division, P.O. Box 330, 1000 AH Amsterdam, The Netherlands.

Special regulations for readers in the USA - This publication has been registered with the Copyright
Clearance Center Inc. (CCC), Salem, Massachusetts. Information can be obtained from the CCC
about conditions under which photocopies of parts of this publication may be made in the USA. All
other copyright questions, including photocopying outside of the USA, should be referred t o the
publisher.

No responsibility is assumed by the Publisher for any injury and/or damage to persons or property
as a matter of products liability, negligence or otherwise, or from any use or operation of any meth-
ods, products, instructions or ideas contained in the material herein.

This book has been printed on acid-free paper

Printed in The Netherlands


This book is dedicated
to Junice Hurtlund Weuuer, my wife.
who hus niude life interesting
This Page Intentionally Left Blank
PREFACE

I was a student at Penn State (and in the Navy), 1942-1952, shortly after the
basic tenets of clay mineralogy had been established. I took numerous courses in
sandstone petrology and limestone petrology, primarily from P.D. Krynine. Shales
were cursorily described as being gray, green, red, or black, and thick or thin. After
hearing a lecture by Tom Bates on clay minerals, I thought, “Wouldn’t it be
interesting to determine what clay minerals occur in the various shales.” With
various excursions, I have been doing that ever since.
I am proud to have helped pioneer in the use of clay minerals (physils) for
geologic interpretation. Several years ago I decided we finally knew enough about
the clay mineralogy of “muds” and “shales” that I should write a book on clay
petrology to go along with the many books on sandstone and limestone petrology.
This is it. I would like to take this opportunity to pay my respects to G. Millot for
his 1962 precocious book “The Geology of Clays”.
I tried to make this book comprehensive enough so that it could be used both as
a reference book for all geologists and as a text for the specialist. Professors will
have to be selective in their choice of material.
The book is basically a critical summary of a large portion (approximately 1600
quoted references) of the clay mineral literature that relates to geology and geologic
processes. In some instances I took the liberty of offering interpretations that
differed from those of the original authors. I hope I have not offended too many. T o
you authors whose papers I missed, I apologize; I wish I were knowledgeable
enough to have been able to include more non-English publications.
Ray Ferrell kindly reviewed the book and made many helpful suggestions. Ray
can not be held accountable for errors; there was too much material for one person
to review in detail. Thanks Ray, I appreciate it. I am deeply indebted to Pat Rice
(now with the USGS), Linda Schwanke and my wife who did most of the typing
(word processing), and to my graduate students Martha Carr and Susan Stell who
helped with the editing. They were helpful and compassionate during my harass-
ment.
The Georgia Tech administrators generously allowed me the time to complete
this book, now it’s back to obtaining research funds.
This Page Intentionally Left Blank
Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ix

01/04/19 (1) Chapter I. Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

01/04/19 (1) Classification of Physils . . . . . . . .............................. 1


01/04/19 (1) Structural Formula . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ...... 3
Nomenclature . . . . . . . ........................................... 5
Phyllosilicates and Rocks . . . . . . . . . . . . . . . . . . . ................. 5
Size Distribution and Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
NewTerms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

Chapter 11. Structure and Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Kaolinite, Dictite, Nacrite .......................... 13

Chemistry . . . . . . . . ............. 22

........ ................

Mite . . . . . . . . . . ............ 32

...... 44
...... 46

Nontronite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
Trioctahedral Smectites . . ............. 54
Mixed-layer Illite/Smectite (1,’s) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
X-ray . . . . . . . . . . ..... ..... .... ......... 51
Chemistry . . . . . . . ..... ..... .... ......... 62
Regular Mixed-Layer ............................................ 66
Chlorite and Chloritics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Chlorite . . . . . . ...... ........................... 67
Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67
X-ray . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ... 68
X

Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . ....... . 72
High temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
Low temperature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
Chloritic Physils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77
Vermiculite . . . . ... ...... ....... .................. 82
... ...... ....... .................. 82
.................................................... 83
Macroscopic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83
Microscopic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 84
Glauconite Celadonite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86
Structure . .. ..... ....... 87
Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Glauconite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88
Celadonite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
Talc, Kerolite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
.. ........... 93
.................................................. 94
Sepiolite, Palygorskite . . . . . . .. 95
Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
Chemistry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98
Sepiolite .... ........ ........................ 98
Palygorskite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 99

Chapter Ill. Soils and Weathering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103

Soil Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103


Physical Weathering and Erosion ..................................... 103
Chemical Weathering . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............. 104
Soil . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............. 104

Zonal Soils . . . . . ........................ 109

Vertisols . . . . . . . . . . . . . . . . . . ........... 109

........

Ultisols . . . . . ............. ........................ 111

Inceptisols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ........
Climate (and Other Factors) ......... ........................ 112

Clay Content . . . . . . . . . . . . . . . . . . . ............ ............ 113

Chemical Weathering Processes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


......................... 122
................................................... 122
Experimental . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ..........
Natural . . . . . . . . ...................................... 130
Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Organic . . . . . . . . . . . . . . . . . . . . . . . ............ 134
Microbial Alteration . . . . . . . . . . . . . . ............ 137
xi

Pedogenic Formation of Physils . . . . . . . . . .......................... 142


Kaolin and Related Materials . . . . . . . . 142
Tropical Climate . 142
Temperate Climate ................................ 146
Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 152
........................ ................. 153
Tropical Climate . ......................................... 155
Temperate Climate . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . _... 156
Discussion . . . . . . . . ................................ 160
Vermiculite . . . . . . . . , . ................................ 161
Occurrence of Trioctahedral Vermiculite . . . . . . . . . . . . . ................ 166
Occurrence of Dioctahedral Vermiculite . . . . . . . . . . . . . . . . . . . . . . . . . . . , . . . . . . . . 168
.................................... ........ 169
................ 170
Mica-Illite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .. 174
Palygorskite and Sepiolite . . . . . . . . . . . . ................ 182
................................................... 185
....................................... .. 186

Chapter IV. Continental Transport and Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189

Rivers ...................................... 189


Intr .......................................................... 189
Transportation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 189
Mineralogy . . . . . . . . . . . ........................... 193
Estuaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20 5
Settling. ..................................... 205
Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 220
Biodeposits . . . . . . . .......................................... 224
Physils in Estuaries . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 230
Marshes and Tidal Flats ................................................ 244
Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 244
Physils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 246
Deltas . . . . . . . . . . . . . . . . . 248
Formation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 248
.......... 25 1
..................... 255
Amazon Delta ............................................... 258
PoDelta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ...................... 262
NileDelta . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263
264
Lakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 265
........................................ 265
SalineLakes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 270

Chapter V. Marine Transport and Deposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279

Water Transport . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279


Introduction . . . . . . . . . . . . .. ................................. 279
Shelf . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 279
Slope, Rise, and Abyssal Plain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 285
Sediment Distribution . . . . ....................... .............. 285
xii

.............. 285

................................

............... 308

Atlantic Ocean .................................................... 314


.............. 315

Expandable Physils ............................

Other Physils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Suspended Physils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Pacificocean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 327
............... 327
.............. 333
Indian Ocean . . . .............. 339
Southern Indian Ocean ................................................ 339
........................... 341
Middle Indian Ocean . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 342

Chapter VI. “Authigenic Marine” Physils . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 345

Exchange Reactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ...... 345


Physil Dissolution and Precipitation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Formation of Physils from Marine Volcanics . . . . . . . . . . . . . . . . . . . . . . . . .
Introduction . . . . . ............... 351

............... 354
High To Low Temperature Transition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 363
............... 365
Hydrothermally Altered Sediments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 371
Ambient Marine Smectite
Palygorskite and Sepiolite ............... 379
Miscellaneous “Authigeni .......................................... 386
Physils That Grow in Shallow Marine, Brackish, and Evaporitic Environments . . . . . . . . . . . . . . 387
Glaucony and Bertheirine . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
Origin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 387
Isotopic Age Dating ................ 397
Palygorskite and Sepiolite . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 400
Evaporite Physil ... 407
Brief Summary ................ 414

Chapter VII. Diagenesis Metamorphism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417

Diagenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Early Diagenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Middle Diagenesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 418
................. 418
...
Xlll

Smectite + Illite . . . . . . . . . . . . . . . . . . . . . . ..... ............. 418


.............. ..... ............. 430
........................................ 440
........................................ 443
...............................

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Illite-Mica . . . . . . . . . . . . . . . . . . . . . . . . . .............. 454
Petrology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 460
............ 461
Anchizone . . . . . .

Layer Silicate Films . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 478


Lattice Fringe Images . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ............. 481
......................................................... 483
Oxygen Isotopes . . . . .................... ............ 489
Potassium Argon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 491

................................................. 496
Temperature and Tectonics ...........

Organic and Physil Paleothermometers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5 13


Coalification Process . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 513
Shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
Organic Material . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 517
Physil-Organic Paleothermometry . . . . . . . . . . . . . . . . . . . . . . . . . 520

Chapter VIII. Physils in Sandstones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 525

........................ .................. 525


f Detrital and Authigenic Physils .................. 526
Composition . . . . . . . . . . . . . . . . . . . . . . .................. 529
Morphology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
Texture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 529
................. 529

................. 531

Smectite . . . .

Limestones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 560
xiv

Chapter IX. Evolution of Physils and Continents . . . . . . . . _ _ _ _ . _561

...... 561
Precambrian . . . . . . . . . . . . . . . _ . _ _ _ 568

Cambrian . . . . . . . . . . . ............... ....................... 516

Ordovician . . . . . . . . . . . . . . . . . . . .
North America . . .
Europe, Africa . . . ........... ........ 582

North America ................................................ 585

....................... .................... 589

Europe. Africa, South America . . . . . . . . . . . . . . . . . ...................... 594


Carboniferous . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ...................... 598
North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 598
.............. 598
..................................... 602
Underclays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 610
Discussion Eastern and Midcontinent Regions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 615
Western North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 616
General Discussion .................... .............. 618
.............
..........................
Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
South America . . . . . . . . . . . . . . . . . . . . . . . . . .................... 622
Tonsteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Permian . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 625
North America ................... 625
Europe, Africa, South America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 629
Triassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .......................... 630
North America . . . . . . . . . . . . . . . . . . . . . . . . . . .......................... 632
Europe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 634
England . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 636
Germany . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 638
France ....................... . 638
Switzerland . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 641
Other Areas . . . . . . . . . . . . . . . . . . . . . . . . . . ................ 641
Spain . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . ................ 641
Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 643
Israel and Sinai . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 644
South America ........................... ... 644
........................................... 645
Jurassic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 645
North America ............. ................................. 645
Europe,Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 649
......................... ... 652
......................................... 653
North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 655
Western North America ................ ............ 655
Coastal Plains . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 661
xv

.............. 665
.............. 668
South America . . . . . . . . . . . ............... 669
Cenozoic . . . . . . . . . . . . . . . . . . ............... 670

Western North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .


GulfCoast . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 676
Eastern North America . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 678
Europe . . . . . . . . . . . . . . . . . . . . . . . . . . . . .................. 679
Africa . . . . . . . . . . . . . . . . . . . . . . . . . . . . .................. 690
South America . . . . . . . .

Paleoatlantic . . . . . . . . . . ...................

.................................................. 695
Mesozoic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 696
Cenozoic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 699

Chapter X . Lithification and Petrology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705

Compaction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 705
Sedimentary Structures ............................................ 712
Composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 714
Color . . . . . . ................................................ 720
Thin Sections ........................... 721

References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 723

Authorlndex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 787

Subjectlndex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 809
This Page Intentionally Left Blank
1

Chapter I

BACKGROUND

01/04/19 (1) Classification of Physils

The clay minerals are small hydrous layer silicates and are part of the phyllosili-
cate family. As discussed in Chapter 11, many of the hydrous layer silicates in clays,
muds, soils, shales, slates, etc. are coarser than clay ( < 2 or < 4 pm). Primarily for
this reason I have suggested (Weaver, 1980) using the term physil, which has no size
connotation, to refer to the low-temperature ( 5 400°C) hydrous layer silicates.
The layer silicates are constructed of planes of atoms forming tetrahedral and
octahedral sheets arranged in various combinations (Bailey, 1980a). The tetrahedral
sheets are composed of tetrahedra linked with adjacent tetrahedra by sharing
oxygen ions at three corners. The shared oxygen (basal oxygens) form a hexagonal
pattern. The fourth tetrahedral oxygen (apical oxygen) of all tetrahedra is per-
pendicular to the sheet and forms part of the adjacent octahedral sheet (Figure 1-1).
The tetrahedral cations are primarily Si and Al, and rarely Fe3+. The octahedral
sheet consists of cations (Al, Fe, Mg) that are octahedrally coordinated by shared
apical oxygens plus unshared OH’S that lie at the center of the hexagonal hole
formed by the basal oxygens.
The smallest structural unit contains three octahedra. I f all three octahedra are
occupied with cations,the sheet is classified astrioctrahedral. If only two octahedra
are occupied and one is vacant,the sheet is classified as dioctahedral.
The combination of one tetrahedral sheet and one octahedral sheet is called a 1:1
layer. The unshared plane of anions in the octahedral sheet consists of OH anions.
A 2:l layer consists of an octahedral sheet sandwiched between two tetrahedral
sheets. Simplified sketches of the major physil groups are shown in Figure 1-1.
Where the tetrahedral cations are all Si and the octahedral cations all A1
(dioctahedral) or all Mg (trioctahedral), the layers are electrostatically neutral
(pyrophyllite and talc) and held together by Van der Waals bonds. Most physils
have some isomorphous substitution of lower charged cations for higher charged
cations. This produces a negative layer charge which is neutralized by positively
charged material between the layers (interlayer material). Interlayer material can be
individual cations (micas), hydrated cations (expanded physils), hydroxide groups
(chloritic physils), and hydroxide octahedral sheets (chlorite). The latter combina-
tion, chlorites, is referred to as a 2 : 2 or 2 : l : l structural unit. The 1:1, 2:l and 2 : 2
layers can occur interstratified in various combinations to produce a wide variety of
interstratified or mixed-layer physils. Two physils, sepiolite and palygorskite, have a
0000000 0000000

0000000
T
c sinp z
0

Serpentine-kaolin ( X - 0 ) Talc-pyrophyllite ( X - 0 ) Mica ( X - 1 0 ) and


brittle mica ( X - 2 0 )

ooocoGo
Tetrahedral cation

IM
I

0 : Octahedral cation
@ : Interlayer cation

I
0 = Exchangeable cation
14 4 - ;5 6 A 14 0 -;4 4A

0 I Oxygen
8 z Hydroxyl group
Q = Water molecule
@ :Oxygen + hydroxyl
Smectite ( X - 0 25 - 0.6) and Chlorite ( X is variable) ( i n projection)
Vermiculite ( X - 0 6 - 0.9

Fig. 1-1. [OlO] view of structures of major physil mineral groups. From Bailey. 1980. Copyright 1980
London Miner. Soc.

Table 1-1
Classification of phyllosilicates related to clay minerals (modified from Bailey. 1980h).

Layer Group Suh-group Species **


type (x = charge per
formula unit) *
1:l Serpent he-kaolin Serpentines Chrysolite. antigorite.
lizardite. amesite. herthierine
(x - 0) Kaolins Kaolinite. dickite. nacrite

2:1 Talc-pyrophyllik Taka Talc, willemseite


(x -
0) Pyrophyllitea Pyrophyllite
Smectite Saponites Saponite. hectorite, stevensite
(x - 0.2-0.6) Montmorillonites Montmorillonite, heidellite.
nontronite
Vermiculte Trioctahedral vermiculites Trioctahrdral vermiculite
(x - 0.6-0.9) Dioclahedral vermiculites Dioctahedral vermiculite
Mica Trioctahedral micas
(.w - 1.0) Dioctahedral micas Muscovite, paragonite. illite.
phengite. celadonite, glauconite
Brittle Mica Trioctahedral hrittle micas Clintonite. anandite
(x - 2.0) Dioctahedral hrittle micas Margarite
2:1 : 1 Chlorite Trioctahedral chlorites Clinochlore. chamosite. nimite
( x variable) Diwtahedral chlorites Donhassitr
Di.trioctahedra1 chlorites Cookeite. sudoite

2:l Sepiolite- Sepiolites Scpiolite. loughlinite


inverted palygorskite
rihhons ( x variable) Palygorskites Palygorskite (attapulgite)
* .Y refers to an O l o ( O H ) , formula unit for smectite, vermiculite. mica. and hrittle mica.
** Only a few examples are given.
3

chain or ribbon structure. These physils consist of 2:l layers with limited width in
one direction. Note that the terms plane, sheet, layer, and unit structure refer to
increasingly thicker parts of a layered arrangement.
Table 1-1 contains the classification of physils as approved by the AIPEA
Nomenclature Committee (Bailey, 1980b). The classification is based on the layer
type, layer charge, type of interlayer material, and type of octahedral sheet (dioc-
tahedral or trioctahedral). Further subdivision is based on chemical composition
and the geometry of superposition of individual layers and interlayers. In addition,
there is a wide variety of mixed-layered or interstratified physils which contain two
or more different layers. The various physils are discussed on the following pages.
For more details see Brindley and Brown (1980) and Weaver and Pollard (1973).
Later on I will report many data on the composition of physil suites. It should be
pointed out that though the various physils can be identified with considerable
accuracy, quantitative analysis is far from quantitative. Due to variations in com-
position, strucutre, interstratification, morphology, grain size, preparation tech-
niques, etc., it is virtually impossible to obtain quantitative analyses of natural
physi\ mixtures. If preparation techniques are standardized, precision can be excel-
lent and small, relative difference in abundance can be trusted.

01/04/19 (1) Structural Formula

The character of the various physils can best be visualized from their structural
formula. T o be sure everyone knows how these formulas are derived, I will show an
example. The technique is that of Ross and Hendricks (1945). The intent is to
interpret the chemical composition in terms of atomic ratios in the octahedral
sheets, the tetrahedral sheets, and the interlayer material. Two basic assumptions are
made. The cation charge is equal to the anion charge; the anion charge is known
from structural studies of coarser grained physils: for the 2:l physils the anions are
O,,[OH], = 22; for the 1:l physils the anions are O,,[OH], = 28; for chlorites the
anions are also O,,[OHI8 = 28. The other assumption is that the number of
tetrahedral cations is 4.0.
First the percent values must be converted to atomic proportions. The calcula-
tions are based on the distribution of single ions, and SO,, MgO, FeO, CaO,
containing single cations, are divided by the molecular weights; A1,0,, Fe,O,, K,O,
and Na,O are divided by one-half the molecular weight.
Next it is necessary to obtain the factor ( K ) by which the atomic proportions are
to be multiplied to produce the structural formula. The atomic proportion of each
cation is multiplied by valency of the cation, the values totaled, and divided into 22
(for 2:l physil):

K=
Zx4+Ax3+Bx3+Cx2+Dx2+E+F
4

Thus, a conversion factor, K, is obtained so that the total cation charge can be
made equal to the anion charge.
The next step is to determine the composition of the tetrahedral sheet: K[Z + Y]
= 4. The Z or Si value is known. The total cation population is 4. Y is the atomic
proportion of A1 necessary to bring the tetrahedral cations to a total of 4.
The formula can now be calculated by multiplying the atomic proportion values
by K:
Tetrahedral sheet = K Si K Y +
Octahedral sheet = K Fe + K Mg + K [A - Y]
Interlayer cations = K Ca + K Na + K K
The tetrahedral charge is K Y. The octahedral charge is 3[ K(A - Y ) + K Fe] +
2 K Mg - 6. The total layer charge should be very close to the total charge of the
interlayer cations.

Percent Factor Atomic


Proportions
SO2 50.10 60.06 0.834 Z
,4120, 25.80 51.00 0.506 A
Fe20, 2.97 79.92 0.037 B
FeO 0.00 71.85
MgO 2.95 40.32 0.073 C
CaO 0.39 56.08 0.007 D
Na,O 0.18 31.00 0.006 E
K2O 7.99 47.10 0.170 F
H2 0 7.36

1.518
0.1 11
0.146
0.014
0.006
0.170
~

5.301
K [ Z + Y ]= 4
4.1 50[0.834 + Y] = 4.00
Y = 0.964 - 0.834 = 0.1 30

Tetrahedral
K Si = 4.1 50 x 0.834 = 3.46
K Y = 4.150XO.130= 0.54
4.00
Octahedral
K Fe =4.150x0.037 = 0.15
K Mg = 4.150X0.073 = 0.30
K [ A - Y ] = 4.150X[O.506-0.130] ~ 1 . 5 6
xm
5

Interla.ver
K C a = 4 . 1 5 0 ~ 0 . 0 0 7 = 0 . 0 (3 ~ 2 )
K Na= 4.15OX0.006= 0.02
K K = 4.15OxO.170= 0.71
0.79
IAl i.s6FeO’.:s Mg0.30lISi 3.46A10.54IOiO(OH)2 . Ca0.03NaO.02K 0.70
Octahedral Tetrahedral Anions lnterlayer
Tetrahedral Charge K Y = 0.54
Octahedral Charge 3(1.56 +0.15)+ 2(0.30) - 6 = 0.21

01/04/19 (1) NOMENCLATURE

01/04/19 (1) Phyllosilicates and Rocks

Despite the relative abundance of clay-rich rocks in the geologic section, they
have never been satisfactorily classified. Early attempts at classification were
hindered by the lack of knowledge of their mineralogy. Although the mineralogy of
fine-grained sediments is now fairly well understood, there is still relatively little
information on the texture of these rocks.
The term shale and other terms for fine-grained rocks have been used for many
years and have been continually redefined. For a review of the history of the word
clay, see Mackenzie (1962), and for shale, see Tourtelot (1960). Most definitions
predate the development of modern instrumentation and are based on field observa-
tions and intuition; the diversity of definitions further complicates matters.
The main nomenclature problem is the dual meaning of the term clay: size and
mineralogy. Against my better judgment I decided it was necessary to introduce a
new word, physil, an abbreviated form of phyllosilicate. The term physil is applied
to all sheet silicate minerals and has no size connotation. A simple classification is
proposed that is based on the percentage of physils and grain size.
As a textural term, petrologists consider clay to be any material finer than
mm (3.9 pm) (Wentworth, 1922). Others have placed the upper boundary at 2 p m
(Atterberg, 1905), at 5 p m (U.S. Bureau Soils), and at 20 pm (Correns, 1969). As a
rock term a clay has been defined as a plastic earth composed of hydrous aluminum
silicates and of fine grain size (Grim, 1968; Pettijohn, 1975). The term has both a
textural and mineralogical definition. The mineralogical connotation should be
removed. “Classification of clays and silts and their indurated equivalents on almost
any basis gives little satisfaction to other than the classifier and often not even to
him” (Twenhofel, 1950).
Clay minerals are considered to be finely crystalline hydrous aluminum (Mg, Fe)
layer silicates. The term clay minerals and, to some extent, the word clay has been
used as a synonym for the term phyllosilicates but with the added connotation of
fine grain size. No other mineral group is defined on the basis of both size and
6

mineralogy. The “clay minerals” should be referred to as phyllosilicates and, i f


needed, a modifying size term should be added.
Much confusion is caused by applying the terminology of unindurated sediments
(clays, muds, etc.) to indurated and diagenetically altered sediments. Thus, the term
argillaceous implies clay. Clay, in turn, contains phyllosilicates of clay size. How-
ever, in consolidated rocks many of the phyllosilicates are silt-sized. Actually.
argillaceous might be more accurately defined as an adjective describing a rock
which contains an appreciable content of phyllosilicates of clay or silt size. O n the
other hand, geologists often prefer to use ambiguous words.
There is no common term that is used to describe a rock composed predomi-
nantly of sheet silicate minerals and which has no connotation of grain size. The
words claystone, mudstone, lutite, and pelite are defined on the basis of grain size
rather than mineralogy, though the rocks are assumed to have a high phyllosilicate
content. In turn, the phyllosilicates are assumed to be clay-sized. In practice the
identification is usually based on the presence of phyllosilicates rather than o n size.
In many instances the phyllosilicates are silt-sized. Thus, many claystones, lutites.
etc. are in reality siltstones. Some of the more widely used definitions of fine-grained
rocks are summarized in Table 1-2.
Petrologists have tended to use textural data from unconsolidated sediments to
classify the clay-rich rocks. Among other things, they assume that the grain size of
unconsolidated material is the same as that of buried sediments and ignore the
diagenetic processes which change the mineralogical and textural properties of
phyllosilicate-rich rocks.
The definition of shale includes the structural term fissile. Fissility is the
tendency of a rock to split along relatively smooth surfaces parallel to the bedding
(Pettijohn, 1975). Fissility is caused by laminations and the parallel orientation of
the sheet minerals and organic material (Ingram, 1953; Ciipson, 1965; O’Brien,
1970; Pettijohn, 1975; Spears, 1976; and many others). Fissility is decreased when
calcareous and siliceous materials are present. As the amount of cement increases
and as orientation becomes more random the size of the fracture blocks increases.
Fissility, or mineral orientation, can be initiated by depositional. diagenetic, or
metamorphic processes. Grim et al. (1957) note that in shales with a high content of
phyllosilicates, the phyllosilicates were > 5 p m (silt) or larger before fissility could
develop. Alling (1945), Ingram (1953), and McKee and Weir (1953) have proposed
classifications based on the splitting property of layered rocks.
As noted by Ingram (1953) fissility is a derived property and, in addition to being
a function of particle orientation and bedding, is determined by weathering.
temperature, and water content. All fine-grained rocks with oriented sheet minerals
are massive when under overburden pressure. As the overburden is removed or a
sample is removed from beneath the overburden, exfoliation occurs in some rocks
and subparallel cleavage develops. Other rocks require considerable weathering
(water penetration?) before cleavage develops. Rocks containing swelling phyllosili-
cates commonly d o not develop fissility until they have lost an appreciable amount
of their water by dehydration. There are no shales in the subsurface, only potential
shales.
7

Size Distribution and Composition

Shallow buried, recent, and ancient sediments with a high content of phyllosili-
cates commonly contain 50 to 80% clay and 20 to 40% silt (Table 1-3). Phyllosili-
cates are present in approximately the same amounts as the clay-sized material.

Table 1-2
Definitions of Some Phyllosilicate-Rich Rocks.
Term Definition Reference
Claystone Indurated clays. Retain considerable coherence on
being wetted after drying. Twenhofel. 1950
Weakly indurated, composed predominantly o f
"clay-sized" particles ( < 10 m). Flawn. 1953
Massive rock in which clay predominates. Ingram, 1953
Contains two-thirds clay. Folk. 1974
Contains more than 7 5 8 clay. Picard, 1971
Indurated clay. Pettijohn. 1975

Mudstone Partly indurated argillaceous rock which slakes readily


to mud when repeatedly dried and wetted. Shrock. 1948
Clays and silts mingled with water form muds and
both claystones and siltstones have heen termed
mudstones. Twenhofel, 1950
Forty sedimentationists recommended the term mud
be dropped. (Why not drop mudstone?) Shepard, 1954
Rocks with subequal silt and clay. Folk. 1974
Blocky or massive claystones. Pettijohn, 1975

Mud rock Massive claystone or siltstone. Contains at least


50% silt and clay. Ingram. 1953
Terrigenous rocks that contain more than 50% silt
and/or clay. Folk. 1974

Shale Claystone and siltstone with cleavage parallel


to bedding. Twenhofel, 1950
Fine-grained rock containing 50 to loo%,clay-size
particles with clay minerals constituting at least
25% of the total rock volume. Picard. 1953
More indurated than claystone. Flawn, 1953
Fissile claystone, siltstone, and mudstone. Ingram, 1953
Fissile mudstone. Pettijohn. 1975
Fissile mudrock. Folk, 1974

Argillite Siltstone or shale with a high degree of induration. Twenhofel. 1950


Indurated argillaceous rock without visible parting.
cleavage. or foliation.
Less than half the micaceous paste and clay
minerals have been reconstituted. Flawn. 1953
Massive shale. Grim, 1968
Metamorphosed mudstone or shale with no cleavage. Pettijohn. 1975
Claystone. Millot. 1970
8

Table 1-3
Fine-Grained Sediments.
Location R Clay % Silt R Sand Reference
~~

Mississippi River 40 Potter et al., 1975


Mississippi Delta Swamps 60 H o and Coleman. 1969
Mississippi Delta Fringe 50-70 30-50 Shepard, 1950
Offshore Texas > 50 Shepard and Moore, 1955
Upper Cretaceous and 58 41 1 Andrews and Overshine, 1974
Eocene. Indian Ocean
Cretaceous. off 79 20 1 Thayer et al.. 1974
West Coast Australia
Late Tertiary. Ross Sea 50 35 15 Barrett, 1975
Pliocene- Pleistocene, 57 42 1 Bode, 1973
N W U.S. and Canada
Lower Cretaceous, 70 30 Boyce. 1972
Blake-Bahama Outer Ridge
Late Tertiary. Lower 70 30 Boyce, 1972
Continental Rise. Virginia

Marine muds in cold climates tend to contain more feldspar and fewer phyllosili-
cates (Barrett, 1975; Bode, 1973) than those deposited in warm regions.
Many fine-grained rocks that have been relatively deeply buried contain 10 to
40% clay size material but the phyllosilicate content is in the range of 60 to 80%.
However, some Paleozoic sediments have both the clay content and the phyllosili-
cate content in the 70 to 80%range. As mineralogy does not have any fixed relation
to grain size, particularly in older rocks, the two parameters must be considered
independently.
Pettijohn (1975) estimated, on the basis of very limited chemical data, that the
average shale was composed of two parts silt and one part clay. Thus, most shales
should be siltstones. Two-thirds of the twelve samples Pettijohn used to calculate
the composition of the average shale were glacial and recent clays (Grout, 1925). All
samples were from Minnesota. I t is unlikely that the group of samples are “typical”.
The “two parts silt-one part clay” shale has the composition shown in Table 1-4
under the Pettijohn column.
Also shown in Table 1-4 are two other estimates of the mineral composition of
the average shale. Yaalon’s (1962) data were based on normative calculations using
the chemical composition of the average shale compiled by Clark (1924). The data
of Shaw and Weaver (1965) were based on direct mineral analyses (x-ray diffrac-
tion) of 400 North American shale samples. Recent analyses confirm the values of
Shaw and Weaver.
Blatt and Schultz (1976) used a chemical dissolution technique to determine the
amount of quartz in 16 mudrocks. The amount of quartz plus chert averaged 28%,
of which 96% was quartz. The sand-silt-clay percentages for the size distribution of
quartz plus chert was 12, 73, and 15%, respectively.
Scotford (1965) made detailed analyses of 158 Ordovician shale samples from
Ohio, Kentucky, and Indiana. The rocks are described as being massive when dry
9

Table 1-4
Average shale.

Pettijohn Yaalon Shaw and Weaver


(1975) ( 1962) (1965)
Quartz 29 20 31
Feldspar 11 8 4
Phyllosilicates 43 59 61
iron Oxides 5 3 0.5
Carbonates 9 7 4
Other Minerals 2 3 2
Organic Matter 0.3 - 1

but show good fissility when wet. This further demonstrates the limits of fissility as
a basic criterion for classification. The mean size distribution of the shale samples
was 3% sand, 59% silt, and 38% clay.
The quartz and feldspar comprised approximately one-quarter of the shale by
weight and the phyllosilicates 61%.The carbonate content was 14%. The rocks were
texturally siltstones, yet contained 61 % clay minerals (phyllosilicates). The silt
fraction of these shales contained 28%quartz, 55% phyllosilicates, and 17%carbonate
minerals.
Cambrian Conasauga shale samples (25) from northwest Georgia (Weaver and
Associates, 1984) have only 8 to 30% of the material finer than 2 pm; the amount
decreases with increasing depth of burial. Phyllosilicates are abundant in the coarser
fraction and increase with increasing depth of burial.
In contrast, mineral and size analyses of an Upper Ordovician Sylvan shale
sample from Oklahoma (Hower et al., 1963) showed it contained 61% less than 2
pm material. Quartz comprised 23.8% of the total sample; phyllosilicates, largely
illite, comprised the other 76.2%.The > 2 pm fraction contained 46.0% quartz and
53.2% phyllosilicates. The < 2 pm fractions contained 9% quartz and 91% clay
minerals. I t should be noted that even though 61% of the material was < 2 pm,
more than half of the silt-sized material was phyllosilicates.
The grain size of phyllosilicates in shales is in part controlled by source (detrital
effect) but it is also determined by diagenetic mineral growth, which is determined
by depth of burial and temperature.
X-ray and size analyses of Ordovician K-bentonite beds indicate they contain
more than 95% phyllosilicates but only 30 to 70% of the material is clay size
(Weaver, 1953). The shallow buried montmorillonitic Upper Cretaceous Pierre shale
has an average of 70% of the minerals less than 4 pm (Tourtelot, 1962).
Size analyses of montmorillonitic Oligocene-Miocene clays and claystones
(1,250 m to 5,500 m depth) from the Gulf Coast (Hower et al., 1975) showed that 50
to 71% of the material was finer than 2 pm. Eighteen Miocene samples (390 m to
3,588 m depth) from the Gulf Coast (Cooke, 1977) contained from 55 to 78% less
than 4 pm material and averaged approximately 65%.
10

I t is virtually impossible to analyze a shale and obtain any meaningful infornia-


tion about the size of the phyllosilicate particles at the time of deposition. Phyllo-
silicates are deposited as isolated flakes, books, floccules, pellets, clasts. and rock
fragments.
The majority of the phyllosilicates in unconsolidated clays, recent marine muds,
and phyllosilicate-rich rocks that have not been deeply buried can be dispersed into
particles finer than 4 p m (usually 2 pm). With burial, temperature and pressure
increase and the average grain size of the phyllosilicate suite systematically increases
and many of the phyllosilicate plates and aggregates increase in size and eventually
become silt-sized. Unconsolidated clays commonly become siltstones at burial
temperatures higher than 150 to 200”.
At burial temperatures as low as 50°C (commonly 2,000 to 4,000 m depth)
montmorillonite starts to convert to mixed-layer illite-montmorillonite and increase
in size. The increase in size is probably continuous with increasing temperature. up
to and through the metamorphic stage, as the montmorillonite is converted to illite,
although this has not been definitely established. Weaver (1960) and Kubler (1964)
have demonstrated that during advanced diagenesis and early metamorphism the
“sharpness” of the 10 A illite peak increases and the width decreases. I t seems likely
that flake-size also increases.
Hower et al. (1975) reported that during burial diagenesis of smectite, chlorite
increased abruptly from zero to 10% at a depth where the temperature was
approximately 70°C. The increase occurred in both the clay and silt size fractions.
There was no further increase with increased depth and temperature. In the
Cambrian Conasauga Shale abundant silt-sized chlorite developed during the late
stages of diagenesis ( - 250°C). In general it appears that during diagenesis chlorite
may become coarser than illite.
Though this temperature-size relation is probably most pronounced for
montmorillonite-rich sediments, a similar textural change should occur in all sedi-
ments containing phyllosilicates.

New Terms

There is need for a term equivalent to feldspar. carbonate. and zeolite that
includes the whole family of phyllosilicates and has no connotation of grain size. I
am proposing that the term physil be used as an abbreviated form of the term
phyllosilicate. Physils would include all hydrous aluminum-magnesium-iron silicates
with a layer structure and would include the palygorskites and sepiolites which can
be considered to have a layer structure with limited width in one direction. The term
has no size implications and includes both the “clay minerals” and the coarser micas
and chlorites. The preceding discussion has shown that many of the physils in
fine-grained rocks are silt-sized or larger, The term clay should be used only in the
textural sense, preferably to describe the material finer than & mm (4 pm). The
term claystone should have the same connotation as sandstone strictly size. The
~

term shale will continue to be used, but it is recommended that it be used only as a
11

Table 1-5
Classification o f Fine-Grained Rocks Containing Phyllosdicates.
Texture Composition
Unindurated Indurated
> 50% < 50% > 50% < 50%
Physils Physils Physils Physils
~

50% silt physil physilitic physil physilitic


I 1
( 230- 10 mm) silt silt siltstone siltstone

50% clay physil physilitic physil physilitic


( <A, mm) clay clay claystone c 1ay st on e

field term to describe fissile siltstones and claystones. Most such rocks can be more
accurately called fissile physilites.
The terms physil clay, physil silt, physil silty clay, etc. can be used to refer to
unconsolidated material containing more than 50% physils. Physilitic (calcareous
physilitic clay, etc.) can be used to describe unconsolidated sediments containing
less than 50% physils (Table 1-5).
Consolidated or lithified rocks containing more than 50% physils would be called
physilites. Physilites would include most rocks classed as shales, claystones, mud-
stones, and siltstones. Slates and perhaps schists could be referred to as meta-physi-
lites. If the physil content is larger than 50%, the rock can be described as physil
claystone, physil siltstone, calcareous physil claystone, etc. If the physil content is
less than 50% the terms physilitic claystone, physilitic siltstone, physilitic sandstone
would be applicable. The terminology is compatible with that used for the other
rock types, e g , graywacke sandstone, arkosic siltstone, oolite sandstone.
Sedimentary rock names should not have a structural connotation. Structural
terms would serve as prefixes. Thus, rocks should be described as fissile physilites.
fissile physil claystones, laminated physil claystones, massive physil siltstones,
massive physilitic siltstones, etc. The two- and three-word descriptions are not as
succinct as single words, but the ambiguity is considerably reduced.
This Page Intentionally Left Blank
13

Chapter II

STRUCTURE AND COMPOSITION

01/04/19 (1) KAOLIN

The name kaolin was derived from the Chinese kauling”, meaning “high ridge”,

the name of a hill near Jauchau Fu. China, where the material was mined centuries
ago (Grim. 1968). The kaolin minerals include four distinct species kaolinite,
dickite, nacrite, and halloysite (Ross and Kerr, 1931: Ross and Kerr, 1934).

01/04/19 (1) Kaolinite, Dickite, Nacrite

01/04/19 (1) Structure utid s-ruv


Kaolinite is a 1:l physil composed of a Si tetrahedral sheet and a Al octahedral
sheet (dioctahedral) combined so that the oxygens at the apex of the Si tetrahedrons
extend into and are part of the octahedral sheet. The other anions necessary to
complete the octahedral coordination of the A1 in the octahedral sheet are OH (Fig.
2-1). Adjacent 1:l layers are held together by hydrogen bonds extending from one
plane of OH ions forming one side of the octahedral layer to the basal oxygens of
the tetrahedrons of the adjacent layer. The ideal charge distribution in the layers is
as follows:

Ions Charge Distance A


60” 12- 0.0
4si4 + 16 + 0.60
’.
4 0 2 2(OH) 10- 2.19
4AI“ 12+ 3.27
6(OH)- 6- 4.37
Interlayer Distnnce 3.0

More so than the other physils, the charges within the layers are nearly balanced
and the unsatisified layer charge is near zero; however. minor substitutions produce
a small layer charge.
The tetrahedral sheet of the kaolin minerals is larger than the octahedral sheet
and considerable lattice distortion is required for the two sheets to become more
nearly equal in width (Brindley, 1951: Zvyagin, 1960; Newham, 1961; Radoslovich,
1963; Bailey, 1980). The kaolin minerals are dioctahedral, having only two of these
octahedral positions filled with Al”: this causes further distortion. Mutual repul-
14

1
I
I
I
I

J
I?
-Y

tetrahedral octahedral

0 Oxygen 0 Hydroxyl Sdicon 0 Aluminum

Fig. 2-1. Upper: Diagrammatic sketch of the 1:l kaolinite layer (After Gruner. 1932). From Grim, 196X.
Clay Mineralogy. Lower: The tetrahedral and octahedral sheets of kaolinite. From Brindles. 1961.
Copyright 1961 London Miner. Soc.

sion between adjacent Al'+ cations causes movement of the coordinating anions
towards one another along the shared octahedral edges. This enlarges the unshared
octahedral edges around the vacant site (Fig. 2-1). The result is two relatively small
filled positions, a larger vacant octahedral position, and a stretching and thinning o f
the octahedral sheet (Fig. 2-1). Much of the structural data for kaolin minerals is
based on analyses of the relatively coarse grained and well crystallized dickite.
15

The tetrahedra adjust to the octahedral sheet by a tetrahedral rotation of


somewhere between 7" and 20" (Suitch and Young, 1983), which shortens the
tetrahedral layer and produces a ditrigonal, rather than hexagonal, pattern of basal
oxygen ions (Fig. 2-1). Further adjustment is required to accommodate the dif-
ference in size of the octahedra. The enlarged edges around the vacant octahedral
site are larger than the distance between the tetrhedral apices of oxygens. This misfit
is corrected by tilting the tetrahedra so that the apices of adjcacent tetrahedra point
A
slightly away from each other. Thus, there is a 0.1 to 0.2 elevation of the bridging
oxygens between each pair of tilted tetrahedra. The inner OH bonds are also
affected. One 0 - H bond points into the octahedral vacant position and the other
away from the octahedral sheet and toward the unoccupied center of an oxygen
triangle formed by the two apical oxygens and shared basal oxygen of two adjacent
SiO, tetrahedra. All six of the inner surface hydrogen atoms appear to be nearly
equally involved in the hydrogen bonding between kaolinite layers (Suitch and
Young, 1983).
Kaolinite analyses are not as accurate but indicate comparable features to
dickite: shortened shared octahedral edges and counter-rotations of octahedral
triads (+3", -5" vs. +6.5", -4"); 0 - H - 0 bonds of about 3 A; shortened AI-OH
surface bonds; one OH raised out of the surface. Some contrast may be significant:
the c-axis is larger and the b-axis is slightly smaller than dickite; the surface oxygen
is elevated from the layer, not depressed into it. The a-axis displacement is slightly
greater than - a/3. Blount et al. (1969) found that in nacrite adjacent tetrahedra are
rotated 7.3" in opposite directions, bringing the basal oxygens closer to the AI
cations in the same layer and the surface hydroxyls of the layer below. The upper
and lower anion triads in each Al-octahedron are rotated by 5.4" and 7.0" in
opposite directions as a result of shared-edge shortening.
Table 2-1 lists the x-ray powder data for kaolinite. Typical patterns are shown in
Figure 2-2. Though the calculated 001 value is 7.14 A, due to layer stacking
A A
imperfections (folds, bubbles, 10 and 14 interlayers), the value reaches 7.20 A
and even higher (Brindley, 1980.)
Structures resulting from different stacking sequences of layers are termed
polytypes. Polymorphism (Brindley, 1961; Bailey, 1963) is most easily analyzed in
terms of two factors: (1) the direction and amount of the interlayer shift, and (2) the
location of the vacant octahedral site in successive layers. Kaolinite and dickite have
identical interlayer shifts and give x-ray powder patterns that are similar with
respect to most of the stronger reflections. The differing location of the vacant
octahedral site in the two structures governs the symmetry and the Z-axis (c-axis)
periodicity of each mineral and accounts for the observed differences in the powder
reflections of medium to weak intensity. Nacrite has a sequence of interlayer shifts
that is entirely different from that in kaolinite and dickite, and its powder pattern is
also markedly different.
Successive kaolin layers are held together by long (approximately 3.0 A) hydro-
gen bonds formed by the O H - 0 pairing between octahedral hydroxyl ions at the top
of the layer and tetrahedral oxygens at the base of the overlying layer. The layers are
stacked together so that these O H - 0 distances are approximately equal. Hydrogen
16

TABLE 2-1
X-ray powder data for kaolinite (Bailey. 198Oa)
A. Kaolinite (Indexing modified from Brindley. 1961)
d (ohs.) I d (calc.) hkl d (ohs.) I d (talc.)
-
7.16 10 t 7.14 001 1472 243
4.46 4 4.47 020 1.467 2 1.469 061
4.36 5 4.36 I io 1.467 332
4.34 110 1.455 332
4.18 5 4. I72 iii 1.453 155
1.452 48
4 13 3 4.120 1T i 1.452 330
3.845 4 3.842 027 1.450 Oh 1
3.741 2 3.739 021 1.429 4 1.428 005
3.573 10 t 3.571 002 1.405 242
1.403 2
3.417 IT] 1.402 205
3 372 4 3.372 111 1.391 335
1.390 2
3 144 7 3.147 112 1390 nfi?
3.097 3 3.096 1 TI 1.375 044
2.832 022 1.371 2 1373 373
2.753 3 2.750 022 1.369 371
2.562 130 1.338 4 1.339 135
2.558 6
2.561 201 1.307 135
1 305 6B
2.526 4 2 531 I 3i 1.302 204
2.491 8
2.497 137 1.292 2 1.291 26i
2.488 200 1285 2hl
1.282 5
2.379 6 2.381 003 1282 401
2.340 202 1.264 3 1 266 267
2.338 9
2.335 I31 I .248 262
1.246 1
2.288 8 2.292 131 1.244 400
2.247 2 2.249 132 1.237 I 77
2.186 1 2.183 20 1 1.235 3 1.236 170
2.131 3 2.128 023 1.233 422
2.061 2 2.060 222 1.220 045
1.992 203 1.219 3 50
1.989 6
1.985 112 1.217 I 1.217 351
1.939 4 1.938 132 1215 3T3
1.896 3 1,895 137 1.214 ii6
1.R69 2 1.870 042 1.200 3 1.200 420
4
1.843 133 1.191 427
1839 1.190 3
1.834 202 1.190 006
1.809 114 1.170 404
1.809 2 1.168 2
1.805 223 1.167 3 s
1.781 4 1.785 004 1.124 1 1.125 264
1.707 2 1.708 222 1 097 26Z
1.094 1
1.685 2 1.689 241 1 092 402
1.665 204 1.085 405
1.082 2
1.662 7 1.662 133 1.082 263
1.662 151 1057 1 1.058 263
1 620 I51 1.049 336
1.619 6 1.049 2
1.619 133 1.047 065
1.584 4 1.587 134 1.042 2oi
1.039 2
1.548 224 1039 136
1 .542
1.543 134 1.038 046
5%
1.541 313 1.021 2 1 020 007
1.536 20 3 1019 136
1.013 2
1.490 060 1013 334
1.489 8 1.4Rb 33i
1.485 3Ji

bond arrangements, although differing in detail, can be formed by several different


positions (shifting along a- and b-axes and/or rotation) of the layers relative to one
another.
There are six different ways that the oxygens of the next kaolin layer can be
paired with the hydroxyls of the reference cell. For each of the six possibilities the
second kaolin layer can be rotated six times at 60" making a total of 36 different
superposition patterns for the kaolin layers. First, those stackings are eliminated
which have the greatest amount of cation-cation superposition (repulsion) in con-
secutive layers. This leaves two one-layer and twelve two-layer cells. The two
single-layer cells are those of kaolinite and its mirror image. When allowances are
made for the directed OH bond and the buckling of the oxygen layer, only two of
the two-layer unit cells are usable. These two two-layer structures are those of
dickite and nacrite. Neither of the kaolinite structures satisfies the pucker criterion;
this may explain why kaolinite crystals are seldom as large as those of dickite and
nacrite.
Kaolinite and dickite have identical layer sequences in which each layer is shifted
by -1/3a, relative to the layer below (Fig. 3). The two structures differ only in
regard to the distribution of the vacant cation site in successive octahedral sheets. If
the two minerals were trioctahedral (no vacant cation positions in the octahedral
sheet) they would be identical and have a one-layer monoclinic (IM) structure.
The three possible octahedral sites, only two of which are filled in the kaolin
minerals, lie on a mirror plane and B and C lie on opposite sides of the mirror
plane. Choice of either B or C as the vacant site imposes triclinic symmetry on the
structure due to loss of the symmetry planes. In well-crystallized kaolinite each layer
is identical and has octahedral site C (or B) vacant. In dickite the vacant site
alternates between C and B in successive layers to create a two-layer structure (Fig.
2-3). The alternation of vacant sites in dickite tends to balance the stress distribu-
tion in the two layers so that the cell shape remains monoclinic. Thus dickite can be
considered as a regular alteration of right- and left-handed kaolinite layers.
In poorly crystallized kaolins it is possible that the vacancy does not always occur
in the same octahedral site in each layer but is a random interleaving of right- and
left-handed kaolinite crystals (random choice of C or B as the vacant site in
different layers).
Nacrite has a six-layer stacking sequence in which each layer is shifted by
-1/3a, rather than 1/3b (Blount et al., 1969). Alternate layers are rotated 180'.
The vacant octahedral site rotates + 60" between layers, creating a two-layer
periodicity along the inclined z-axis.
In well-crystallized kaolinite and dickite, displacement of one kaolin layer with
respect to the other is along the a-axis. Most kaolinites (Brindley, 1961) are also
shifted in the direction parallel to the b-axis. The effects of this shift along the
b-axis on x-ray reflections is used as a measure of "degree of crystallinity". Actually,
what is measured is the amount of stacking disorder.
The feature most commonly observed is that reflections with the k index (b-axis
reflections) not a multiple of 3 (k # 3n) tend to be weak or missing, while reflections
with k = 3n tend to be largely unaffected.
The OH ions in the hydroxyl layer lie in lines parallel to the b-axis and at an
interval of b/3. Therefore, the structural layers can be displaced parallel to b by
nb/3, without altering the OH-0 bonds between the adjacent layers. Such displace-
ments can be expected to occur rather easily because no marked energy changes are
involved.
Layer displacements of b/3 cause phase changes (in the x-ray reflections) of
k x 12O", so that if k is not a multiple of 3, phase changes of Oo, 120°, and 240"
18

KAOLINITE
(Keokuk, 10~0)

t
KAOLINITE
(Georgia)

KAOLlNlTE
(Georgiol

KAOLlNlTE
(Pugu, Tongonyika)

I
I
- -
HALLOYSITE
(Indiana, llO'Cl

1 1 1 I I 1
2o Degrees, 28 O' 60
A
Fig. 2-2. X-ray diffractometer patters (Cu K a radiation) of four kaolinites and one halloysite (7 form).
Crystallinity decreases from top to bottom. From Brindley, 1980. Copyright 1980 London Miner. Soc.
19

X n X n

0 Al ~VuCunCy
Fig. 2-3. Normal projection onto (001) of the octahedral portions of three layers (labeled 1. 2. 3) of the
kaolinite and dickite structures, showing the distribution of cations and vacancies over the A. B, and C
octahedral sites. In both structures each layer is shifted by -1/3a, relative to the layer below. T h e
projected Z-axis vector is shown as a solid-line arrow. After Bailey. 1963, Amer. Miner.. 48. 1196-1209.
From Grim, 1968. Copyright 1963 Min. Soc. Anier.

occur randomly and the corresponding hkl reflections are cut out. When k is a
multiple of 3, the phase changes are multiples of 360" and the corresponding hkl
reflections are unaffected. When this is not the case, two-dimensional coherence is
still possible; and in place of sharp hkl reflections, one may obtain hk bonds of
scattering.
On the basis of calculated theoretical profiles of (02,ll) band Plancon and
Tchoubar (1977) concluded that the apparent stacking disorder is due to the
displacement from one layer to the other, or from one domain to another, of the
vacant octahedral position.
There appear to be all gradations between well-crystallized kaolinite and that of
complete randomness in the b direction, and in the distribution of octahedral Al.
Fig. 2-4, after Murray and Lyons (1956), and Fig. 2-2 demonstrates the variation in
the character of x-ray patterns of kaolinites of different crystallinities. Johns and
Murray (1959) recommended that the relative intensities of the (021) and (060)
reflections be used as a measure of kaolinite crystallinity. Values range from 0.0 to
1.0 with the more crystalline kaolinites having higher values. Hinckley (1963)
suggested a measure based on the degree of resolution of the l i 0 and l l i reflections
(Fig. 2-5).
Fire clays, ball clays, and flint clays are kaolinite-rich clays, usually of the b-axis
disordered variety, which contain a relatively high impurity content. Illite,
20

71 5 4 3 2.5 2 1.5
1 , I , , , , J

lO.0Ol 10.002
0.060
A Ill I, I I I 1 I 1 , l I I I I 1

L -
7A 5 4 3 2.5 2

Fig. 2-4. X-ray powder diagrams of samples of kaolinite arranged in order of crystallinity. From Murray
and Lyons, 1956. Copyright 1956 Natl. Acad. Sci.

montmorillonite, diaspore, boehmite, quartz, and organic material are the minerals
usually associated with these deposits. Few, if any, of the kaolin minerals in these
clays have been concentrated enough to afford meaningful chemical data.
I .5
21

iio

CRYSTALLINITY

WELL CRYSTALLIZED
WAOLlNlTE

I I I I I I
180 200 220 24 26O 200 DEGREES 28
Fig. 2-5. Crystallinity index for kaolinite. From Hinckley, 1963. Reprinted with permission from Clays
Clay Miner., 13. Copyright 1963 Pergamon Journals, Ltd.

Approximately a dozen examples of randomly interstratified kaolinite/sniectitc


(K/S) have been described (Sudo and Hayashi. 1956; Schultz et NI.,1971; Wiewibra.
1972; Calvert and Pevear. 1983: and others). Quantification curves for the de-
termination of the ratio K/S from X-ray patterns have been calculated by Tomita
and Takahashi (1986). Generally K/S formes by the weathering of montmorillonite
and is probably more abundant than generally realized.
High-resolution electron microscopy (Lee rt NI.. 1975) and differential heats of
K-Ca exchange (Talibudeen and Goulding, 1983) indicate that kaolinite commonly
22

contains interstratified layers and islands of mica, vermiculite, and smectite. The
amount of 2:l material is too small, commonly less than a few percent, to be
detected by x-ray analysis. It is believed to be a residue of incomplete weathering of
mica to kaolinite. The 2:l inclusions presumably account for much of the K and
Mg, and high C.E.C. values found for some kaolinites. Thicker mica layers inter-
leaved with kaolinite plates can often be observed with the optical microscope.
Kaolinite mostly occurs as flakes or plates ranging in shape from near hexagonal
to ragged-edged anhedral. There is apparently little relation of shape to crystallinity.
In many kaolinite-rich deposits books or stacks built of crystal plates are relatively
abundant (Fig. 8-18).
A wide variety of organic complexes (formamide, urea, acetate, glycerol, benzi-
dine, etc.) can penetrate between the layers of kaolin and expand the layers from 7
A to 10 to 15 A.Halloysite is more easily expanded than kaolinite and various
techniques have been devised, using these organic compounds, to differentiate
between these two minerals and to distinguish between the other various types of
kaolin (MacEwan and Wilson, 1980; Theng et ul., 1984).
21/04/19 (1) Chemistry
Kaolinite is by far the most abundant species of the kaolin group. Although
hundreds of chemical analyses of this clay have been made, there is little known for
certain about the exact composition of most samples. The ideal composition for
kaolinite Si,AI,O,,(OH), is 46.54% SiO,, 39.5% A1,0,, 13.96% HzO; however, in
nature, this exact composition is seldom, if ever, found (Table 2-2).
Fe,O,, TiO,, MgO, and CaO are nearly always present in kaolinite samples, and
K ?O and Na ,O are usually present. Most samples either have excess SiO, or A1 .O,.

Table 2-2
Chemical Composition of Kaolinite

1 2 3 4 5 6 7
SiOz 44.15 41.03 44.62 43.84 41.76 46.40 45.1
A1 2 0 3 38.99 38.40 35.43 38.85 37.86 39.52 37.x
HzO+ 12.96 14.05 12.84 13.54 13.28 13.90 13.4
Total 96.10 93.48 92.89 96.23 92.90 99.82 96.3

Fez03 1.06 0.59 0.93 0.46 0.99 0.09 0.45


TiOz 0.62 1.60 0.78 1.03 2.01 none 2.1
MgO 0.14 0.22 0.45 0.36 0.28 0.15 0.09
others 1.59 2.46 3.03 0.77 2.53 0.33 1.01
1.93 1.82 2.14 1.92 1.88 2.00 2.03
I . Cornwall. England
2. Murfreesboro, Tennessee U.S.A.
3. Zettlitz, Czechoslovakia
4. Macon. Georgia U.S.A.
5. New Jersey, U.S.A.
6. Keller ct d.(1966). Keokuk geode, Iowa, U.S.A.
7. Warde (1950). South Africa (Flint Clay)
* 1-5 from Van der Marel. H.W. (1958).
23

Mineral impurities such as quartz, anatase, rutile, pyrite, limonite, feldspar, mica,
montmorillonite, and various iron and titanium hydroxides are commonly present in
addition to a number of other minerals. Si and Al, in the form of hydroxides,
apparently can occur as coatings on the kaolinite layers. Although many of these
impurities are usually identified, seldom is the analysis sufficiently quantitative to
determine if all the deviation from the ideal composition is due to the impurities.
Ross and Kerr (1931) reported on a series of kaolinites in which the SiO,:R,O,
mole ratio varies from 294:lOO to 185:lOO; the theoretical value is 200:lOO. EDX
analyses (Weaver, 1976) and electron microprobe analyses (Jepson and Rowse,
1975) indicate the Si/AI ratio of kaolinite varies from particle to particle. The latter
authors report values ranging from 0.992 to 1.082. Both dickite and nacrite have
ratios near 200:lOO and contain relatively few impurities.
Alumina (wet chemical analysis) is commonly present in excess in Georgia
kaolinites, usually in amounts of 1 to 2 percent. Although some of this may be
substituting for Si in the tetrahedral sheet, it is likely that much of it is adsorbed on
the face and edge of the kaolinite flakes as A1 hydroxide, perhaps, in part, acting as
a cement.
Chemical dissolution techniques indicate kaolinite from Cornwall contains 3.1 to
4.9 percent of easily soluble SiO, and 1.5 to 5.9 percent of easily soluble A1,0,
(Follett et al., 1965). Most of this material is presumably present as amorphous
material. Values for Georgia kaolinite are 2.1 and 0.95 percent. Keller et al. (1966)
found a well-crystallized, nearly pure kaolinite in the protected voids of geodes in
the Mississippian Warsaw Formation of Iowa. The SiO,/AI,O mole ratio is 2.00;
no TiO, and only 0.09% Fe,O, is present. The dehydroxylation temperature of this
well-crystallized kaolinite is approximately 100" higher than that of " typical"
kaolinite. A number of kaolinites were analysed by Van der Mare1 (1958) and such
parameters as crystallinity, surface area, and cation exchange capacity determined.
Though most of the measured parameters show good negative or positive correlation
with each other, they show little correlation with composition.
Most of the TiOz in kaolinite deposits is in the form of 0.1 pm anatase pellets
that can be dispersed and flocculated independent of the kaolinite. The anatase
contains significant amounts of Fe and Mg, although most of the Fe and Mg is
present in other forms (Weaver, 1976). Leaching experiments by Dolcater et ul.
(1970) indicate that on an average 15 percent (range 0 to 30 percent) of the Ti in
seven samples (six from Georgia) was in the kaolinite structure.
Much of the iron (1 to 2 percent) is present as iron oxides and pyrite: however,
some is complexed with the TiO, and some occurs in the micas and montmoril-
lonites commonly present. After removing most of the mineral impurities and
leaching, the Georgia kaolinites retain about 0.3% Fe,O,, suggesting that this
amount is probably present in the lattice. Mossbauer analyses by Malden and
Meads (1967) and Fysh et al. (1983) showed that Fe was present in the octahedral
sheet of the kaolinite layer. A typical formula for "cleaned" commercial kaolinite is
S i 4 . ~ 4 A l , . , * F e " , ~ ~ T i ~ ~ , , "(Georgia)
0 , ~ ( O H )(Rengasamy,
~ 1976). Most studies of the
Fe content of kaolinite have been conducted on commercial kaolinite: soil kaolinites
commonly have a higher Fe content. Analyses of 27 tropical ferruginous soil
24

samples from India indicated that as much as 6% Fe,O, is present i n the kaolinite
structure: Si,A13,2,Fe;:,:30,,,(0H),. Octahedral Fe” ranged from 0.1 1 to 0.X2. The
b-dimension (060) increased from about 8.92 A to 9.02 A as the Fe content
increased (Rengasamy et al., 1975).
Electron paramagnetic resonance (EPR) studies comfirm that structural Fe’ ’ is
present in the octahedral sheet of kaolinite and further indicate it is present in two
different sites, I iron and E iron (Angel and Hall, 1973; Meads and Malden. 1975;
Herbillon et a/., 1976; Mestdagh et a/., 1980). As total Fe,O, increases, the degree of
kaolinite crystallinity decreases (Hinckley Index); however, Mestdagh et al. ( 19x0)
showed that the decrease in crystallinity correlated with I iron and E iron had n o
effect on crystallinity. Total Fe,O, ranged from 0.06 to 2.32 (24 samples); I iron
ranged from 0 to 1.81. The difference in the location of the I and E sites has not
been established.
The cation exchange capacity of the kaolinite minerals is relatively low, but due
to the omnipresent impurities it is difficult to determine true values. Van der Marel
(1958) lists values ranging from 3.6 to 18.0 meq/lOOg, which is the generally
accepted range. He showed an excellent linear relation between C.E.C. and surface
area (other analyses fall on this same trend), which would occur if 2:l expandable
layers were present.
Part of the exchange capacity is believed to be due to broken bonds at the edges
of the flakes. This charge is reversible, being negative in a basic environment and
positive in an acid environment and zero at neutrality (Schofield and Samson.
1953). Sumner (1963) noted that iron oxides (and aluminum hydroxides) are present
on the surface of kaolinite and materially influence the measured C.E.C. Iron oxides
behave amphoterically. Thus, in a basic solution, iron oxide acts to increase the
C.E.C. of a kaolinite-iron oxide complex and in an acid solution tends to produce a
low value. He demonstrated that the negative charge o n kaolinite was constant
between p H 2.4 and 5.0, but above this increased with increasing pH. and con-
cluded, along with others, that this indicated kaolinite has a permanent negative
charge below pH 5 due to isomorphous replacement within the structure.
McBride (1976) concluded, on the basis of electron spin resonance studies, that
kaolinite had no edge exchange sites and that pure kaolinite should have a C.E.C. of
- 1 meq/100g which is due to minor isomorphous substitution. Schofield and
Sampson (1953) calculated that only one Al” need replace a Si4’ in 400 unit cells
to produce a C.E.C. of 2 meq/lOOg. I t seems likely that pure kaolin has an
extremely low C.E.C. and that values appreciably larger than one are due to the
presence of interstratified expanded layers or islands.

21/04/19 (1) Allophane, Imogolite, Halloysite

Allophane, imogolite and halloysite are treated as a group because they com-
monly occur together and are frequently genetically related. Allophane is a near-
amorphous Si-AI complex; imogolite is a paracrystalline Si-AI tubular mineral;
halloysite is a tubular or spherical kaolin.
25

fluffy

Fig. 2-6. Electron micrograph of allophane (round particles) and imogolite (fibers) from Andept soil. Bar
equals 500 A. Courtesy K. Wada.

21/04/19 (1) A llophane


Allophane has long been described as a fluffy amorphous Si-Al-H ,O precipitate.
Kitogawa (1971), Henmi and Wada (1976), and Wada (1979) showed that it consists
of 30 to 50 A (diameter) hollow spheres (Fig. 2-6) with walls composed of defect
kaolin or imogolite-like layers. The wall structure implies the existence of long-range
order in two dimensions, though X-ray diffraction effects do not occur because of
the form and fineness of the spheres.
Most allophanes have SiO2/A1,O3 mole ratios between 1 and 2. The A1 occurs in
both 4- and 6-fold coordination. The amount of A1 in 4-fold coordination increases
as the SiO,/Al,O, ratio increases (Ross and Kerr, 1934; Henmi and Wada, 1976).
The H,O content commonly ranges from 25 to 40%.Allophane is relatively “pure”.
Fe,O, and MgO are usually present in amounts less than 0.57%. Up to 2-3% CaO is
commonly present. The Ca may occur as an exchange cation.
The cation exchange sites occur inside the allophane spheres. C.E.C. values are
quite variable and depend on the type of exchangeable cation, concentration, and
pH. Wada (1977) reports values ranging from 20 to 45 meq/100g; Aomine and
Jackson (1959) suggested an average value of 100 meq/100g.

21/04/19 (1) Imogolite


Imogolite is a tubular paracrystalline material that is commonly associated with
allophane (Fig. 2-6). It has a long-range order that allows the formation of tubes a
micrometer long with an external diameter of about 20 A. Imogolite consists of a
26

Allophone I

Weaver and
Pollard, 1973
Fig. 2-7. X-ray diffraction patters of allophane and imogolite. From Wada. 1977. Reprinted by
permission of Soil Sci. Soc. Amer.. Inc.

single gibbsite sheet curved in the form of a tube, with orthosilicate groups attached
to the inside of the tube, each group replacing three OH groups around an empty
octahedral site. The long and flexible tubes combine to form a thread with
somewhat parallel but imperfect alignment of tubes (Wada and Yoshinaga, 1969;
Wada et al., 1970; Wada, 1981). Crystallinity is sufficiently well developed so that
broad x-ray diffraction peaks are observed at 12-20, 7.8-8.0, 5.5-5.6, 3.3, and 2.25 A.
X-ray patterns of imogolite and allophane are shown in Fig. 2-7.
The SiO,/AI20, ratio is close to 1.0. Typically the SiO,:AI,O,:H,O + ratio is
1.1:1:2.5 or a structural formula of (Also, OH,,, 4 H 2 0 ) 2 (Si80,,)OH,) (Wada and
Yoshinaga, 1969). Imogolite is capable of absorbing up to 46g H20/100g dry clay.
C.E.C. values are reported to be in the range of 3 to 20 meq/100g but vary with
cation type and solution concentration (Wada, 1977).

21/04/19 (1) H alloysite


Halloysite is a crystalline mineral commonly associated with and derived from
less well crystallized allophane and imogolite. It usually occurs as rolled, cylindrical,
or quasi-spherical particles (Fig. 2-8). Two principal types exist: a less hydrous form
with a composition near that of kaolinite, A12Si20,(OH),, with a basal spacing near
7.2 A, and a more hydrous form with a composition near Al2Si2O5(OH), 2 H 2 0
A
and a basal spacing near 10.1 (Brindley, 1980). The more hydrous form was called
halloysite and the less hydrous form metahalloysite. It is now recommended (Bailey.
1980) that the terms halloysite (10 A) and halloysite (7 A)
be used. Halloysite (10 A)
transforms to halloysite (7 A) at about 70°C.
Halloysite (7 A) is a highly disordered form of kaolinite. In halloysite (10 A),
2 H 2 0 per formula unit occurs between the 1:l layers. The water molecules are
arranged in a hexagonal network and linked to each other and to 0 and OH ions of
21

Fig. 2-8. Electron micrograph of spherical and tubular halloysite. Bars = 0.2pm. From Dixon and
McKee, 1974. Copyright 1974 the Clay Miner. SOC.

the adjacent kaolinite-type layers by hydrogen bonds. However, infrared spec-


troscopy studies suggest hydrogen bonds are not present and the interaction
between water sheets and the kaolinite-like layers are due to dipole attraction (Yariv
28

TABLE 2-3
X-ray powder data for halloysites

7 A-halloysite 10 A-halloysite
001 and d (obs.) I d (calc.) d (ohs.) I 00 1
hk indices indices
10.1 10 00 1
001 7.41 6 7.21
02.11 4.43 10 4.460 4.46 8
002 3.603 4 3.607
3.40 5 003
13,20 2.562 4 2.575 2.56 5
? 2.493 - -
003 2.405 1 2.405
? 2.340 - - 2.37 3
I
04.22 2.222 2.229 2.23 3
I
004 1.805 1.803
15. 24, 31 1.680 2 1.685 1.67 3
06. 33 1.484 5 1.487 1.48 5
I
26.40 1.283 1.287 1.28 1
17. 35. 42 1.233 1; 1.236 1.23 1
I
006 1.203 1.202
I
OX. 44 1.110 1.114
I
28. 37, 51 1.023 1.023
I
19, 46. 53 0.970 0.973
I
39. 60 0.858 0.859
I
2. 10; 48. 62 0.842 0.843

7 A-halloysite. Brindley and Robinson (1948).


10 I\-halloysite. Mehmel (1935).

and Shoval, 1975). There is some disagreement as to whether the tubular and
spherical morphology is due to layer disorder or to curved layers (Brindley, 1980).
Table 2-3 lists the x-ray powder data for halloysites. Because of the morphology.
the broad hk diffraction peaks of halloysite are normally enhanced when oriented
slides are x-rayed. Bates (1959) explained the curvature in terms of misfit between a
larger tetrahedral layer and a smaller octahedral layer. His study of the chemical
composition of kaolinite and halloysite indicated the latter had more H,O + . He
believed the excess water weakened the interlayer bonds and prevented the oxygens
of the surface of the tetrahedral sheet from stretching the basal hydroxyls of the
adjacent octahedral sheet, thus maintaining the misfit between the smaller oc-
tahedral and larger tetrahedral sheet and forcing the layer to curve. Radoslovich
(1963) does not believe this explanation is compatible with his observation on the
relative ease with which the tetrahedral layer adjusts to the octahedral layer.
However, it is necessary to have unequal stresses in order to create the necessary
curvature. He believes the unbalanced stresses are the expansion due to AI-AI
repulsion across shared edges and a contraction within the layer of surface hydrox-
29

Original After dehydration


Fig. 2-9. Idealized morphological change of halloysite (10 A ) particle by dehydration. From Kohyama et
al.. 1978. Copyright 1978 The Clay Miner. Soc.

yls, probably by OH-OH bonds in the hydroxyl triads around vacant octahedral
sites.
Fig. 2-9 is the idealized model of morphological changes that occur by dehydra-
tion (Kohyama et a/., 1978). Halloysite layers are rolled spirally into tubes. The
upper picture is a cross-sectional view and the lower the transmission electron
micrograph images of the tubes. Each layer means a “unit” equivalent to the
crystallite thickness; the width of stripes in the images is similar to the crystalline
thickness. The “unit” consists of a small number of unit-cells of 20.7 or 14.9 A A
thickness. The width of the stripes indicates the number of unit-cells ranges from 2
to 10. When halloysite is fully hydrated the “units” are tightly connected with each
other, but become separated during dehydration as the unit cell shrinks froni 20.7 to
14.9 A.
Halloysite chemical analyses (Table 2-4) show more variability than analyses of
the other kaolin minerals. Part of this may be real, but part is due to the
fine-grained poorly crystalline nature of the material which makes impurities more
difficult to detect. The content of low temperature water depends largely on the
dehydration history of the clay and has little meaning otherwise. For most samples
high-temperature water is larger than the theoretical value of 13.96%.The high value
is believed due to water trapped between layers during dehydration. Si0,:AI .O,
ratios range from 165:lOO to 206:100, although most values fall in the 1XO:lOO to
200:lOO interval. The average ratio for 22 samples is 188:100, which is identical to
the ratio reported for the well-crystallized kaolinites of Georgia (Hinckley, 1963).
The SiO,:AI,O, ratios for the various kaolin minerals suggest that most of the
variations are a function of inpurities. Ninety percent of the analyses show a ratio
less than 2:1, suggesting that an excess of AI,O, is much more common than an
excess of SiO?. Much of this A1,0, is probably present as aluminum hydroxides
adsorbed on the surface and between the layers of the kaolinite particles.
30

Table 2-4
Chemical Analyses of Halloysites.
1 2 3 4
SiO? 44.08 45.20 44.3 44.64
A1 2 0 , 39.20 38.96 39.1 34.89
Fez01 0.10 0.21 0.4 2.00
MnO - - - - _

MgO 0.05 0.08 - 0.01


CaO none - - 0.09
K .O 0.20 - - 0.06
Na 0.20 - - 0.32
TiO. - - 0.1 0.10
H20- 1.44 - 4.0 2.60
H,O+ 14.74 15.35 13.4 14.30
FeO none ~ - ~

P20, - - - 0.08

Total 99.81 99.80 101.3 99.91 I'

Si02:A120, 191:100 197:lOO 192: 100 218:lOO


" Includes 0.82% SO,
Hallq~.sires
1. Ross and Kerr (1934). Hickory, N.C.J.G. Fairchild, analyst.
2 . Alexander er al. (1943). Djebal Deber. Libya dried at 110°C. L.T. Alexander. analyst
3. Loughnan (1957). Bedford, Indiana. G.T. See, analyst.
4. Keller ei a/. (1971). Los Azufres Thermal area. Michoacan. Mexico.

The average FezO, (0.43%)and TiO, (0.08%) values for halloysites are less than
those for kaolinite. However, iron-rich halloysites (ferrihalloysites) have been de-
scribed which contain from 8 to 25% Fe20,. The Fe presumably occurs in the
octahedral layer, filling from 15 to 50% of the occupied positions (Weaver and
Pollard, 1973).
For a discussion of the origin of kaolin minerals, see Chapter 111 and Chapter
VIII, pp. 540-542.

SERPENTINE

Structure

The serpentine group of minerals has a 1:l layer structure consisting of a


trioctahedral sheet and a Si tetrahedral sheet. In addition to the coarser minerals.
such as antigorite and chrysolite, the group includes the clay-sized minerals
berthierine (7 A chamosite), amesite, greenalite, and cronstedtite. The 001 spacings
range from about 7.0 to 7.4 A.
In chrysolite and antigorite the Mg-rich octahedral sheet is larger than the
tetrahedral sheet. In the case of chrysolite, articulation is accomplished by tilting of
31

the tetrahedra, causing the 1:l layers to curl into cylindrical rolls with the larger
octahedral layer on the outer convex side. In antigorite the curling tendency is
interrupted periodically by inversion of the tetrahedra so the direction of curvature
is reversed, producing a wave pattern.
In most of the other trioctahedral 1:l minerals there is sufficient R7+substitution
in the tetrahedral sheet (increases size) and octahedral sheet (decreases size) so that
the sheets are more nearly equal in size and the morphology is platy rather than
tublar. Where the tetrahedral sheet is larger, accommodation is made by tetrahedral
rotation (Bailey, 1980).

Chemistry
Most of the coarser serpentines deviate very little from the ideal composition
Mg,Si4OI0(OH),. Minor amounts of Al'+, Fe'+, and Fez+ commonly substitute
for Mg2+in the octahedral sheet. Analyses (28) of serpentine reported by Faust and
Fahey (1962) have an average octahedral composition of 0.05 Al'+, 0.09 Fe3+, and
0.09 Fez+. The platy serpentines commonly have more Al'+ than the tubular
variety. Gillery (1959) synthesized serpentines of varying composition and found
that pure Mg Si serpentines were tubular. When the A1 content of both the
octahedral and the tetrahedral sheets was greater than 0.25 per O,,(OH),, the misfit
of the sheets was minimized to the extent that a platy morphology developed.
The clay-sized 1:1 trioctahedral minerals have almost an infinite variety of
compositions (Table 2-5). Greenalite and cronstedtite are essentially the Fe analog
of Mg serpentine. Si is usually the only cation in the tetrahedral sheet, and Fe is the
predominant, and commonly the only, cation in the octahedral sheet. Minor
amounts of Mg, up to a maximum of 0.8, are commonly present in greenalite
(Guggenheim et al., 1982; Cole, 1980). Both Fe3+ and Fe2+ are reported but it is
believed that most of the Fe was originally present as Fe2+ and some was later
oxidized. These Fe-rich minerals are found primarily in low-grade metamorphic iron
formations. Caryopilite is the Mn analog of greenalite. Both greenalite and caryopi-
lite have an excess of Si and a deficiency of total octahedral cations for a
serpentine-type structure. An extra tetrahedron is inserted for every eight tetrahedra
to facilitate articulation with the larger octahedral sheet. Some of the tetrahedra
may be inverted (Guggenheim et al., 1982).
Berthierine has a relatively high content of both tetrahedral and octahedral Al, as
does amesite (Table 2-5). It differs from amesite in that Fe2+ rather than Mg2+ is
the dominant octahedral cation. A typical formula is
( A10.79 Fei.12 Fe4:.; Mg0.25) ( si1.Z1 0.79 )O.5 (OH )4
Tetrahedral A1 ranges from 0.32 to 0.90; octahedral A1 has a slightly larger range,
0.30 to 1.0. The other octahedral cations have a smaller spread of values. Most
octahedral F e 2 + values are in the range of 1.50 to 1.80 (Brindley, 1982; Bhat-
tacharyya, 1983). The negative tetrahedral charge is balanced by the presence of
R" cations in the octahedral sheet; the R'+ octahedral cations exceed the Al'+
tetrahedral cations, and electrical neutrality of the layers is achieved by vacant
-
octahedral sites ( 0.05-0.3). The ionic bonding between the negatively charged
32

Tiihle 2-5
Chemical Analyses and Structural Formulas of Amesite (1). Greenalite (2). Cronatedtite (3). and
Berthierine (4.5.6).
1 2 3 4 5 6
SiO? 20.95 30.08 16.42 26.40 2237 32.6
Al 35.21 - 0.90 1X.23 23.24 10.25
FC,O, - 34.85 29.72 5.70 n.d. 38.90
FeO X.28 25.72 41.86 2537 40.15 n.d.
MgO 22.88 - ~ 11.35 1.77 4.79
CaO 0.58 - 1.32 0.42 0.09
H?O+ 13.02 9.35 10.17 10.60
H ?O- 0.23 - - 1.05

101.15 100.00 100.39 100.00 xx.12 86.54


Octahedral 1 2 3 4 5 6
Al 1.oo - ~ 0.60 0.86 0.30
Fe’+ - 1.23 0.71 0.23 ~
1S O
Fe?’ 0.33 1.29 2.38 1.13 1.91 I1.d.

MgO 1.64 - - 0.91 0.15 0.37


z: 2.97 2.52 3.09 2.87 2.92 2.17

Tetrahedral
Si 1.01 1.74 1.12 1.42 1.30 1.hX
Al 0.99 - 0.07 0.58 0.70 0.32
Fe‘+ - 0.26 0.81 -

.‘ Includes N a z O = 0.17 and K z O = 0.17


1. Gruner (1944). Amesite, pale bluish-green crystals, Chester, Mass. F.V. Shannon. analysist.
2. Gruner (1963). Greenalite. granular aggregates, Mesabi Range. Minn., No. 45758. Recalculated t o
100.00%.
3. Hendricks ( 1939b). Cronstedtite. Kisbanya, Hungary, Cossner, analyst.
4. Banister and Whittard (1945). Oolites from Silurian siltstone. Wickwar, England.
5. lijima and Matsumoto (1982). Replacing kaolinite at burial temperature 65’ to 150°C‘. Utatsu
District. Japan.
6. Rohrlich ei ul. (1969). Pellets in recent muds. Loch Etive, Scotland

tetrahedral and positively charged octahedral sheets contracts the C axis to 7.05 -
A as compared to 7.15 A for kaolinite.
Recent marine 7 A minerals contain primarily Fe”, whereas the ancient varieties
contain primarily Fe2+. Odin (1984) suggested they are not the same mineral.
Berthierine is believed to form by the alteration of kaolinite, and perhaps other
physils. in moderately reducing environments where Fe is mobilized during the early
stages of diagenesis. Like glauconite, berthierine usually occurs in the form of
pellets, probably fecal, or ooids with a peloidal nucleus (Bhattacharyya, 1983).
For additional information on origin, see Chaper VI.

ILLITE
The term “illite” was proposed by Grim et ol. (1937) for the clay-size mica-type
mineral occurring in argillaceous sediments. They stated “the term is not proposed
33

as a specific mineral name, but as a general term for the clay mineral constituent of
argillaceous sediments belonging to the mica group”. Later Grim (1968) described
illite as “ a mica-type clay mineral with a 10 A c-axis spacing which shows
substantially no expanding-lattice characteristics”.
In fact, most minerals referred to as illites contain some expanded layers. The
presence of approximately 5% expanded layers can be detected by treatment with
A
ethylene glycol which broadens the 10 peak slightly and decreases the 001/003
peak-height ratio. Smaller amounts of expanded layers can be detected by heating
the sample at 300°C and determining if the peak-width at half-height, of the 10
A
peak, decreases. Srodon and Eberl (1984) suggest that 10 material which contains
a few percent of expanded layers be referred to as illitic material, which seems
reasonable.
Heat treatments of samples from a metamorphic sequence showed that the final
few expanded layers persisted to a burial temperature of approximately 350”, the
beginning of epigenesis (Weaver and Associates, 1984). Another problem occurs
A
when a sample contains a good 10 phase and a mixed-layer illite-smectite (I/S)
with a few smectite layers. For some samples it is difficult to determine if two
phases are present or if all the illitic material contains a few smectite layers. As the
I/S is finer grained than the pure 10 A phase, size fractionation can sometimes
solve the problem.
In the literature, the term illite is commonly applied to micaceous minerals
containing from 0 to 20% smectite layers. If the x-ray patterns are not reproduced in
the article, it is difficult to determine the true nature of the “illite”. Material
referred to as degraded illites have a significant proportion of smectite or vermicu-
lite layers. For a more detailed discussion of illites, see Srodon and Eberl (1984).

Structure

As the muscovite structure beautifully illustrates the complex character of the


A
layer structure, it is discussed in some detail. The 10 illite layer is basically similar
to that of muscovite, a dioctahedral sheet sandwiched between two tetrahedral
sheets (Fig. 2-10). For more detailed discussions of the structure of mica, see Smith
and Yoder (1956) and Radoslovich (1960). An octahedral sheet consists of cations
which are surrounded by six anions (0 and OH), three above and three below each
cation. The tetrahedral sheet is composed of tetrahedra which consist of cations
surrounded by four oxygen ions. The three basal oxygens of each tetrahedron are
shared between cations of adjacent tetrahedra. As for kaolinite, the basal oxygens
have a ditrigonal configuration. The apex oxygens of each tetrahedron point in the
same direction, towards the octahedral layer. These apical oxygens take the place of
hydroxyl ions in the octahedral layer and serve to link the octahedral with the two
tetrahedral sheets The geometry is such that four of the six anions surrounding and
bonded to the octahedrally coordinated cations are apical oxygen and two are
hydroxyl OH ions. The hydroxyl ions are situated in the center of a hexagonal
arrangement of the apical oxygens.
34

Fig. 2-10, The muscovite structure. viewed along the a axis. Approximately one-half o f the unit cc1I is
shown. Open circles in increasing size indicate. respectively. silicon. aluminum. tetrahedral oxygcn. :ind
potassium. Hatched circles are hydroxyls. From Brindley and MacEwan, 1953. Copyright 1961 London
Min. Soc.

In muscovite, one of every four tetrahedrally coordinated cations contains Al


rather than Si ions (Si3AIO,"). In illite the Al value is slightly less and the negative
tetrahedral charge is less. In dioctahedral muscovite two of the three octahedral
cation positions are filled with A1 ions and the third is vacant. The sheet is neutral.
In the trioctahedral micas, biotite and phlogopite, neutrality is obtained by having
all three positions occupied by the divalent cations F e z + and Mg. Actually, some
trivalent ions are usually present in the octahedral sheet and the sheet has a positive
charge and/or some vacant cation sites. In illite a portion of the octahedral Al is
replaced by Fe and Mg ions, resulting in a net negative charge for the octahedral
sheet. Muscovite has a layer charge of one, all originating in the tetrahedral sheet.
Illite has a charge slightly less than one and differs from muscovite in that the layer
charge originates in both the tetrahedral and octahedral sheets. The negative layer
35

charge is neutralized by large cations. primarily K'. which lay between the 2:1
layers and serve to bond them together.
The idealized charge distribution of muscovite and illite is as follows:

Muscovite Illitr
K' I+ K' O.x+

60 ~ 12- 60 12
3Si4' l A l ' + IS+ 3.5Si4' 0.5AI" 15.5
4 0 2(OH)- 10- 4 0 2(OH) 10
4AI' ' 12+ 3.4AI" 0.6Mg' ' 11.4'
40-2(OH)- 10- 4 0 2(0H) 10
X i 4 'I A I " 15+ 3.5Si4 ' 0.5AI" 15.5
60 - 12- 60 12
K' 1' K+ O.x +

Radoslovich (1960) demonstrated that the sheets making up the mica layer d o
not have an ideal hexagonal symmetry but are distorted to accommodate the
difference in the size of the cations and the size of the two types of sheets. The
smaller (b-axis) octahedral sheet is the controlling sheet, and the tetrahedral sheet
can easily adjust to match the size of the octahedral sheet by rotating the tetrahedra;
in turn. the octahedral sheet is stretched to match the width of the tetrahedral sheet.
Both the octahedral and tetrahedral sheets deviate from the ideal hexagonal
symmetry. The two filled octahedral sites are smaller than the vacant site. The
coplanar 0 ' s around the filled (Al") sites have 0-0 = 2.8 A and around the vacant
site 0-0 = 3.34 A. This is accomplished by rotation of the upper and lower triads
of anions around the octahedral site (see Fig. 2-1). The octahedral sheet is stretched
so that i t can more closely match the width of the tetrahedral sheet.
The tetrahedral sheet decreases its b dimension by rotating individual tetrahedra
11" to 13" (7"--9f" for biotite and phlogopite). In addition to being rotated. the
tetrahedra are tilted. Tilting is necessary in order to accommodate the unequal
charge distribution caused by an unfilled octahedral site and A13' substituting for
Si4+in the tetrahedral sheet.
The tetrahedral sites Si, and Si, (Fig. 2-10) contain, respectively, no Al3'
(Si,-0 = 1.62 A)
and on an average Sif/'Al~~,(Si,-O= 1.69 A). The lack of bonds
from 0,,, 0,,,and OH towards the vacant octahedral sites allows the anions to be
moved away froni their "ideal" position. Also. the apical oxygens (O,, ) attached to
the larger tetrahedron (AI") can move more freely than Ol,. The shift of 0, is
away from Ol,. The Si, tetrahedra adjust by tilting. The O,-Si,-O,, angle increases
from 109 1/2" to 115 1/2". This directed force also acts o n the interlayer K ion.
In fact, Radoslovich (1962) demonstrates by niultiple regression analysis that the
tetrahedral sheet has little or no effect on the size of the b axis. This is because o f
the ease with which tetrahedra rotate.
Micas b = (8.925 + 0.099K - 0.069Ca + 0.062Mg + 0.116Fe: '
+0.098Fe3' + 0.166Ti) + 0.03 A
However, the large interlayer K ion appears t o stretch the octahedral layer.
36

O f

0
Fig. 2-1 1. Normal projection onto the a-b face of some of the atoms in muscovite. This clearly shows the
di-trigonal character of the oxygen network. the inner ring of six oxygens around K + , and the rotation
of the tetrahedra from the ideal structure. From Radoslovich. 1960. reprinted by permission of
International Union of Crystallography.

In the ideal muscovite structure the K ion is 12-coordinated with six hexagonally
arranged oxygens above and six below. Rotation of the tetrahedra causes a ditrigo-
nal arrangement of the oxygen ions (Fig. 2-11) such that three of the 0 ions are at a
distance of 2.81 A and three at 3.39 A. Thus the K ion basically has a six-coordina-
tion rather than 12. The size of the ditrigonal holes are smaller than the K ions, and
adjacent oxygen planes are 3.4 A apart rather than the expected 0-0distance of 2.8
A.
Radoslovich (1963) calculated that the 0,’s carry an average unsatisfied charge
of about -1/8 (Si, = 1/2 Al”+ + 1/2 Si4+= 7 ‘/K) and the K + of +2/8. Thus,
K ’ is attracted towards, and its charge largely satisfied by, one 0, from each layer.
He states,
“In 2M, muscovite the interlayer K + is held in place by six bonds under
compression. on the average. In detail, the K + occupies an equilibrium position
determined by a complex balanced system of interlacing strong bonds reaching
right through the adjacent layers to K’s at the next level, above and below. The
system of bonds is a direct consequence of 2Mt muscovite being a dioctahedral
mineral with 2A13’ octahedrally, and with an ordered arrangement of 2Si and
2Sil,2All,2 tetrahedrally. It is hardly surprising that this polymorph is one of the
most stable micas under natural weathering. This view of the role of K’ in
37

muscovite is far removed from the early concepts of an ion of the right charge
flopping into a hole of comfortable size!"
Bassett (1960) demonstrated that the H in the O H ion lying immediately under
the K ion adjusts to the repulsive force of both the overlying K ion and the
underlying cations in the octahedral layer. In the trioctahedral micas the O H - sits
over the center of a triad of R2' ions and the H + is at the top of the 0. directly
under the K'. As a result of the K-OH repulsion, the K is less firmly bonded and
can relatively easily be exchanged. Because the dioctahedral micas have an unfilled
octahedral position, the balance of repulsive forces causes the 0 - H bond to be
inclined 18" to the cleavage plane and as a result the K ion is more strongly bonded
in the interlayer position.
Illite flakes are too small to afford good structural information; however,
structural data have been obtained for a phengite, which is similar in composition to
illite (Guven, 1970).

(A1I .43 FeG, Fe,:, Mgo.,,, ) (Si 3.3') A1 (1.6, )O,o( O H 1.

The higher Si content in the phengite tetrahedra, as compared to muscovite,


decreases the electrostatic unsatisfied charges o n basal oxygens and thus the
interlayer cation-basal oxygen bonds are weaker. The relative abundance of the
large Mg and Fe2+ cations cause polyhedral misfit and create strains in the
octahedral sheets. This strain limits the amount of R'+ which can be accommodated
in the octahedral sheet. The substitution of the large R" ions for Al in the
octahedral sheet and the Al for Si substitution in the tetrahedral sheet decreases the
misfit between the two sheets. The tetrahedral rotation, 6", is approximately half
that for muscovite. This produces a longer K - 0 bond length. indicating the bonds
are weaker.
The oxygen packing in mica causes an offset of the upper and lower hexagonal
nets two-thirds of one oxygen diameter (1.7 A ) or a/3 (Fig. 2-12). When this is
combined with the ideal hexagonal symmetry of the surface oxygen network, six
simple polymorphs are possible (Smith and Yoder, 1956). However, Radoslovich
(1960) suggested that because the surface oxygens actually have a trigonal symmetry
surface they can best fit together in ways which correspond to n o rotation or
rotations which are multiples of 120" between layers (rather than 60" rotations
allowed by hexagonal symmetry). This allows three polymorphs, l M , 2M,. 3T.
Three assumptions are made:
1 ) The trigonal symmetry precludes 180" rotations between layers.
2) The K ion is displaced from the center of the oxygen network by some m a l l
force.
3) The two K ions on opposite sides of one layer tend to move as far apart as
possible.
The K: at the top of layer A is acted on by a small force in the direction of 0,
(see preceding discussion). In the 1M structure the same K: is acted on by an
One
unit
1oA

I
Fig. 2-12. A projection on (010) showjing one phlogopite layer. together with adjacent parts of the t u n
adjoining layers. 0, and 0, are at the vertices of the tetrahedra. The 0, oxygens overlap with the OH
ions. From Smith and Yoder. 1956. Copyright 1961 London Min. Soc.

opposing force from the bottom of layer B (Fig. 2-13). A more stable state is
reached when the two forces act as nearly as possible in the same direction.
"The nearest permissible approach to this, because of the trigonal symmetry. is at
60" to each other; and the resulting force on K,' will be between the two. The

B r n ' l c mP B RcIITcEl

B Top B TOP C WI'IDY


WSIJLTKVT
Fig. 2-13. ( a ) Forces on the K ' ions in the 1 M structure. from successive layers A and B. ( b ) Forces on
the K t ions in the 2 M , structure, from successive layers A. B and C. ( c ) ditto in 3T htructure. From
Radoslovich. 1Y60 reprinted by permission International Union of Crystallography.
39

force on K l at the top of layer B is then at 120” to that on K: at the top of


layer A. If layer C also rotates relative to B (to likewise reach a more stable
position), then this rotation may be either a further +120”. or else -120”. Of
these, the latter results in a net force on, and displacement of, K: which is
directly opposite the resultant force on K:. If assumption (3) is correct, then this
is the more stable arrangement; and i t is seen that the net effect is an alternating
+120” rotation between layers, as required for the 2M, structure (Fig. 2-13b).
The alternative position of layer C (Fig. 2-13c) corresponds to the 3T structure.
This does not remove K: as far as possible from K:, and would not be so likely
to occur as the 2M, arrangement.”
All three polytypes, as well as a disordered 1Md variety. are found in illites. I t
has been generally accepted that most of the illite i n shales and clays that was not
the 2M, polytype was the 1Md polytype. Thus, a 2MI/1Md ratio is commonly
reported for shales. Srodon and Eberl (1984) suggested that if the 1Md polytype is
defined as having zero ordering, then there are no 1Md illites, but that samples so
described represent varying degrees of 1M ordering. The 1Md/lM ratio is related to
the percentage of illite layers. The presence of the 22-33” band may serve as a
criterion to distinguish the 1Md polytype from turbostatic stacking on the one hand,
and from the 1Md polytype on the other. lllites produced at low temperatures from
smectite, kaolinite, feldspar, and authigenically in sandstones are 1 M or 1 Md illites.
Hydrothermal experiments by Yoder and Eugster (1955) and Velde (1965)
showed that the sequence 1Md + 1M + 2M, was a function of increasing tempera-
ture and hydrothermal run time. Numerous studies (see Diagenesis Chapter) have
demonstrated that 1Md illite, or 1M illite, is systematically converted to 2M, illite
with increasing thermal metamorphism. Weaver and Broekstra (1984) found that
during regional metamorphism 1Md illite was completely converted to 2M, illite or
phengite at approximately 360”C, the beginning of epizone metamorphism.
Table 2-6 contains the x-ray powder data for muscovite, which are similar to
illite. Table 2-7 shows the spacings needed to identify the various polymorphic
varieties of mica and illite.
Typical x-ray patterns of illite are shown in Fig. 2-14. The width of the 001 peaks
is a measure of the “purity” of the illite. Illites with 9 to 10% KzO are composed
largely of 10 A mica-like layers. Illites with less K,O have broader peaks largely
caused by the presence of chlorite (14 A) and smectite (17 A, e g ) layers. This is
indicated by broadening of the 001 (10 A and 14-17 A ) and 003 (3.33 A and 3.5 A )
on the low angle side and the 002 (5.0 A and 4.7 A) on the high angle side. As the
iron content increases, the relative intensity of the 001/002 peaks increase; the
reverse relation holds for K. As the K,O plus Fe,O, contents of illites are inversely
related, the 001/002 ratio is a reasonable measure of the K 2 0 and Fe,O, values. In
the better crystallized micas and phengites, where the K,O values are relatively
constant, the 001/002 ratio is a measure o f the Fe content of the octahedral layer.
Conversely, in the illites where the amount of Fe,O, is relatively constant (2 to 5 % ) ,
the ratio reflects the K,O content (Weaver, 1965).
The 060 values are used to determine whether an illite is dioctahedral (1.50
40

TABLE 2-6
X-ray powder data for muscovite. After Grim. Bray. a n d Bradley (1937).
2M , Muscovi tc
indices d I

002 9.99 S

004 4.98 m
110
4.47 vs
iii
111 4.29 W
002 4.1 1 W
112 3.95 vw
113 3.87 m
023 3.72 m

113 3.55 VW

114 3.48 in
024
3.32 VS
00h
114 3.20 ins
113 3.1 vw
025 2.98 \

115 2.8h m
116 2.78 m
I3i
2.585 W
200
202
2.56 vs
131
008 2.49 W

202: 133 2.46 W

204 2.39
mdh
133 2.38
204; 133 2.245 wd
223 2.185
206 2.14
mdh
135 2.1 3
136
044 2.05 vw
0. 0. 10 1.99 5
206: 137 1.95 W

1.83 vw
138 1.76 W

2. 0. TO 1.65 W

139 1.64 m
Oh0
1.504 S
235
41

TABLE 2-7
Powder diffraction data for distinguishing polymorphic variables of mica. After Yoder and Eugster. 1955.

iii 4.46 111 101 4.46 W


02 1 4.39 \' W

1i i 4.35 W 111 4.30 W


021 4.12 vw 022 4.1 1 vw
112 3.97 VW'

1IJ 3.89 W 104 3.87 W

023 3.74 W
113 3.66 W 105 3.60 W

114 3.50 m
006
Oo3 3.36 vs 024 3.35 vs 009 3.33 Vh
022
114 3.21 m
112 3.07 m 107 3.1 1 W
025 3.00 in
113 2.93 m 115 2.87 W 1OX 238 W

116 2.80 W'


023 2.69 m

A X 6 = 9.00 A)
or trioctahedral (1.54 A X 6 = 9.24 A).
I t can also be used to
indicate the proportion of large (Mg + Fe) ions in the octahedral layer (Maxwell
and Hower, 1967).

Chemistry
Table 2-8 contains the chemical composition and structural formula of a selection
of illites and illitic material. As noted. most illites contain some non-illite layers;
therefore, the analyses tend to reflect the proportion of non-illite, largely montmoril-
lonite, layers. Only a few samples have been described which are composed entirely
of 10 A layers.
The chemistry of "illites" was summarized by Weaver and Pollard (1973). Based
on the analyses of minerals called illi te. they obtained the following average
formula:

These analyses include illites containing from 0 to approximately 15% smectite


layers. The average is reasonably representative of the illitic material in shales.
Information on the composition of pure illite has been obtained by plotting the
compositions of I/S physils and extrapolating the data to 100% illite layers
(Weaver, 1979):
42

I I / I II I j I 1 I I I l l ! ~ I l J I 1 I I i I I I I [ I !I I ! I I j / I ( I I I1 1 1
10 20 30 40 50° 28
Fig. 2-14. Examples of X-ray powder diffraction patterns of pure illite and illitic materials lacking
noncoincident illite smectite reflections. For samples 28 and 41, air-dry patterns are added to the patterns
of the glycolated preparations to show 001 and 003 peak intensities. Sample 28 exemplifies a poorly
oriented preparation where the 34-36'2 region cannot be exploited due to the presence of a 20 reflection.
From Srodon. 1984. Copyright 1984 The Clay Min. Soc.

The layer charge of 0.75 is the same as that Hower and Mowatt (1966) and Srodon
and Eberl (1984) obtained by extrapolating the K content as a function of the
proportion of illite layers in I/S. Plots of cation exchange capacity vs. expandability
give a CEC for end-member illite of 15 meq/100 g.
lllitic material with a layer charge less than 0.75 and K,O less than about 8% can
generally be assumed to contain some non-lOA layers. Illites with less than 5%
smectite layers have layer charges ranging from 0.75 to 1.0, though those with a
charge of 1.0 are usually called phengites. Illites typically have from 50 to 100% of
their layer charge originating in the tetrahedral sheet. For most illites, the trivalent
43

Table 2-8
Chemical Composition of lllites

1 2 3 4 5 6 7 8 9 10
~ -
SiO, 50.55 47.4 49.85 52.87 52.58 49.67 48.95 47.94 49.03 54.18
A1 ,b3 26.14 35.6 23.68 24.90 27.04 27.31 25.89 33.08 17.93 14.20
FeA 0.67 1S O 6.60 0.78 0.72 2.96 5.02 2.23 13.11 12.11
FeO 0.65 1.87 1.19 0.24 2.40 1.31 2.03
MgO 4.25 0.30 1.86 3.60 2.75 1.09 2.25 1.49 2.79 4.01
CaO 0.60 0.12 0.69 0.24 0.29 0.45 0.34 0.39 0.68
Na,O 0.19 0.53 0.34 0.22 0.05 0.10 0.15 0.32 0.10 0.04
K2 0 10.29 9.12 6.64 7.98 8.47 7.26 6.70 9.48 7.84 6.07
TiO, 0.42 0.23 1.40 1.02 0.33 0.23 1.06 0.22
H2O+ 4.59 6.80 6.73 6.61 9.00 7.60 6.00
H20- 0.99 0.0 2.56 2.03 1.50
Total 99.34 99.16 102.54 101.03 99.41 99.93
Octahedral
Al 1.51 1.90 1.41 1.51 1.68 1.71 1.48 1.76 0.99 0.89
Fe” 0.03 0.08 0.35 0.04 0.03 0.16 0.26 0.11 0.70 0.63
Felt 0.03 0.11 0.07 0.01 0.13 0.08 0.12
Mg 0.43 0.03 0.19 0.36 0.28 0.1 1 0.23 0.15 0.30 0.41
Tetrahedral
Al 0.57 0.87 0.53 0.46 0.47 0.53 0.62 0.82 0.51 0.26
Si 3.43 3.13 3.47 3.54 3.53 3.47 3.38 3.18 3.49 3.74
lnterlayer
Ca 0.09 0.02 0.10 0.01 0.04 0.03 0.05 0.03 0.05
Na 0.02 0.10 0.05 0.03 0.01 0.01 0.02 0.01 0.01
K 0.89 0.77 0.59 0.70 0.72 0.65 0.59 0.80 0.71 0.53
Layer Charge 1.03 0.87 0.65 0.95 0.76 0.70 0.68 0.91 0.78 0.64
%
’ Expanded
Layers 0 0 0 <5 6 mxill 5 small

Polytype 2M, 2M, 2M, 2M, IS11 IM IM IM 1Mor I M oi


1Md 1Md
I. Weaver: Precambrian shale bed in Belt limestone. Glacier National Park. Montana; Analyst E.G.
Oslund.
2. Srodon and Eberl. 1984: Hydrothermal, Kauhe. Japan.
3. Mankin and Dodd. 1963: Silurian, Blaylock shale. Beavers Bend. Oklahoma: Analyst J.A. Schleicher.
4. Gaudette. 1965: Laminated pockets in Silurian dolomite. Marblehead, Wisconsin: Analyst J. Witters.
5. Srodon ef d., 1986: Silurian bentonites, Welsh Borderlands.
6. Weaver, 1953: Weathered feldspar in Ordovician graywacke. State College. Pennsylvania: Analyst G.
Kunze.
7. Kodama and Dean. 1980: Mudstone from Eldorado, Saskatchewan.
8. Triplehorn. 1967: Post-depositional in Cambro-Ordovician sandstone. Algeria.
9. Keller. 1958: Lacustrine fine-grained Jurassic sandstone. Blue Mesa. Uravan. Colorado.
10. Porrenga. 1968: Lacustrine or lagoonal Oligocene clay, Aardenburg. Belgium.

octahedral population is larger than 1.5. For additional information on the composi-
tion of illite, see the section on mixed-layer illite/smectite.
Table 2.8 contains chemical analyses of a number of 2 M , and 1M illites which
44

contain few or no smectite layers. The range of compositions fairly well typifies the
range of illites in general.
A number of Fe-illites have been described that are intermediate in composition
between illite and glauconite (Table 2-8, Nos. 9 and 10). These Fe-illites contain
between 7 and 16% Fe,O,. They are either the 1Md or I M polytype. The 002 x-ray
peak is very small relative to the 001 and 003 peaks. They are neoformed or
transformed from other physils in continental basins of a lacustrine or lagoonal type
and are commonly associated with gypsum and dolomite, indicative of evaporitic
conditions (Kossovskaya and Drits, 1970).
Paragonite is 2M, N a mica with an 001 spacing of 9.6 A. I t is normally a
coarse-grained mica found in metamorphic rocks; however, paragonite occurs in the
< 2 fm fraction of slates and incipient slates (late stage diagenesis) (Weaver and
Associates. 1984).
Brammallite is a rare illite that contains about twice as much Na as K (Bannister,
1943).
Sterne et ul. (1982), Juster et al. (1987) and others described illites from
organic-rich black shales in which NH; substitutes for more than half the K'. The
001 spacing is slightly larger than 10.0 A. The NH,' illite (tobelite) can occur as a
discrete mineral and interlayered with K illite.
TEM and SEM pictures of shale illite (Chapter VII) indicate it occurs as thin.
generally angular flakes. Lath shaped illite has been observed forming from K-
feldspar. Authigenic fibrous illite is relatively abundant in sandstone (see Chapter
VIII).
Illite forms by the continental weathering of feldspars, micas, and possibly
volcanics. It forms directly from solution at both low (neoformation) and high
(hydrothermal) temperatures, and by the conversion of smectite to I/S during
burial. Details on origin are presented on the pages indicated:
Soils: pages 174-181
Lacustrine: pages 273-274
Evaporites : pages 408, 414
Metamorphism: pages 454-513
Sandstones: pages 533-536. 551-560

PHENGITE

Some white micas in metamorphic rocks have more Si, Mg, and Fe and less Al
than muscovite. Metamorphic geologists refer to this material as phengite. The
structure of phengite is described in the illite section. Though phengites occur in
schists, they also occur in rocks of lower metamorphic grade. Weaver and Associates
(1984) found phengite formed from biotite and chlorite in the late stage of
diagenesis and throughout the anchizone. During burial metamorphism smectites
are systematically converted to I/S with increasing illite layers and total layer
charge. Most shales have never been exposed to temperatures in excess of - 200°C.
When exposed to temperatures of 200" to 250°C, "illites" contain 5 t o 10%smectite
45

Table 2-9
Chemical Composition of Phengites.
1 2 3 4
SiO, 49.6 50.45 48.13 51.08
A1 2 0 3 22.2 25.77 28.34 20.40
Fe,O, 2.07 0.99 4.07
FeO 1.60 2.54 2.75
MgO 4.82 5.02 2.71 4.05
CaO 0.03 0.29 3.08
NazO 0.07 0.56 0.39 1.16
K2 0 10.5 10.11 11.oo 7.08
TiOz 0.12 0.19 0.95
H,O+ 4.29 5.28
H2 0 - 0.09 0.20

Total 99.29 100.1

Structural Formula
Octahedral
Al 1.39 1.43 1.51 1.14
Fe, + 0.12 0.05 0.20
Fez+ 0.09 0.14 0.16
Mg 0.51 0.50 0.27 0.42

Tetrahedral
Al 0.47 0.61 0.75 0.50
Si 3.53 3.39 3.25 3.50

Interlayer
Ca 0.01 0.02 0.19
Na 0.01 0.07 0.05 0.15
K 0.93 0.87 0.95 0.63

Layer Charge 0.92 0.99 1.40 1.32

Polytype 2Ml 2Ml 2M 1


i C r 2 0 i and Cr,,.,,,
1. Weaver. 1984: Metamorphosed illite, greenschist facies; Cambrian. Conasauga Formation. Georgia:
Analyst K. C. Beck.
2. Emst. 1963: Schist, Tiburon Peninsula. California: Analyst C.O. Ingamells.
3 . Rule and Bailey. 1985: Rio de Oro. Spanish Sahara.
4. Ernst. 1963: Schist. Kotii District, Shikoku, Japan: Analyst H. Haraiiiura.

layers and have a layer charge of 0.7 to 0.8. With increasing temperature, passing
out of the realm of shales, metamorphism continues and the layer charge increases
to approximately 1.0 as the final smectite layers are converted to 10 A layers. The
final product is called phengite; actually i t is a “true illite.”
Table 2-9 contains analyses of a few phengites. For a high percentage of the
analyses in the literature the calculated layer charge does not equal the sum o f
interlayer cations. In other instances the phengites appear to be muscovite with a
46

minor amount of R2+ in the octahedral sheet. The typical phengite has a formula
similar to:
(A1I 37 Fe,: : J e t I: Mg, 38 )(Si 7 11 )
‘41 0 67 )Om ( O H M K 0 xoNa0 ,&a 0 05

The layer charge is approximately 1.0 as compared to 0.70 to 0.80 for most illites. I n
general, phengites have a higher tetrahedral charge (0.65 to 0.70) than illites (0.45 to
0.60). though there is some overlap.

06/03/19 (1) SM ECTITE

Smectites are 2:l physils with a structure similar to that of illite. In general, the
layers are stacked in a random manner, though beidellites have a fair degree of
crystallinity (Grim and Guven, 1978). The layer charge is commonly less than
0.5/0,,,(OH)2. The average layer charge is approximately 0.35. As a consequence of
this low layer charge, smectite is not able to bond interlayer cations with sufficient
force to cause adjacent layers to contract. Exchangeable cations, water, cation-hy-
droxy complexes, and organic molecules can occur in the interlayer space.
The negative layer charge, due to isomorphous substitution, is most commonly
balanced by interlayer cations which are easily exchanged. The weight of cations
that can be exchanged is referred to as the cation exchange capacity. CEC. About
80% of the exchangeable cations are considered to occur in the interlayer space and
20% on the layer edges, bonded by the unsatisfied charge due to ‘‘ broken” bonds.
The exchange capacity is generally between 60 and 130 meq/100 g. The CEC
increases as the layer charge increases, up to a point. If the layer charge is larger
than - 0.6-0.7/0,,(OH), (vermiculite), the interlayer cations are not easily ex-
changed and the CEC decreases. When interlayer cations are not easily exchangea-
ble the layer charge is high enough that the layers can contract to 10 A when K + is
the interlayer cation.
In nature H + , Na+, Ca”, and Mg2+ are the dominant exchange cations. Marine
smectites have N a + and Mg2+ ions as the major exchange cations. In soils, H + ,
Ca+, or Mg” can be the predominant cation. Also, in soils A1 and Fe hydroxides
can precipitate in the interlayer position. These compounds are not easily exchanged
and decrease the CEC of the smectite. For additional information on CEC, see
pages 170-175, 344-348.
At relative humidities around 50 to 75% smectites normally contain one ( N a +
and Li’) or two (Ca2+,Mg”, H’) water layers. The water is believed to have a
hexagonal ice-like structure. At 100% R.H. Na’ and Li+ montmorillonites will
expand to over 40 A and eventually individual layers will dissociate (Norrish, 1954).
Smectites are extremely fine-grained (Fig. 2-15). The layers can be dispersed to
the unit cell level and flake width can be less than 0.1 fm. These small flakes can
have a wide range of layer charges. The layer charge can be determined by the
variation in basal spacing when a smectite is treated with alkylammonium com-
pound of varying chain length (Lagaly and Weiss, 1975). The great majority of
montmorillonites (the most abundant smectite) have a range of layer charges.
47

Fig. 2-15. Electron micrograph of montmorillonite. Sheets are 1 or 2 unit cells thick. Bars = lpm.
48

Fig. 2-16. Different types of charge distribution for intracrystalline heterogeneous nioiitniorillonitea and
the corresponding basal spacings of the n-alkylammonium derivatives. From Lagaly and Weiss. 1975.
Copyright 1976 Applied Pub. Ltd.

Values range between 0.2 and 0.5/0,,(OH),. Most rnontmorillonites have a layer
charge range of 0.10 to 0.15, 0.24 to 0.36, 0.31 to 0.46. About 70% of the interlayer
charges are between 0.28 and 0.42. Fig. 2-16 shows the charge distribution for six
rnon trnorillonites.
Layer charge inhomogeneity can also be detected by the way the expanded layers
react when exposed to solutions containing K f . CiEel and Machajdik (1981) found
that when a variety of smectites were saturated with K f and treated with ethylene
glycol they contained layers of three different thicknesses: 10 A. 14 A, and 16.8 A.
The higher charged layers ( 20.6) fix K and contract to 10 A; the low charged layers
adsorb two layers of ethylene glycol (16.8 A);
those layers of intermediate charge
adsorb only one layer of ethylene glycol. Alternate wetting and drying (40 to 100
49

times) of K saturated montmorillonite can permanently collapse up to 50% of the


A
layers to 10 and produce an I/S (Srodon and Eberl, 1984).
X-RU~
Smectites are easily identified by the ease with which these layers contract and
expand (Fig. 2-17). The amount of interlayer water depends on the type of
exchangeable cation. Under the relatively humid conditions encountered in labora-
tories, smectites with interlayer N a t and K + adsorb one layer of water and have a
-
layer thickness (001 spacing) of 12.5 A. H+, Ca2+,and Mg2+ saturated smectites
contain two water layers and have a thickness of 14 to 15.5 A. A water layer is
approximately 2.5 A thick. Untreated natural smectites, particularly from a marine
environment, commonly contain both 12 and 15 A layers.
For identification purposes the interlayer water is commonly replaced by ethyl-
ene glycol or glycerol. These organic molecules assure that all layers will have

K SATURATED
0 % RELATIVE HUMIDITY

GLYCEROL VAPOR

ETHYLENE GLYCOL

7 10 14
BASAL SPACING, d BASAL SPACING, d
Fig. 2-17. X-ray patterns of soil smectite subjected to various treatments. Glycerol treatment indicates
both montmorillonite (2 layers, 18 A) and beidellite (1 layer, 14 A) are present. From Borchardt, 1977.
Reprinted by permission of Soil Sci. Soc. Amer.. Inc.
50

approximately the same thickness, thus sharpening the 001 reflections, and demon-
strates the expandability of the smectites. Ethylene glycol treated samples com-
A
monly have an 001 x-ray spacing between 16.6 and 17.2 and for glycerol around
17.7 A (Novich and Martin, 1983). The spacings tend to decrease slightly as the
layer charge increases.
The smectite thickness can be reduced to 9.5 to 10.0 A, depending on the type of
interlayer cation, by heating the sample to 300" to 400°C and x-raying immediately.
The interlayer water can normally be removed at 110°C or at 0% relative humidity.
Fe and Al hydroxides and organic material can occur in the interlayer position and
prevent collapse at low temperatures (see pages 170-175).
The 060 spacing obtained from a random sample can show whether a smectite is
dioctahedral ( - 1.50) or trioctahedral ( - 1.53). More specific identification gener-
ally requires chemical data. The Li treatment (Greene-Kelly, 1955; Malla and
Douglas, 1987) can be used to determine if dioctahedral smectites have primarily an
octahedral charge (montmorillonite) or a tetrahedral charge (beidellite). When a
sample of dioctahedral smectite is saturated with Li+ and dried at 300'C overnight,
the Li+ ion will migrate into the vacant octahedral position (there is some disagree-
ment about this). If the layer charge originates in the octahedral layer, i t will be
neutralized and the neutral layers will not expand when treated with ethylene glycol.
If the charge originates in the tetrahedral layer, it is not completely neutralized and
A
it will expand to 17 when treated with ethylene glycol.
51

The formula shows the composition of the Octahedral and tetrahedral sheets, the
location of the layer charge and the amount of exchangeable ions. These formulae
are idealized and for most smectites isomorphous substitution, and charge, occurs in
both the octahedral and tetrahedral sheets. The exchangeable cations have an
average value near 0.33. Individual samples have a range of values of about 0.2 to
0.6. The total exchange cation charge should equal the negative layer charge. It
should be noted that these pictorial formulae are sensitive to impurities in the
sample. The presence of a few percent of free Al or Fe oxides or hydroxides causes
the tetrahedral charge to be overestimated and free Si02 causes it to be underesti-
mated. Total layer charge is increased or decreased. Because of its fine-grained
nature, impurities are relatively common in smectites.
Montmorillonite and Beidellire: Montmorillonite and beidellite are the two Al-rich
smectites. They are both dioctahedral and differ primarily in that the layer charge of
montmorillonite originates primarily in the octahedral sheet and that of beidellite in
the tetrahedral sheet. The latter generally has the larger layer charge. Typical
analyses are given in Tables 2-10 and 2-11. For additional analyses see Schultz
(1969), Weaver and Pollard (1973), and Grim and Guven (1978). Most structural
formulae have more than 2.00 octahedral cations. It is not known if values in excess
of 2.00 are real or due to impurities. Various studies have shown that Si, Al, Fe,
and/or Mg are commonly present as oxides or in the interlayer position. It is likely
that values larger than 2.10 indicate an impure sample.
The average amount of tetrahedral A1 in montmorillonite is about 0.10 (Weaver
and Pollard, 1973), which indicates 2/3 or more of the layer charge originates in the
octahedral layer.
Most montmorillonites have a layer charge between 0.25 and 0.60. As the average
value is about 0.35, the high charge smectites are relatively rare. Grim and Kulbicki
(1961) suggest there are two end member montmorillonites:
Cheto type: (~l~.3,Fe&LMg~.54 W 3 . 9 1 Al0.09)Olo(OW 2 X0.60
Wyoming type: (A11.72 Fei.,f,
Fei.&MgO,lS 3.96 A10.04 )OIO(OH) 2 0.34

The Cheto type has a layer charge nearly twice that of the Wyoming type. The
difference is due primarily to the high Mg content of the Cheto montmorillonite.
Montmorillonites with a high Fe content also exist. They are discussed under
nontronite. Using multivariate analysis of variance and discriminant analysis,
Alberti and Brigatti (1985) were able to divide the montmorillonites into four
subgroups. The octahedral cations were the most important discriminating varia-
bles.
One beidellite has been described (Table 2-11) in which essentially all the charge
originates in the tetrahedral sheet. In other samples an appreciable portion of the
layer charge originates in the octahedral sheet.
Nontronite: Nontronite is the dioctahedral smectite in which Fe3+ is the domi-
nant octahedral cation (Table 2-12). Nontronite is the Fe analogue of montmoril-
lonite and beidellite. The dominant layer charge can originate in either the oc-
tahedral or the tetrahedral sheet. In the case of nontronite, Fe3+- as well as A13+-
can occur in the tetrahedral sheet. Our knowledge of Fe-rich smectites has been
52

Table 2-10
Chemical Analyses and Structural Formulasof some Montmorillonites of Various Origins.
1 2 3 4 5 6 7 8 9
~~

SiO, 53.98 51.14 51.95 61.77 50.53 48.60 51.7 51.20 62.30
A1 2 0 , 15.97 19.76 18.02 19.85 19.31 18.40 20.2 22.14 23.50
Fe,O, 0.95 0.85 0.21 1.95 7.25 1.21 0.77 - 3.35
FeO 0.19 - ~ - - 0.07 - - 0.37
MgO 4.47 3.22 5.10 5.56 2.60 1.88 3.4 3.53 1.95
CaO 2.30 1.62 2.09 1.89 0.72 2.25 2.5 3.72 0.31
Na,O 0.13 0.04 0.04 0.07 0.41 0.35 0.09 - 0.40
K ,O 0.12 0.11 ~
0.09 0.34 0.28 0.84 0.56 0.03
TiO, 0.08 - 0.02 0.24 0.75 Trace - -

H,O' 9.12 7.99 7.60 7.72 7.90 8.44 7.4 - 6.45


H2 0 13.06 14.81 15.60 - 10.66 17.64 13.0 17.08

100.62 99.75 100.65 99.14 100.64 99.47 99.99 99.83 98.66

Octahedral
Al 1.48 1.64 1.43 1.38 1.40 1.68 1.58 1.69 1.64
Fe' ' 0.05 0.05 0.01 0.09 0.40 0.08 0.05 - 0.15
Fe' * - - - - - 0.01 ~ ~
0.02
Mg 0.52 0.36 0.58 0.54 0.28 0.23 0.37 0.38 0.19
2.05 2.05 2.02 2.01 2.08 2.00 2.00 2.08 2.00

Tetrahedral
Al 0.00 0.12 0.15 0.09 0.28 0.06 0.14 0.22 0.10
SI 4.00 3.X8 3.85 3.91 3.72 3.94 3.26 3.78 3.90

Intcrlayer
Ca/2 0.39 0.20 0.32 0.38 0.41
K - - - 0.03 -

Na 0.02 0.02 0.01 0.06 0.02


- - 0.08
Mg/2 ~ ~

C.E.C..
Layer Charge 0.37 0.33 0.67 0.60 0.32 0.30 0.5 1 0.39 0.31
1. Kerr. 1950: Altered rhyolitic and andesitic tuff. Santa Rita. New Mexico: Analyst: Ledoux and C o .
2. K o s and Hendricks. 1945: Nesta penetrating a shale. Montmorillon, France: Analyst: R.C. Wells.
3. Blokh and Sidorenka. 1960: Cavity filling inextrusive rocks. Selongin Daura, U.S.S.K. (called
Nefedyevite).
4. Grim and Kulbicki, 1961: Altered latitic ash. Cheto. Arizona.
S . Rosa and Hendricks. 1945: Upper Cretaceous Bentonite. Booneville. Mi
Fairchild.
6. Yoshikawa. 1960 (Personal communication): Vein cutting granite, Tottori Prefecture. Japan.
7. Fournier, 1965: Hydrothermally altered plagioclase. Ely, Nevada: Analybt: Elmore. Botts. Chloe.
Artis and Smith.
8. R o s s and Hendricks. 1945: Pegmatic clay. Branchville. Connecticut: Analyst: G.J. Brush.
9. Grim and Kulbicki. 1961: Altered volcanic ash. Cook Co.. Wyoming.

greatly increased by the many analyses of deep-sea smectites (Table 6-6) (McMurtry
et (11.. 1983; Singer et al., 1984). The Fe" content of the octahedral sheet of
dioctahedral smectites ranges from 0.0 to 2.0. The montmorillonites contain an
53

Table 2-11
Chemical Analyses and Structural Formulas of Some Beidellites.
1 2 3 1 2 3
59.30 48.03 49.01 Octahedral
36.1 1 23.23 20.5 Al 1.98 1.75 1.34
0.50 0.03 6.85 Fe’+ 0.02 - 0.39
0.10 2.24 2.08 Mg 0.01 0.26 0.28
0.02 3.48 0.17 E 2.01 2.01 2.01
3.98 - 0.58 Tetrahedral
0.11 0.01 0.95 Al 0.52 0.33 0.27
12.1 Si 3.48 3.67 3.73
10.5 Layer Charge 0.50 0.56 0.52
Total 100.12 * 99.12 * * 103.0
* Loss on ignition: 6.3% (ignited weight).
** Loss on ignition: 22.10%.
1 . Weir and Green-Kelley, 1962: Black Jack Mine, Beidell, Colorado. gouge clay
2. Nadeau el (11.. 1985: Unterrupsroth. France. veins in phonolite.
3. Sawhney and Jackson. 1958: Houston soil.

average of 0.15 octahedral Fe. When Fe values are between about 0.15 and 1.0, the
physils are referred to as Fe-rich montmorillonites; when Fe comprises more than
50% ( > 1.0) of the octahedral layer, the physil is called nontronite. As shown in
analysis No. 4, most marine “nontronites” contain several percent of K 2 0 and are
apparently mixed-layer illite/nontronite.
Montmorillonite and beidellite form from a variety of materials (volcanic glass,
mica, feldspar, hornblende, etc.) under a wide variety of environmental conditions
(marine, lacustrine, soil, hydrothermal). The main requirement is that chemical
weathering not be overly aggressive. Details on the origin of montmorillonite and
beidellite are presented on the pages indicated:
Soils: pages 153-161
Lacustrine: pages 271, 273
Marine: pages 334-338, 356, 360-361, 366-379, 400-407
Sandstones: page 551
Nontronite commonly forms from basaltic material under weathering. lacustrin
or low temperature hydrothermal conditions. The warm waters are effective in
leaching the Fe from basalt. Recent studies indicate hydrothermal nontronites are
abundant in the ocean ridge areas and in adjacent deep-water basins. Nontronite
also forms from Fe-silicates and oxides. For more information see the pages
indicated :
Soil: page 156
Lacustrine: page 273
Marine: pages 354-371
54

Table 2-12
Chemical Analyses and Structural Formulas of Some Iron-Bearing Dioctahedral Montmorillonites and
Nontronites.
1 2 3 4
SiOz 39.52 41.88 39.92 51.6
A1203 18.48 11.90 5.37 0.05
Fe,O, 12.60 26.20 29.46 -
FeO - - 0.28 27.7
MgO 3.27 0.10 0.93 3.2
CaO 2.72 0.67 2.46 0.30
Na,O 0.10 - - 0.07
K2O 0.50 0.52 - 2.08
Ti02 1.88 - 0.08 0.01
H,O' 12.10 1.67 7.00
H20- 9.04 10.78 14.38
Total 100.21 99.72 99.88
Octahedral
Al 1.oo 0.51 0.03
Fe3+ 0.72 1.58 2.02 1.59
Fe2+ - - ~

Mg 0.34 0.27 0.31


x
~

2.06 2.10 2.07 1.90


Tetrahedral
Fe3' 0.11
A1 0.72 0.63 0.50 -
Si 3.28 3.37 3.50 3.89
Interlayer
Ca - 0.35 0.16
K - - 0.23
Na - 0.02 0.03
Mn - 0.06
1. Alietti, 1960: Hydrothermal alteration of basaltic tuff.
2. Ross and Hendricks, 1945: No. 66 Alteration zone in gneiss, Spruce Pine, North Carolina; Analyst:
L.T. Richardson.
3. Kerr, 1950: Alteration of basalt, Manito, Washington; Analyst: W. C. Bowden.
4. McMurtry et ul., 1983: Marine hydrothermal muds from Galapagos mounds.

Hisingerite is an Fe-rich, poorly crystalline material that is found in hydrother-


mally altered basalts on the continent and in the deep sea. X-ray reflections are
broad and the structure is in doubt. It appears to have a 2:l structure and may be
the precursor of nontronite (Brigatti, 1981; Shayan, 1984).
Trioctahedral Smectites: These are the Mg-rich smectites: stevensite, hectorite,
and saponite (Table 2-13 and 2-14). Stevensite is essentially a pure Mg silicate
similar to talc. The primary difference is that stevensite has a layer charge created
primarily by vacant positions in the octahedral layer. The layer charge and CEC
tend to be less than those of the other smectites.
55

Table 2-13
Chemical Analyses and Structural Formulas of Stevensite (1.2) and Hectorite (3) Samples.

1 2 3 1 2 3
SiO, 57.30 55.02 55.17 Octahedral
A120, None 1.12 0.33 Al - 0.08 0.02
FeA 0.32 ~ 0.12 Fe3+ 0.02 - -
FeO None 0.70 Tr Fez+ - 0.05 -
MnO 0.21 - Tr Mg 2.88 2.57 2.65
MgO 27.47 24.89 24.51 Mn 0.02 - -
CaO 0.97 0.54 0.90 Li 0.10 0.33
x
~

NazO 0.03 0.94 2.20 2.92 2.80 3.00


K 10 0.03 0.43 0.08 Tetrahedral
Li - 0.36 1.14 Al 0.00 0.02 0.00
TiO, - 0.08 0.01 Si 4.00 3.98 4.00
coz - 0.30 0.63 Interlayer Cations
F None 3.22 4.75 Ca 0.30 0.04 0.04
HZ0' 7.17 6.42 2.84 Mg 0.04 ~

H,O- 6.69 1.66 8.93 Na 0.13 0.28


K 0.04 0.01
Total 100.19 101.68 101.87 C.E.C.
meq/100g 36.00 75.1 115.00
Layer Charge 0.14 0.44 0.33
* Includes 0.21 CI. 0.05 PzO,.
Faust and Murata. 1953: Hydrothermal. pseudomorphous after pectolite. Springfield, New Jersey:
Analyst: K.J. Murata.
Faust et d.. 1959: Authigenic lake deposits associated with marl. Eastern Morocco: Analyst: J.J.
Fahey.
Ames and Goldrich, 1958: Hot springs alteration of zeolite in alkaline lake, Hector, California;
Analyst: S.S. Goldrich.

Hectorite is characterized by the presence of Li' substituting for Mg2+ in the


octahedral layer. Only one sample (No. 3) has been reported that contains apprecia-
ble Li.
Whereas stevensite and hectorite have little or no tetrahedral Al, the more
abundant saponite contains appreciable tetrahedral Al. Tetrahedral compositions
range from Si3.,5Alo.25to Si3,10Al,).90.When the tetrahedral charge is larger than
about 0.50. the excess negative charge is neutralized by the substitution of R'+,
largely Fe3+,in the octahedral sheet. Some samples contain large amounts of Fe'+,
nearly equal in amount to Mg. The octahedral sheet commonly has a positive
charge.
The Mg-smectites obviously crystallize in high Mg environments. These include
hydrothermal alteration of basaltic material in both marine and continental environ-
ments. Hydrothermal saponite is relatively abundant in the ocean ridge areas. but
apparently less abundant than the F e smectites. Most Mg smectites formed in
evaporitic lacustrine environments where both Mg and Si were concentrated during
56

Table 2-14
Chemical Analyses and Structural Formulas of Some Saponites.
1 2 3 4
SiO, 50.01 43.62 39.68 40.44
A1 2 0 3 3.89 5.50 3.93 14.52
Fe20, 0.21 0.66 19.82 1.85
FeO - - 1.12 2.37
MnO - 0.06 0.19 0.15
MgO 25.61 24.32 11.21 21.11
CaO 1.31 2.85 2.37 2.32
Na 2O - 0.08 - 0.42
K2 0 - 0.04 - tr
TiO, 0.04 0.00 0.37 -
CO, - - - -
H 20t 12.02 5.48 6.16 6.31
H,O- 7.28 17.42 15.11 10.65
Total 100.37 100.03 99.96 100.14

Octahedral
Al 0.04 0.02 0.04 0.38
Fe3+ 0.01 0.04 0.00 0.1 1
Fez+ ~ - 1.45 0.15
Mg 2.85 2.91 1.52 2.40
Mn - - - 0.01
c 2.90 2.97 3.01 3.05
Tetrahedral
Al 0.30 0.50 0.38 0.92
Si 3.70 3.50 3.62 3.08
Layer Charge 0.45 0.50 0.32 0.44
1. Cahoon, 1954: Hydrothermal alteration of dolomitic limestone, Milford. Utah; Analyst: W. Savour-
nin.
2. Mackenzie, 1957: In vesicles in basalt, Allt Ribhein, Skye; Analyst: J. B. Craig.
3. Sudo, 1954: Authigenic clay matrix in Tertiary iron sand bed, Moniwa. Japan.
4. Mongiori and Morandi, 1970: Hydrothermal alteration of poligenic breccia, Rossewa. Italy.

evaporation and the pH was relatively high. For additional information see the
following pages:
Lacustrine: page 272
Marine: pages 361, 363
Evaporites: page 410
Laboratory synthesis of smectites (Harder, 1972) has provided some general rules
for their formation. At low temperatures (3-20°C) smectite precipitates only in
solutions under-saturated in silica. If the silica concentration is high, the precipitates
remain amorphous. A pH greater than 7 is required. An inverse and apparently
linear relation exists between the Mg2+concentration and the pH with regard to the
precipitation of smectites. Only 10 ppm Mg2+ are required in the solution to form
smectites at pH 10, whereas at pH 7, 100,000 ppm Mg’+ are necessary. If the Mg’ ’
content is lower at any given pH value the precipitate remains amorphous. How-
ever, it seems plausible that the amorphous material could crystallize into smectite
once the Mg concentration of the fluid increased.
A number of rare smectites have been described:
Sauconite = Zn2+predominant in the octahedral layer (Ross, 1946).
Pimelite = Ni2+ fills the octahedral layer, similar to stevensite (Spangenberg, 1938).
Medmontite = Cu2+ predominant in the octahedral layer (Tchukhrov and Anosov,
1950).

MIXED-LAYER ILLITE/SMECTITE (I/S)

Mixed-layer or interstratified physils are those in which individual crystals are


composed of basic unit layers of two or more types. It is quite probable that the
great majority of physil minerals are composed of interstratified layers of differing
composition. In most instances these differences are not detected by routine
methods of analysis. From a practical standpoint, physils are classified as inter-
stratified when the layers are sufficiently different in character and sufficiently
abundant that the presence of the two or more layer types can be established by
x-ray analysis.
The layers may be regularly or randomly interstratified. The latter are by far the
most common and are probably the second most abundant clay physil, following
“illite” (which in most cases is a mixed-layer physil). Regular interstratified physils
have a definite periodicity and some are given specific mineral names (rectorite,
tosudite, and corrensite). The randomly interstratified physils are described in terms
of the type and proportion of the two or more types of layers. Many of them exhibit
some degree of regular interlayering.
The two contiguous layers of a regular mixed-layer physil act as a single crystal
and produce a 001 x-ray reflection which is equal to the total thickness of the two
layers, i.e., 10 A illite layer and 14 A chlorite layer = 24 A reflection. Other basal
reflections are regular submultiples of this distance.
Random mixed-layer physils produce a non-integral series of x-ray reflections.
Each reflection occurs between adjacent 001 reflections from the two different
layers and is positioned according to the proportion of the two types of layers (Fig.
2-18). Methods of calculating the expected 001 positions and intensities are dis-
cussed by MacEwan et al. (1961) and Reynolds (1980). In addition to the spacing of
the two layers, the proportion of two layers, the structure factors, the size of the
crystallites, and the tendency towards regularity influence the position and intensity
of the reflections. Mixed-layer illite/smectite (1,’s) is by far the most abundant
mixed-layer physil.

X-ray
Reynolds and Hower (1970) and Reynolds (1980) calculated a large number of
profiles for mixed-layer illite/smectites. I/S physils containing less than 40 to 50%
58

10.3

2 01

1) K+ - 12.4 A

Fig. 2-18. Observed diffractometer traces for a 10% montrnorillonite-90% illite mixed-layer system in
each of four states of niontmorillonite expansion. Under each trace the vertical lines show where 001
reflections would occur from pure phases, short lines for expanded layers, long lines f o r mica layers.
Thick horizontal lines show regions where reflections are to he expected from mixed-layer systems. From
Weaver, 1956. Copyright 1956 Miner. SOC.Amer.

smectite are mostly randomly interstratified. As the proportion of illite layers


increase, an IS ordering develops. These IS units are randomly or regularly inter-
stratified with illite layers. As the proportion of smectite layers decreases to about
15%, an IS11 ordering develops. Srodon (1980, 1984) and Tomita and Takahashi
(1985) have developed methods which apparently give more precise information,
59

particularly about the nature of the ordering. Srodoh’s method involves the use of a
high angle reflection, between 33 and 35”29. which eliminates the interference from
reflections from discrete illite, which is commonly a problem in the low-angle
region.
The mixed-layer I/S physils are also referred to in terms of Reichwite numbers:
RO = random, R1 = IS, R2 = 11.5, and R3 = IS11 (Jagodzinski, 1949). Srodoh and
Eberl (1984) suggest that 11s ordering does not exist or is rare.
I/S physils have been considered to consist of packets of approximately 2 to 15
layers that are either randomly or regularly interstratified. An alternative explana-
tion has been proposed by McHardy et al. (1982) and Nadeau et al. (1984).
Basically they found that particle thicknesses observed in the TEM were consider-
ably thinner than those calculated from x-ray peaks; they proposed that I/S physils
are mechanical mixtures. TEM analyses showed that pure smectite is composed of

ILLITE-SMECTITE LAYER SEQUENCE


INTERSTRATIFIED INTERPARTICLE
MODEL MODEL

LAYERS PART1CLE
SM E CT I TE
SY ECTITE 1nm
SM ECT I TE
1nm
SM E CT I TE

ILLITE ILLITE
2nm
SMECTITE

ILLITE
ILLITE
ILLITE 3nm

SMECTITE L -

ILLITE

ILLITE ILLITE
4nm
ILLITE

Fig. 2-19. Diagrammatic representation of the one-dimensional structure of interstratified illite/smectite


layer sequence showing smectite and illite layers as defined by the interparticle model 1 nm = 10 A). Dark
bands represent tetrahedral sheets; stippled bands represent octahedral sheets of the silicate layers: K’a
are planes of nonexchangeable cations (predominantly potassium), and E + nH,O + 0 exchangeable
cations and interlayer water k organic molecules such as ethylene glycol or glycerol (not to scale). Note
that a single 2:1 silicate layer is identical to an elementary smectite particle in the interparticle model but
not identical to a smectite layer in the interstratified model. in which the octahedral sheets are taken as
the layer boundaries. After Nadeau and Bain. 1986. Copyright 1986 The Clay Min. Soc.
60

SIEMENS D-500
200

17.56
L 1 /I A
26.68
34.90
45.26
A
111
6%S

... ,luL8 26.49


44.00
- 1 1 6 47.20 41’1~1
ISlA

Two-Theta (Degreer)
Fig. 2-20. X-ray powder diffraction patterns of I/S encompassing range of expandabilities Patterns are
of oriented, ethylene glycol-solvated preparations using CuKa radiation. Peak positions and sample
identification are given in the figure. ‘%S= percentage of smectite layers (expandability); RO = random
interstratification; R1, R3, = types of ordered interstratifications. Some A values are: 5.24 = 16.86 A.
6.44=13.73A. 9.73=9.090A, 7.14=12.21 A, 9.28=9.530A, 8 . 9 6 = 9 . 8 6 9 A . From Srodoh et al.. 1986.
Copyright 1986 The Clay Min. Soc.

particles approximately 10 A
thick, equivalent to one 2:l layer. Ordered I/S is
composed of particles that are approximately 20 A
thick; these particles are
apparently two 2 : l layers bonded by K ions. When sedimented on a glass slide.
“fundamental” particles stack on top of each other. Water (or glycol, etc.) layers
occur between the fundamental particles. The x-rays see an ordered IS physil.
Similarly, ordered IS11 consists of thicker illite crystals that contain four 2:l layers
bound by three planes of K, forming illite crystals about 40 A
thick. Water layers
develop between these crystals, creating a “mixed-layer’’ physil. Illites having no
expansive aspect have crystal thicknesses that average 70 A. The difference between
the interstratified and interparticle models is illustrated in Fig. 2-1 9.
Fig. 2-18 presents a simplified version of how the x-ray diffraction peaks of an
I/S physil vary when the thickness of the smectite layers are varied. Fig. 2-20
contains x-ray patterns of a series of I/S physils where the proportion of smectite
layers ranges from 88 to 6%. Reynolds (1980) constructed a series of calculated
x-ray patterns for a variety of I/S physils. Table 2-15 shows the shift in I/S peak
positions as a function of the I/S ratio. Srodoh (1980) developed a more precise
method for determining the proportion of illite and smectite layers (Fig. 2-21).
Basically, he uses higher angle peaks which can be measured more precisely and also
takes into account variations in thickness of the glycolated smectite layers. Later
(Srodon, 1984) he described a method of identifying I/S physils in the presence of
discrete illite. A simple technique developed by Watanabe (1981) and modified by
61

Table 2-1 5
Relations Between Compositions and Peak Positionsfor some Common Interstratifications: Spacings in
A. After Reynolds, 1980.
S illite lllite - (glycol) montmorillonite
Random Ordered
d d d d
0 8.52 5.62 8.52 5.62
10 8.60 5.59 8.60 5.59
20 8.67 5.57 8.71 5.55
30 8.76 5.54 8.82 5.50
40 8.90 5.50 8.93 5.44
50 9.06 5.44 9.03 5.39
60 9.26 5.37 9.21 5.34
70 9.52 5.28 9.40 5.28
80 9.83 5.16 9.66 5.22
90 10.00 5.07 9.99 5.09
100 10.16 5.01 10.16 5.01

Srodon and Eberl (1984) can be used to determine the I/S ratio and the degree of
ordering (Fig. 2-22).
As the proportion of illite layers in I/S increases, the - 10 A
peak (non-e.g.)
becomes sharper and decreases in width. Weaver (1960) proposed the "sharpness
ratio" (S.R.), later referred to as the Weaver Index (W.I.), to quantitatively describe
this phenomenon. The S.R. is the ratio of the height of the l0A peak to the height of
A
the peak flank at 10.5 (S.R. = 10 A/l0.5 A). Kubler (1964) proposed a similar
measure, crystallinity index (C.I.) or Kubler Index (K.I.), which is the 10 A peak

I I
2 3 4 5 6. 28 42 43 44 459 21a
Adz

Fig. 2-21. The plot for measuring the smectite:illite ratio based on the angular distance d. hetween
reflections in 42" - 48'2 region. The reflection between 42" and 45"2 is used to select the proper
thickness of the ethylene glycol complex for smectite-dominated compositions. Dashed parts o f o f the
curves represent the composition range beyond which Ad, cannot he measured because o f merging of
analytical reflection. From Srodoh. 1980. Copyright The Clay Miner. Soc.
62

I I I I I

0 1 2 3 4 5
A 28 I (DEGREE)
Fig. 2-22. Modified Watanabe (1981) diagrams showing how the distances between the first three I/S
x-ray peaks can be used to identify the type of ordering and estimating the ratio. g = 0 = random:
g = 1 = IS; g = 3 = ISII. From Srodoh and Eberl. 1984. Reviews Min., 13. 495--517. Copyright 19x3
Miner. Soc. Amer.

width at half-height, in mm or degrees 20. The W.I. increases and K.I. decreases
with increasing grade of diagenesis-metamorphism. For more details, see Chapter
VII.

Chemistry
I/S is the most abundant clay physil. In addition to the obvious I/S minerals. a
large portion of the “illites” and “smectites” are actually I/S physils. The composi-
tion of most 1/S is determined by the proportion of the two types of layers.
Commonly, but not always, I/S physils are formed during burial diagenesis where
montmorillonite is converted to illite as the temperature increases. Similar changes
occur in many hydrothermal deposits. With increasing temperature the amounts of
tetrahedral Al, interlayer K, and average layer charge increase. Si decreases.
Octahedral Fe and Mg may or may not decrease. However, I/S physils with varying
ratios can form during weathering and in alkaline lakes at normal surface tempera-
tures. I/S physils also form authigenically in sandstones, during burial. These may
or may not be temperature dependent. Eberl et al., (1986) and others have shown
that, in the laboratory, when smectite is saturated with K and alternately wet and
dried, up to 70% of the expanded layers can contract to form illite-like layers. The
same process may operate in soils.
Table 2-16 lists the chemical analyses and structural formula for a number of I/S
physils in bentonites (1 through 7). Most reliable I/S analyses are of bentonites o r
63

Table 2-16
Chemical Composition of I/S.
1 2 3 4 5 6 7 8
~ ~

Si02 53.1 54.8 51.6 52.58 60.01 52.44 56.4 56.19


Al ?O, 19.2 20.7 23.6 27.04 23.06 26.38 22.1 25.92
Fez(), 3.34 2.89 3.57 0.72 2.10 0.31 1.7 3.97
FeO 0.24 0.73 0.00
MgO 3.04 2.92 2.06 2.75 4.40 3.57 5.0 I .7x
CaO 0.69 0.91 0.1 1 0.24 0.22 0.66 0.4 1.63
Na,O 1.49 1.61 0.92 0.05 0.17 0.16 1.7 0.17
K ?O 1.51 3.61 6.24 8.47 2.65 7.85 7.3 4.54
TiO? 0.05 0.07 0.19 0.33 0.22 0.44 0.10
HlO+ 6.61 6.91 4.78 LO1
H2O- 2.03 3.07 6.93 5.80
Total 101.01 100.47 99.76 101.53 100.00

Structural Formula
Octahedral
Al 1.47 1.46 1.55 1.68 1.50 1.62 1.46 1.60
Fe'+ 0.18 0.15 0.19 0.03 0.10 0.02 0.09 0.19
Fez+ 0.01 0.04
Mg 0.33 0.30 0.21 0.2~ 0.38 0.36 0.49 0.17
Tetrahedral
A1 0.16 0.22 0.39 0.47 0.17 0.46 0.26 0.37
Si 3.84 3.78 3.61 3.53 3.83 3.54 3.74 3.63
Interlayer
Ca 0.05 0.07 0.01 0.01 0.15 0.05 0.1 1
Na 0.31 0.32 0.18 0.01 0.01 0.02 0.02
K 0.14 0.32 0.56 0.72 0.22 0.68 0.62 0.67
Layer Charge 0.55 0.79 0.75 0.76 0.53 0.82 0.62 0.66
% Expanded
Layers 76 45 20 6 60 10 20 51

Ordering R R R IS11 I/S R


1 through 4 recalculated to 100% of ignited material
8 preheated at 400°C.

1, 2. 3. Srodoh and Eberl. 1984: Carboniferous hentonites. Upper Silesia, Poland.


4. Srodon and Eherl. 1984: Silurian hentonites. Wenlock, Pentre Beirdd. Powys.
5. Bystriim, 1956: Ordovician bentonite, Kinnekulle. Sweden.
6. Weaver, 1953: Ordovician bentonite. Salina. Pennsylvania. U.S.A.
7. Huff and Turkmenoglu, 1981: Ordovician bentonite, Kentucky. U.S.A.
8. Foscolos and Kodama, 1974: Lower Cretaceous shale. British Columbia. Canada.

hydrothermal deposits from which relatively pure concentrates can be obtained. I t is


difficult to obtain pure I/S from shales, but analyses indicate the compositions are
similar to those of the bentonite I/S. In samples 1 through 6 the proportion of illite
64

Fig. 2-23. Plot of total layer charge. tetrahedral charge and octahedral charge vs. expandability. Silurian
bentonites. Wales; Cretaceous bentonite from Montana disturbed belt, U.S.A.: Upper Carboniferous
bentonites, Upper Silesia, Poland; Hydrothermal, Akita Prefecture. Japan. Modified from Srodoh and
Eberl. 1984.

layers is believed to be related to burial temperature, with the number of illite layers
increasing with temperature.
The only consistent compositional trend that parallels the increase in illite layers
is an increase in K and tetrahedral Al. Total layer charge does not increase
systematically as the proportion of illite layers increases. Samples 2, 3, 4, and 6 have
a total layer charge ranging from 0.75 to 0.82, yet the number of smectite layers
ranges from 6 to 45%. However, the relative amount of tetrahedral charge increases
as the number of illite layers increases. This suggests the tetrahedral charge, closer
to the interlayer K, is more effective than the octahedral charge in fixing K .
However. overall the total layer charge of I/S physils increases from 0.4-0.5 to
0.7-0.8 as the proportion of illite layers increases from about 10%to near 100%.The
relation of total layer charge to the percent illite layers varies for various strati-
graphic sections (Fig. 2-23), in part this is related to the variations in the amount of
octahedral and tetrahedral charge. In general the octahedral charge remains rela-
tively constant and the tetrahedral charge increases as the total layer charge
increases (Fig. 2-23). The composition and layer charge of the original smectite
apparently influences the rate of formation of illite layers.
Sample 7 is believed to have formed at a temperature lower than that of I/S
physils of equivalent I/S ratio (Huff and Turkmenoglu, 1981). This sample has a
65

>0.55K/Layer, Perhaps Variable

-
E
v)
so 1 k,Random

' 40 Ordered \

2o
lo 1 I I I I I I
0.1 0.2 0.3 0.4 0.5 0.6 0.7 1 3

Fig. 2-24. Plot o f fixed potassium vs. expandability illustrating proposed interpretation of illitization
process. From Srodoh et al.. 1986. Copyright 1986 The Clay Min. Soc.

relatively high Mg content and the octahedral charge is larger than the tetrahedral
charge. The authors suggest that the high charge necessary to produce illite layers
was a product of parent material and pore fluid composition rather than thermal
history. Sample 8 is a shale sample from the Cretaceous Buckinghorse Formation of
Canada. Some of the expandable layers are believed to be vermiculite.
Weaver and Pollard (1973) noted that if all the K in I/S physils is assigned to the
illite layers, the amount of K per illite layer increases as the proportion of illite
layers increases. Thus, when less than half the I/S layers are illite, the concentration
of K is 0.55 per illite layer and the concentration of K increases as the proportion of
illite layers increases. Srodoh et ul. (1984), using additional data, have refined this
trend (Fig. 2-24) and shown that 1/S containing more than 50% smectite layers
(random) contains approximately 0.55 K, and I/S with less than 50% smectite
layers (ordered) contains about 1.0 K in the illite layers. By extrapolation they
determined that the end-member illite derived from the alteration of smectite should
have an average of 0.75 K per illite layer, a total charge of about -0.8 and a cation
exchange capacity of 15 meq/100 gm.
Srodon et al. (1986) suggest that as a few individual illite layers form in the
original smectite, the K ions in the individual illite layers are separated by layers
containing interlayer water (smectite). Thus, there is little K'-K ' repulsion, and a
charge of about 0.55 will cause collapse of the interlayers. As illite layers become
more abundant, K f - K + repulsion becomes more likely and a higher charge is
required to cause layer contraction. The typical illite is presumed to be a mixed-layer
Io,ss/I,,,~; the ideal illite is I,.,, (phengite).
66

The cation exchange capacity (CEC) and surface area of I/S is proportional to
the percentage of smectite layers. CEC values Tor the smectite layers are on the
order of 100 meq/100 g (Weaver and Pollard, 1973; Srodon ef (11.. 1986).

Regulur Mixed-lu.ver Phjsils


Regular mixed-layer physils which consist of approximately 50% 10 A layers and
50% expanded layers are called rectorite or sometimes allevardite. These physils
characteristically have a well developed superlattice x-ray peak in the 23 to 27 A A
A
range, depending on the nature of the inter-layer material. The 10 layers can be

Table 2-17
Composition of Regular Mixed-Layer Physils

1 2 3
SiO, 54.11 49.93 51.98
Al ?O, 40.3x 25.97 39.35
Fez(), 0.15 2.01 1.44
FeO - 0.30 -
MgO 0.78 2.77 0.25
CaO 0.52 1.11 3.46
Na,O 337 0.09 2.43
K ,o 0.29 3.46 1.ox
TiO? 0.01 0.42 ~

H20+ 7.45
H20- 6.49
Total 100.24 100.00 101.02

Structural Formula

Octahedral
Al 2.01 1.64 1.94
FcIi 0.01 0.11 0.07
Fez+ 0.02 -

Mg 0.07 0.29 0.02


Tetrahedral
Al 0.80 0.50 0.86
Si 3.20 3.50 3.14
I nterlayer
Ca 0.07 0.17 0.22
Na 0.44 0.01 0.28
K 0.02 0.3 1 0.08

Layer Charge 0.60 0.63 0.79

Ordering IS IS IS
1 ignited weight basis. 3 ignited weight basis. 0.03 LizO.
1. Kodama. 1966: Fort Sandeman District. Buluchistan. Pakistan.
2. Rateyev et ul.? 1969: Samarskaya. Luka. U.S.S.R.
3. Nishiyama and Shirnoda. 1981: Tooho Mine, Aichi Prefecture, Japan.
paragonite (Na), illite-muscovite (K), or margarite (Ca). K, Na, o r Ca can be the
predominant interlayer cation (Table 2-1 7). Rectorite is not very abundant and
commonly is a hydrothermal mineral.
I/S forms by weathering of feldspars and volcanic rocks; i t forms authigenically
in sandstone pores and perhaps in marine and alkaline lacustrine environments: it
forms during burial metamorphism (from smectite). and under hydrothermal condi-
tions.
Details on the origin of I/S are given on the following pages:
Soils: pages 174-181
Lacustrine: pages 272-274
Diagenesis: pages 41 8-442
Metamorphism: pages 454-51 3
Sandstones: pages 533-536, 551-560

CHLORITE A N D CHLORITICS

Chlorite

A
The fundamental 14 unit layer of trioctahedral chlorite consists o f negatively
charged 2:l mica-like layers that alternate regularly with positively charged hy-
droxyl sheets (Fig. 2-25). The negative charge is created largely by the substitution
of Al’* in the tetrahedral sheet. The main cations in the 2:l octahedral sheet and
the octahedral interlayer sheet are Mg and Fe”. Al and Fe3+ substitute for the
divalent cations and produce a positive charge. In addition to the electrostatic
bonding, hydrogen bonds exist between the interlayer hydroxyls and the basal
oxygens of the 2:l layers above and below. Dioctahedral chlorites and di, trioc-
tahedral chlorites (dioctahedral 2:l layer and trioctahedral interlayer sheet) are
much less abundant than trioctahedral chlorites.. Many chlorites formed under low
temperature conditions have incomplete interlayer sheets.

Structure
The 2 : l layer is similar to that o f mica and has an a/3 stagger between the
tetrahedral sheets. In the interlayer sheet the octahedral cations can occur in two
different positions. I and I1 (Fig. 2-26). A given interlayer sheet, I or 11, can be
placed on a 2:l layer in six ways that will provide hydrogen bonds o f similar length
between the hydroxyl and oxygen surfaces. The six positions can be divided into
two sets of three positions. a and b. The repeat 2:1 layer can also be positioned six
ways with respect to the interlayer sheet. There can be 12 different one-layer
polytypes, designated according to the type of interlayer sheet ( I or 11). the position
(a or b) of the interlayer sheet on the initial 2:l layer, and the position of the
repeating 2:l layer on the interlayer sheet (Bailey, 1980).
68

c 8inB
* 14.21

! N0.l 0

0 0

Fig. 2-25. Idealized chlorite structure (type Ilb). (a) 010 projection. The anion planes are numbered. ( b )
001 projection of initial talc sheet. omitting tetrahedral cations. The arbitrary initial axes are not
necessarily the same as those of the entire structure. From Bailey and Brown. 1962. Amer. Miner., 47.
819-870. Copyright 1986 Miner. SOC.Amer.

Of the various possible polytypes only four have been found: Ilb, /? = 97", Ib,
/? = 90", Ib, p = 97", and Ia, /?=97". Bailey and Brown (1962) established that the
IIb polytype is the stable high temperature chlorite and is the type normally present
in metamorphic rocks. The authigenic, low temperature chlorites are the type I
polytype. Based on a study of chlorites in sandstones, Hayes (1970) suggested the
following stability sequence (from low to high temperature):
Ib, -+ Ib ( p = 97") + Ib (/?= 90") + IIb
Ib, is disordered chlorite which lacks an h01 diffraction band in the 2.4-2.5 A
region. There is little relation between chemical composition and polytype.

X -RU-V
Chlorites are characterized (oriented slide) by an integral sequence of 001 lines
A
based on a repeat of d = 14 (Table 2-18). 001 values range from approximately
14.1 to 14.4 A;
these thickness values are controlled by the chemical composition.
The layers do not expand when treated with ethylene glycol and contract only a few
tenths of an angstrom when heated to 500-700°C.
When heated to about 500°C the 001 reflection increases by a factor of 2 to 5
times and the higher-order 001 reflections decrease in intensity. This reaction is
69

x Brucite h y d r o x y l s f r o m no. 5 onion p l a n e


0 Talc o x y g e n 6 f r o m no. 4 anion p l a n e
I Brucite octahedral cations, brucite s h e e t I
II B r u c i t e o c t a h e d r a l c a t i o n s , b r u c i t e s h e e t II
Fig. 2-26. Superimposition of brucite sheet upon initial talc sheet in 001 projection. ( a ) Brucite cations
may occupy either sites I or I1 above no. 5 OH plane with no. 6 OH plane emplaced t o provide
octahedral coordination for the occupied sites. (h) and (c) Two methods (a and b) o f positioning hrucite
sheet I upon initial talc sheet to provide hydrogen bonds between no. 4 0 and no. 5 OH planes. The two
resultant assemblages. designated la and Ih. are interrelated by shifts of the brucite sheet by 1/3 a,, in
the directions indicated by the small arrows. (d) and (e) Talc-brucite assemblages Ila and Ilb. From
Brown and Bailey, 1962. Amer. Miner., 47, 819-870. Copyright 1962 Miner. Soc. Amer.

characteristic of chlorites. Other 14 A minerals, smectites and vermiculites, contract


to near 10 A when heated. When kaolinite and chlorite are both present, interpreta-
tion can be a problem. As chlorite is much more soluble than kaolinite, a variety of
chemical treatments has been devised to distinguish between the two (Brown and
Brindley, 1980); however, the two physils can usually be identified from the x-ray
A
pattern of untreated samples. Chlorite has a 4.75 (003) peak which kaolinite does
not have. Further, it is usually possible, when both physils are present, to resolve the
A
3.58 A (002 kaolinite) and 3.50 to 3.55 (004 chlorite) peaks.
In order to identify the chlorite polytype it is necessary to x-ray random powder
mounts. Due to the interference of peaks from other minerals it is necessary to x-ray
70

TABLE 2-18
X-ray powder data for chlorites: spacings in A. From Bailey. 1981

IIh Ib la
Ilk1 d I d I (1 I
00 1 14.15 8 14.4 6 14.2 6
002 7.05 10 7.15 10 7.10 10
003 4.72 6 4.79 4 4.73 4
02: 11 4.60 2 4.63 4 4.63 1
004 3.54 10 3.59 7 3.55 8
005 2.83 4 2.87 2; 2.84 3
20T 2.66 1; 2.68 4 -

200 - 2.66 4
I
203 2.59 5 2.61 1; 2.59
1
201 2.54 8 2.55 2.55 1
20? 2.44 7 2.475 6 -
202 2.38 4 2.39 1 2.395 6
204 2.255 4 2.29 1 2.27 1
- I
203 2.20 -
I
205 2.06 1: 2.105 2 2.07
I
007 ~
2.045 -

204 2.00 6 2.01 1 2.01 3


I I
206 1.88 2; 1.91 1.89
205 1.82 2; - -
I
15: 24: 31 1.74 1 1.758 2 1.76
I
207 1.715 1.74 1 -

206 1.66 1; - 1.67 7 1,


-

208 1.565 3 - 1.57 1


060 1.538 7 1.548 6 1.549 6
062 1.503 2; 1.515 3 1.515 3
I
063 1.462 1.478 1 1.472 I
I
209 ~ - 1.439
I
0. 0. 10 1.434
1.414 1 1.420 3
064 1.420 1
208 1.392 2; - -

samples that are nearly pure chlorite. Bailey (1980) lists the characteristic diffraction
spacings for the four polytypes.
Chlorites formed at low temperatures commonly have incomplete or imperfect
brucitic or gibbsitic interlayer sheets or are interstratified with expandable smectite
or vermiculite layers (mixed-layer Ch/S or Ch/V). I t is believed that the incomplete
sheets consist of brucitic islands. Ethylene glycol can apparently move into the
inter-island space and expand the layers to 16 to 17 A. These chlorites, called
swelling chlorites. collapse more easily when heated than d o normal chlorites but
are much more “heat stable” than smectites and vermiculites (Martin Vivaldi and
MacEwan, 1960).
71

Reasonable estimates of the chemical composition of chlorites can often be made


from x-ray data. If the difference in scattering power and size of the various atoms
in the chlorite species is sufficiently large, the positions and intensities of x-ray
reflections are measurably affected. The b and c parameters are the most useful.
The 060 reflections of trioctahedral chlorite range from 1.53 to 1.55 Aand vary
linearly with compositional variations in the octahedral sheet. One formula for the b
parameter is (Radoslovich, 1962):

b = 9.23 A + 0.03 Fe2+(+0.0285)


Thus, the 060 value can be used to calculate the amount of Octahedral F e z + in
chlorites.
The basal spacing depends on the dimensions of the 2:l layer and the hydroxide
sheet and the forces holding them together. This bonding force is a function of the
amount of charge in the tetrahedral and octahedral sheets (particularly the inter-
layer sheet) and is therefore proportional to the amount of A1 substitution in the
tetrahedral sheet (Brindley, 1961). Brindley has given the following linear relation:

d(001) = 14.55 A - 0 . 2 9 ~

where x represents the amount of tetrahedral Al.


Graphs based on relative peak intensity have been constructed which can be used
to determine how the heavy atoms (Fe + Mn + Cr) are distributed between the two
octahedral sheets (Petruk, 1964).
Brindley has also described a method (after Brindley and Gillery, 1956) of
calculating the Fe content and the difference in composition of the two octahedral
sheets using the intensities of the first five peaks. The common presence of
vermiculite layers and faults in clay-sized chlorites makes the application of this
technique difficult. In this same article Brindley and Gillery show how variations in
basal intensities can be used to identify dioctahedral chlorites and chlorites with
dioctahedral mica-like layers and trioctahedral brucite sheets. The most diagnostic
feature of dioctahedral chlorites is the presence of a relatively strong 003 reflection.
Bailey (1975) reviewed the various x-ray techniques used to determine the
chemical composition of chlorites and concluded that tetrahedral A1 and octahedral
Fe can be estimated with an average error of about 10%. However, the standard
chlorites used were well crystallized and relatively coarse. Chlorites in shales are less
well crystallized, may contain some 7 A and/or expanded layers, and may contain
two or more varieties of chlorite.
A comparison of chlorite compositions calculated from x-ray data with those
obtained by chemical analysis of chlorites from a Paleozoic shale-slate sequence
showed there was very poor correlation between the two types of analyses (Weaver
el ul., 1984). Fe values calculated from X-ray data were high for most samples and
tetrahedral A1 was high for all samples. These results could be produced by the
A A
presence of some 7 layers interspersed with the 14 layers: TEM studies indicate
this is common. These analyses were made on 4 to 8 p m chlorites. Compositions
based on x-ray patterns of the < 2 pm chlorites are likely to be extremely
12

Mg Chlorite

V
a
0

cl J

Fig. 2-27. X-ray diffractometer patterns of Mg and Fe chlorite5 and chloritic shale

inaccurate. At best one should be able to determine that a chlorite is Fe-rich (low
intensity odd orders) or Mg-rich (first four 001 peaks similar intensity) (Fig. 2-27).

Chemistry
High temperature: Chlorites are hydrous silicates of magnesium, aluminum, and
iron. Most are trioctahedral, but dioctahedral and mixed dioctahedral-trioctahedral
types exist. Because of the wide range of ionic substitutions, a great many names
73

I THURINGITE I I I
EXPLANATION

Fe,O, < 4 percent

- -_ I

.I
~-
SHERIDANIIE,
I
I
a*:
’..I
-1 I
I I

CLINOCHLORE
A

200 220 240 260 280 300 3 20 340 360 380 400
FORMULA POSITIONS OCCUPIED BY L

Fig. 2-2X. Classification of chlorites based on the two principal types of ionic replacement. From Foster.
1952.

have been given to various members of the chlorite group. The structural determina-
tions, chemical analyses, and classification of chlorites are based largely on studies
of macroscopic chlorites. Chlorite is abundant as a clay-sized mineral, probably
comprising from 10 to 15% of the total physil suite. Some of the chlorite is detrital.
derived largely from metamorphic rocks, but much of it appears to be either
authigenic or diagenetic. In shales chlorite is usually coarser grained than illite.
The chemistry of the chlorites has been reviewed by Hey (1954), Foster (1962),
and Deer et al. (1962). Hey and Foster have presented classification schemes. All
chlorites have some Al substituting for Si which gives the tetrahedral sheets a net
negative charge. This charge is balanced by the substitution of A l and Fe” for Mg
and FeZf in the two octahedral sheets.
Foster (1962) devised a classification based on the composition of the tetrahedral
sheet and the relative Fe content of the octahedral sheet (Fig. 2-28). Most chlorites
tend to have a much higher tetrahedral Al content than the 2 : l physils. The
tetrahedral charge is balanced by the substitution of Al and F elt in the two
octahedral sheets. Foster assumed that the R7+cations were evenly distributed
between the two sheets, but structure analyses indicate that this is not necessarily
true. Foster found:
“... that octahedral Al is not often closely equivalent to tetrahedral Al, being
higher in about one-fourth of the formulas, and lower in about one-half. If lower.
some other octahedral trivalent cation, usually Fe7+,is present and proxies for Al
in providing sufficient octahedral positive charge to balance the tetrahedral
negative charge. Octahedral occupancy is close to 6.00 only if the sum of the
octahedral trivalent cations is approximately equal to tetrahedral Al. I f i t is
greater, as is most usual, octahedral occupancy is less than 6.00 formula positions
14

by an amount equal to about one-half the excess of octahedral trivalent cations


over tetrahedral Al, indicating that the excess octahedral trivalent cations replace
bivalent cations in the ratio of 2:3.”
Octahedral occupancy ranges from 5.46 to 6.05. R” octahedral occupancy ranges
from approximately 10 to 47% of the filled positions. with most of the values
ranging from 15 to 35%. In general, these values are similar to those for the other
physils although the larger values are larger than those found in any of the sheet
structures except some of the chamosites. These high values are not restricted to the
iron-rich chlorites as they are in the other physils. In some instances high R3+
occupancy values in the chlorites may indicate that one of the octahedral sheets is
dioctahedral rather than trioctahedral.
In most of the analyses reported by Foster, octahedral Al is dominant over
octahedral Fe” although in a few examples the reverse is true. I t was thought that
the Fe20, in chlorites was the result of secondary oxidation of FeO; however,
Foster found that i n many chlorites much of the Fe3+ was necessary to maintain a
charge balance.
The chlorite formulas, similar to biotite and trioctahedral physil formulas,
indicate that as the Fe”:Mg ratio increases. the amount of R3+ tends to increase.
Naturally occurring sheet silicates with a pure Fe” octahedral sheet apparently d o
not exist because of the inability of the tetrahedral sheets to adjust to such a large
octahedral sheet. Foster’s data show that the ratio of Fe2+:R7+ranges from 0.00 to
1.00, indicating a complete range of isomorphous substitution between Fe” and
Mg.
Experimental hydrothermal studies in a portion of the system MgO-A1203-Si0,-
H z O (Nelson and Roy, 1958) showed that a complete sequence of 14 Achlorites
could be synthesized ranging in composition from amesite (Mg, Al )( Si A1 )
,,
O,,,(OH),, to pennini te (Mg Al o.5 )(Si 3.5 Al ,).s)O,,,(OH)x. A certain minimum
amount of R7+substitution is necessary in order to provide sufficient layer charge t o
bind the various layers. The chlorite structure is not stable with more than one-third
of the octahedral positions filled with Al; if more Al is present, a three-phase
assemblage is produced, containing a chlorite of the amesite composition.
From the same compositions, but at lower temperatures (below 400-500°C). a 7
A structure of the kaolin type is developed. Nelson and Roy (1958) called these
materials septechlorites. It was not established whether the 7 A phase was metasta-
ble or not.
Trioctahedral chlorite occurs in all types of sedimentary rocks. I t is seldom the
dominant physil but probably occurs in more than 80% of sedimentary rock,
commonly in amounts < 30%. Much of the chlorite in shales is detrital. though i t is
not clear how much. It is also formed authigenically and diagenetically.
I t is difficult to obtain pure samples of chlorite from shales and few chemical
analyses are available. X-ray data can be used to obtain some estimate of the
composition, though the results are questionable.
The 001 reflections (14.25 to 14.05 A) of most chlorites in shales indicate that
they should have from 1.0 to 1.9 tetrahedral A1 per four tetrahedral positions. The
75

Table 2-19
Chemical Composition and Structural Formula [O,,,(OH),]of Some Low-Temperature Chloritez.

1 2 3 4 5 6
S102 23.13 30.7 31.X 35.63 29.5
A1 2 0 , 21.61 20.5 19.9 34.x7 52.0
- -
Fe20, 26.3 5.01 0.5
FeO 41.06 2.9 - 0.43 0.3
M go 4.75 31.2 9.3 X.63
CaO 0.04 ~ - 1.6
Na ~

0.24
K2 0 - 0.46
H - 14.15 13.9

Octahedral
Al 1.40 1.25 2.00 1.30 3.01 4.30
Fe’+ ~ ~

0.22 0.34 0.03


Fe?+ 3.82 0.23 2.00 1.95 - 0.02
Mg 0.79 4.44 1.36 2.29 1.1x
Ca 0.01
z: 6.02 5.92 5.36 5.79 4.53 4.35

Tetrahedral
Si 2.57 2.94 3.27 2.X9 3.26 2.70
Al 1.43 1.06 0.73 1.11 0.74 1.30

1. Iijima and Matsumoto, 1982: Chamosite from berthierine. Triassic coal beds, Utatsu. Japan.
2. Bodine and Standaert. 1977: Clinochlore. Silurian rock salt. Retsof. New York.
3. Weaver and Associates. 19x4: Chlorite porphyroblasts in incipient slate (250°C). Cambrian. Con-
asauga Formation. Georgia.
4. McDowell and Elders. 1980: Hydrothermal chlorite (250°C). Salton Sen geothermal field
5. Sudo and Sato, 1966: Di. dinctahedral chlorite. hydrothermal. Akita. Japan.
6. Caillere. 1962: Di. dioctahedral chlorite in bauxite deposits of Pyrenees Orientales.

tetrahedral Al in the majority of these chlorites falls within the narrower range o f
1.0 to 1.6 (14.25to 14.10 A). An average value would be close to 1.4.(Most of the
macrochlorites lie within a similar range.) Available 060 values are limited; however,
the relative intensities of the 001 reflections indicate that most shale chlorites have
from 25 to 67% (1.5to 4.0)of the octahedral positions filled with F e and that, in
many samples, there is more Fe in the interlayer hydroxide sheet than in the 2:l
layer. The Fe values are probably high. T h e presence of kaolinite and faults in the
chlorite tend to give anomalously high values for Fe.
LOWteniperurure ( < 300°C): Authigenic and diagenetic chlorites can be divided
into Fe-rich chamosite and Mg-rich clinochlore types with a wide range of inter-
mediate species. Table 2-19 lists some typical chemical analyses a n d structural
formulas. Unfortunately, it is difficult to determine from the older literature which
chamosites are chlorites (14 A)and which are berthierine (7 A). The chamosite in
Table 2-19 was formed at a burial temperature of approximately 160°C from
berthierine. The berthierine formed by replacing kaolinite in a coal swamp. Schoen
(1964) calculated from x-ray data that the secondary chlorite in the Silurian
16

ironstones had the following composition:


(MgAI),.A Fe)3.5(Si7.65 A1 1 35 )OIO (OH),
Mg-rich chlorites are characteristically found in evaporite deposits, such as
dolomite and halite. Under evaporitic, alkaline conditions the Mg concentration and
Mg/Ca ratio are sufficiently high that Mg-chlorite is the stable phase. Most of the
material formed under these conditions are not complete chlorites but mixed-layer
chlorites/ expanded or “chlorites” with incomplete interlayer sheets. Sample No. 2
is a well crystallized chlorite from the Silurian halite deposits of New York. It is
possible the “completeness” of this chlorite has been improved by burial diagenesis.
Chlorite also forms during burial diagenesis and metamorphism. During low-tem-
perature diagenesis chlorite apparently formed as a by-product” of the conversion

of montmorillonite to I/S (see Chapter VII). The amount of chlorite produced is


small and tends to be Fe-rich. Abundant chlorite is commonly generated during the
late stages of diagenesis of shales and the formation of slaty cleavage. Sample No. 3
-
is chlorite porphyroblasts from shales containing incipient slaty cleavage ( 250°C).
The Fe is slightly more abundant than the Mg. but with increasing temperature Fe
decreases and Mg increases until the ratio is about 1.0 in well developed slates
(Weaver et al., 1984). Sample No. 4 is a typical hydrothermal chlorite from the
Salton Sea geothermal area.
In di, dioctahedral chlorites both octahedral sheets are dioctahedral. One variety
is called donbassite. I t is not clear whether there are other varieties. Dioctahedral
chloritic material forms in soils when Al and Fe hydroxides precipitate in the
montmorillonite and vermiculite interlayer space; however, these layers are usually
poorly developed. Well crystallized dioctahedral chlorites are largely of hydrother-
mal origin but also form during weathering. In order to maintain electrical neutral-
ity the octahedral sheets must contain more than 4 A1 per O,,,(OH),. The reported
tetrahedral charges range from 0.60 to 0.80. Sample No. 5 is a typical hydrothermal
sample of di, dioctahedral chlorite; other analyses are given in Weaver and Pollard
(1973). Soil di, dioctahedral chlorites have a significantly higher amount o f tetra-
hedral Al - 1.20 to 1.30 (Sample No. 6) - than the hydrothermal variety.
Another type of chlorite consists of a dioctahedral 2:l layer and a trioctahedral
interlayer, di, trioctahedral chlorites. Physils of this type have been called sudoite
and cookeite (LiAI,)(Si,AI)O,,(OH),. Sudoites have been found primarily in
hydrothermal deposits, but also occur associated with Ordovician K-bentonite beds
(Weaver, 1959). They are the IIb structural type with a general composition near
(A], 7Mg2.3 NSi3 3Al0.7)OlI,(OH),
Eggleton and Bailey (1967) describe a sample in which the 2:l octahedral layer
contains 2.0 A1 and the trioctahedral interlayer has a composition o f
Mg2,3All,,(OH)6. In a sample from Japan the 2:1 octahedral sheet contains
(Mg,,Al, 7 ) and the interlayer sheet (Mg,3Al,.,) (Shirozu and Higashi, 1976).
For additional information on the origin and distribution of chlorite, see the
following pages:
Soils: pages 170- 174
Marine: pages 352-354. 363-366. 371
Evaporites : pages 407-410, 414
Diagenesis: pages 435-436, 443-446
Metamorphism: pages 462-469, 503-510
Sandstones: pages 533, 543-550, 560

Chloritic Physils

Chlorites and “would-be” chlorites are more complex and less well understood
than any of the other physils. This is due in part to the fact that chlorites have two
octahedral sheets and it is difficult to determine the composition and completeness
of the two sheets. The 2:l physils have an almost infinite range of compositions and
structural properties. 2:l layers can have a relatively high negative layer charge and
be bound together by a K + ion bridge (illite-mica). 2:l layers can have a low layer
charge (smectite) or high layer charge (vermiculite) and be separated by water
layers. 2:l layers with a wide range of layer charges can have Mg, Al, Fe hydroxide
compounds between the layers. These are the chlorite minerals. The interlayer
hydroxides can range from complete, well bonded interlayers (chlorite), to weakly
bonded interlayers, to interlayer hydroxy islands, to a random or regular distribu-
tion of interlayer material and water layers between 2:l layers.
These incomplete chlorites can form in soils by the degradation of chlorite or,
more commonly, by the precipitation of hydroxy A1 or Fe layers in the interlayer
space of smectites and vermiculites (frequently formed by weathering of illite-mica).
In alkaline lacustrine, evaporitic, and brackish environments hydroxy Mg interlayers
often precipitate in the inter-layer space of expanded physils. Perhaps the most
distinctive feature of the physils with incomplete hydroxy interlayers is that they d o
not contract to near 10 A when heated at several hundred degrees Celsius, but they
do partially collapse when heated at 400 to 500°C (Rich, 1968; Barnhisel, 1977).
Caillttre and HCnin (1949) and Slaughter and Milne (1958) were able to precipi-
tate interlayer A1 hydroxide sheets (as well as Mg and other hydroxides) between
montmorillonite layers forming chlorite-like physils. Hsu and Bates (1964) were able
to deposit hydroxy-aluminum polymers between layers of vermiculite as long as the
NaOH/AI ratio was less than 3. Carstea (1968) found that the formation of
hydroxy-Al interlayers in montmorillonite and vermiculite increased with increasing
temperature. Brindley and Kao (1980) systematically precipitated hydroxy-Al and
hydroxy-Mg with different OH values between montmorillonite layers. The OH/AI
ratios ranged from 2.17 to 2.98 and interlayer A1 from 0.35 to near 2 (dioctahedral);
the OH/Mg ratios ranged from 1.03 to 1.92 and interlayer Mg from 0.31 to near 3
(trioctahedral). Al and Mg nitrite solutions were added to suspensions of
montmorillonite and varying amounts of NaOH were added. Before the NaOH was
added, the interlayer material had a composition near [Al(OH)?]+and [Mg(OH)]+.
With the addition of NaOH, interlayer compositions moved progressively towards
AI(OH), and Mg(OH)2. The hydroxy-Mg products have less tendency to swell in
ethylene glycol and have greater thermal stability than the hydroxy-Al products. As
7x

MG

MG

MG

MG

ze DEGREES

Fig. 2-29. Magnesium hydroxide-montmorillonite complex: diffraction patterns of samples dried at


100°C showing effect o f increasing magnesium atoms per structural unit. From Slaughter and Milner,
1958. Reprinted with permission from Clays Clay Miner. 7th Conf. Copyright 1959 Pergamon Books.
Lid.

the interlayer hydroxy compounds become better developed there is less tendency
for the product to expand with ethylene glycol. When the interlayer Mg fills 2.50 of
the 3.0 interlayer positions, the 14.5 to 14.8 A layers no longer expand and d o not
contract when heated at 300°C. At lower Mg values a mixed-layer chloritic/
expandable layer product forms. These various products appear to be reasonably
representative of the variety of chloritic minerals formed under natural conditions.
Fig. 2-29 shows the gradual change in peak intensities as Mg hydroxide is
precipitated in a montmorillonite suspension (Slaughter and Milne, 1958).
Mixed-layer chlorite/smectite and chlorite/vermiculite form under weathering
(Al, Fe) and evaporitic (Mg) conditions. I t also forms hydrothermally and during
burial diagenesis. It can also be formed by stripping the interlayer hydroxy sheets
from chlorite during weathering. Both dioctahedral (from smectite) and trioc-
tahedral (chlorite and biotite) chloritic material forms in soils; those formed under
evaporitic, hydrothermal, and diagenetic conditions are largely trioctahedral (Mg).
The development of mixed-layer Ch/expanded layer rather than expanded layers
with scattered hydroxy islands is favored by higher temperatures.
Regularly interstratified 1:l Ch/S has been given mineral names. The dioc-
tahedral variety is called tosudite and the trioctahedral variety corrensite. Tosudite
is primarily a hydrothermal physil and commonly forms as an alteration product o f
volcanic rocks. In some instances I/S is the precursor physil (Ichikawa and
Shimoda, 1976). The octahedral sheets have a high Al content and most of the layer
charge originates in the octahedral layer. Similar dioctahedral chloritic material
19

develops in acid soils where hydroxy Al and Fe precipitate been some of the
expanded layers of smectite and stripped illite-micas. The regularity and crystallin-
ity is less well developed than for their hydrothermal counterparts (Weaver and
Pollard, 1973).
Corrensite was originally defined as a regular 1:1 interstratification of trioc-
tahedral (Mg) chlorite and swelling chlorite (Lippmann, 1954). The term is now
generally applied to regular 1 :1 chlorite/ smectite (usually saponite) and
chlorite/vermiculite physils. Corrensite (regular Ch/S) and Ch/S (random) are
relatively common physils and form under a wide variety of conditions. The
random, non-1:l variety is by far the most abundant. Ch/V is relatively rare. The
main requirement for the formation of Ch/S is the availability of abundant Mg and
a fairly closed system. For a compilation of chemical data. see Brigatti and Poppi
(1984).
Ch/S and corrensite are common in evaporite rocks (marls, dolomites, anhydrite,
and halite) (Lippmann, 1954; Millot, 1970; Peterson, 1961; Kopp and Fallis, 1974).
In many instances the mixed-layer phases form by hydroxy Mg precipitating
between layers of smectite or stripped illite. The chloritic physils formed under
evaporitic conditions tend to have a relatively high Mg and low Fe content (Table
2-20).
Ch/S and corrensite are widespread in basalts where they have formed both by
weathering (Sarkisyan and Kotelnikov, 1972) and by hydrothermal activity (Krist-
mannsdottir, 1978). In the latter instance they generally form in the general
temperature range of 100 to 250°C.Chloritic minerals formed from basalts char-
acteristically have a relatively high content of Fe (Table 2-20).
The abundant Ch/S in Ordovician carbonates probably formed from volcanic
ash deposited in mildly evaporitic (tidal flat) waters (p. 579). It also forms
diagenetically in sandstones, particularly i f volcanic material is present (Almon ef
al., 1976).
It is a relatively common weathering product of chlorite (Johnson, 1964). In fact,
in many areas it is difficult to collect a shale outcrop sample in which the chlorite
has not altered to Ch/S or Ch/V.
Ch/S has been found in metamorphosed limestones (Wilson and Bain, 1970),
basalts, and volcanic detritus (Suchecki et a/., 1977). It is also found adjacent to
dikes which intrude shales (Blatter et al., 1973) and underlying basalt flows (April,
1980) (Table 2-20). The hydrothermal fluids and heated pore waters apparently
mobilized Mg from the basalt and the sediments and converted the 2 : l physils to
Ch/S or Ch/V.
Corrensite persists in sediments heated to temperatures in excess of 200"C,
perhaps as high as 300°C. Suchecki et a/. (1977) suggested that this is due to the
trioctahedral character of the Ch/S. As the H + associated with the octahedral OH
ions are oriented perpendicular to the c-axis, the repulsion between the interlayer
cations and the 2:l layers is high enough so that corrensite characteristically has a
A
superlattice peak at 28-29 (14 + 14 A) A
which shifts to 30-31 (14 + 17 A) when
treated with ethylene glycol and - A 2 A)
(14 + 10 when heated at 500°C (Fig.
2-30). For a true 1 : l corrensite the 001 peaks are regular submultiples of the 28-31
80

Table 2-20
Chemical Composition and Structural Formula of Dioctahedral and Trioctahedral Ch/S* and Ch/V
1 2 3 4
SiOz 42.14 41.20 35.59 39.04
All01 37.38 12.1 15.13 15.32
Fe103 0.30 1.74 6.65 13.86
FeO - 0.39 9.15 5.57
MgO 0.08 22.0 14.76 25.23
CaO 1.65 1.4 1.86 0.30
Na20 0.15 0.7 1.13 0.68
K2 0 1.40 0.22 1.13
HIOt 15.4 7.30
H2 0 13.51 6.80 6.62

Octahedral
Al 3.01 0.93 0.78 0.35
Fe’+ 0.02 0.1 1 0.45 0.80
Fez+ - 0.03 0.68 0.35
Mg 0.01 2.97 1.97 2.87
1 3.04 4.04 3.88 4.37
Tetrahedral
Si 3.43 3.67 3.18 2.97
Al 0.57 0.33 0.82 1.03

In terlayer
K 0.14 0.02 0.13
Na 0.02 0.02 0.20 0.10
Ca 0.14 0.13 0.18 0.03
* For additional chemical analyses see Brigatti and Poppi (1984).
1. Shirnoda, 1969: Tosudite. hydrothermal from tuff. Fukushirna Prefecture. Japan.
2. Bradley and Weaver. 1956: Corrensite (Ch/S). Mississippi limestone, Jumper Canyon. Colorado.
3. Kirnbara. 1975: Corrensite (Ch/S), Miocene tuff, Japan.
4. April. 1981: Corrensite (Ch/S). contact between basalt dike and Jurassic red beds.

A peak. As the proportion of chlorite layers increases, above 50%, there is less shift
of the 001 peak when the sample is treated with ethylene glycol and when heated.
Conversely, when the chlorite layers decrease in abundance and the smectite layers
increase, the 001 peaks shift more when heated or treated with ethylene glycol.
Additional information on mixed-layer Ch/S, Ch/V. and chloritic physils can be
found on the following pages:
Soils: pages 161-174
Marine: pages 363-366
Evaporites: pages 407-414
Diagenesis: pages 443-446
Sandstones: pages 533, 543-550, 560
81

Untreated

Ethylene Glycol

550' C

Fig. 2-30. X-ray diffraction pattern of oriented slides of Juniper Canyon corrensite. From Bradley and
Weaver, 1956. Arner. Miner., 41.497-504. Copyright 1964 Miner. SOC.Arner.
82

VERMICULITE

Like chlorite, most of o u r understanding of vermiculite is based on a study of


macroscopic crystals. Vermiculite occurs as a clay-sized material in sediments,
although much of it occurs as a mixed-layer clay. Vermiculite is somewhat of a
wastebasket term. Basically, the name is given to the 2:l expandable minerals which
generally have a layer charge larger than that of smectite ( > 0.6 per 0 , , , 0 H 2 ) .
Vermiculite can be either trioctahedral or dioctahedral. Most macroscopic vermicu-
lites are trioctahedral, whereas both types are common as clay-sized minerals.
Probably most vermiculites form by the degradation of pre-existing sheet silicates
and much of their character is determined by that of the starting mineral. Vermicu-
lite results when K is stripped from micas or the hydroxy interlayer sheets from
chlorite. During weathering the layer charge is usually reduced slightly by the
oxidation of ferrous Fe or by hydroxylation of 0 to OH.

X-ruy
Gruner (1934)established that the more common trioctahedral vermiculite had a
monoclinic unit cell with a 5.3 A, b 9.2A, c between 28.57 and 28.77 A, and 0
97'09'. The (060)value for the trioctahedral variety ranges from 1.51 to 1.53 andA
for the dioctahedral variety 1.49 to 1.50A. For additional structural data see Bailey
(1980).
The large size of many vermiculite flakes allows the nature of the interlayer
material to be defined more specifically than for the clay-sized minerals. Mat hieson
and Walker (1954) and Mathieson (1958) have shown that the interlayer water
consists of two discrete sheets. The water molecule sites within each sheet are
arranged in a near hexagonal array with each site being equivalently related to a
single oxygen in the slightly distorted silicate layer surface. Only about two-thirds of
the available water molecule sites and one-ninth of the exchangeable cation sites
were occupied in the specimen Mathieson and Walker examined. This is equivalent
to approximately one Mg per one and one-half unit cells vs. six Mg per unit cell for
a brucite sheet. The Mg ions lie between the water sheets and are octahedrally
coordinated with the water molecules. They point out that their model assumes
static conditions. There is actually a constant migration of the interlayer water
molecules and cations at normal temperatures.
Under normal laboratory conditions the two water layers afford a 14.36 A
spacing when placed in water; at 100% R.H. the spacing increases to 14.81 A
(Walker, 1956).Walker showed that the interlayer water is removed in discrete steps
as the temperature is increased. Initially the two layers of water are rearranged to
A
produce a 13.82 A spacing; next an 11.59 phase is produced as one layer of water
A
is completely removed; this is followed by a 20.6 phase which consiscs of regular
alterations of one water layer (11.59A)and no water layers (9.02A): a 9.02 A phase
was produced when all interlayer water was produced. This latter spacing varies as a
function of the size of the interlayer cation.
The temperature of dehydration is variable, depending on the particle size and
composition of the sample. Due to the high layer charge, temperatures as high as
83

500°C are necessary to drive out all the interlayer water; however, water is readily
readsorbed until the vermiculite is heated to 700°C.
The O O t peaks of well developed vermiculites have a distinctive series of
intensities. The 001 (14.4 A) peak is the most intense; the 002 is the least intense;
and there is a progressive increase in intensity from the 002 to the 005 peaks. Fig.
3-20 ontains x-ray patterns of soil vermiculites and mixed-layer mica/vermiculites.

Chemistry
Mucroscopic: The expanded 2:l minerals have a continuous gradation of cation
exchange capacities ranging from approximately 25 to 250 meq/100 g. Weiss et al.
(1955) and Hofmann (1956) suggested the division, between smectite and vermicu-
lite be made at 115 meq/100 g. This is equivalent to a layer charge of 0.55 per
O,,(OH)z unit of structure; Walker (1958) has suggested the boundary value should
be placed slightly higher. Harward et al. (1969) reported an average CEC of 162
meq/100 g for 13 vermiculites with a range of 135 to 189 meq/100 g.
Why is there a need for a boundary? The expansion, contraction, and adsorption
properties of expandable 2:l physils vary as the particle size, layer charge, and
tetrahedral charge vary. In general, as these three properties increase, and they
frequently (but not always) increase together, the layers are better able to contract
to 10 A when potassium is placed in the interlayer positions; they imbibe one layer
of ethylene glycol or glycerol rather than two; the interlayer water is better
organized and fewer layers (two) can be accommodated in the interlayer position.
These processes are generally operative when the total layer charge is larger than 0.6
to 0.7. The location of the charge appears to be a factor. When a high proportion of
the charge originates in the tetrahedral sheet, less total charge is required to restrict
the expansion between layers.
The generally accepted criteria for calling an expanded physil vermiculite or
smectite is the response of the physil to ethylene glycol and/or K. Mg saturated
vermiculites adsorb only one layer of ethylene glycol and have a spacing of 14 A;
smectites, except for some beidellites, adsorb two layers of ethylene glycol and have
a spacing of 17 A. When saturated with K, at - 50% r.h., vermiculites collapse to
10.0 to 10.5 A and smectites to approximately 12 A (Harward et at., 1969). Some
samples do not meet both criteria. Weaver (1958) described a sample derived from
the weathering of illite that expanded to near 17 A but contracted to 10 A when K
saturated. These samples are relatively rare. At the extremes, smectites (layer charge
< 0.4) and vermiculites (layer charge > 0.7) are quite distinct; in the intermediate
area, assignment can be difficult and probably is not all that significant.
Many microscopic vermiculites have been created by the removal of potassium
from biotite and phlogopites, and the interlayer hydroxyl sheet from chlorites.
Biotite can easily be altered to vermiculite in the laboratory by exposing it to a
dilute Mg solution. Biotites have a theoretical exchange capacity of 250-260
meq/100 g. Oxidation of octahedral iron and hydroxylation of oxygen reduces this
somewhat, but the resulting charge is generally significantly larger than that of
smectite and usually enough to cause lattice contraction when potassium is accessi-
ble. Muscovites which are leached of potassium should have an even higher charge,
84

as these contain less ferrous iron. Vermiculites derived from chlorite should gener-
ally have a lower layer charge than those derived from micas.
Most coarse grained vermiculites are trioctahedral. Many of the finer soil
vermiculites are dioctahedral, formed from illite or authigenicly from solution. The
authigenic vermiculites commonly form in soils developed on basic igneous rocks
(Walker, 1975).
During the process of alteration of non-expanded sheet structure minerals to
vermiculite, an intermediate mixed-layer phase is commonly produced.
Foster (1963) calculated the structural formulas for 25 macroscopic vermiculites,
and Brigatti and Poppi (1980) compiled the chemical data on 124 vermiculites.
Foster's vermiculites have octahedral compositions similar to phlogopites, Mg-bio-
tites, Mg-chlorite, and saponite; but due to the oxidation of much of the iron they
tend to have a higher proportion of trivalent octahedral cations. The tetrahedral
sheets have the relatively high charge characteristic of biotites and chlorites.
Octahedral R3' occupies up to 40-45 percent of the filled octahedral positions; this
is a higher value than is found for any other trioctahedral sheet structure clays
except for the high iron 1:l clays. The high R'+ values in vermiculite are probably
due to the oxidation of iron during weathering.
Most of the vermiculites listed by Foster apparently were formed by the leaching
of K from biotite. Biotite has a negative layer charge near 1.00 per O,,(OH), units.
Vermiculites have layer charges ranging from 1.08 to 0.38 with only five of 25 values
being less than 0.57. The reduced layer charge is due to the oxidation of ferrous iron
and the hydroxylation of oxygen. When the layer charge is reduced, the potassium is
replaced, commonly by hydrated Mg ions, and the layers expand. The average
structure of Foster's 25 vermiculites is:

(A]0 14Feo3.;9Mg2.40 )(Si 2.87A1,.13 )010(OH)ZX0.71


Microscopic,: Clay-sized vermiculite and vermiculite layers interstratified with
mica or chlorite layers are quite common in soils where weathering is not overly
aggressive. (A few references are: Walker (1949), Brown (1953), Van der Marel
(1954). Hathaway (1955). Droste (1956), Rich (1968). Weaver (1958), Gjems (1963).
Millot and Camez (1963), and Barshad and Kishk (1969).) Most of these clays are
formed by the removal of K from biotite, muscovite, and illite or the brucite sheet
from chlorite. This is accompanied by the oxidation of much of the iron in the 2:l
lattice.
Some of the expanding physils in soils have the attributes of vermiculite, some of
smectite, and some have features of both. The variation in properties is largely
related to the layer charge. The charge is dependent o n the original charge on the
2:l layers of the parent mineral, the amount of ferrous iron in the octahedral sheet.
and the nature of the leaching conditions. Artificial weathering of biotite (Ismail.
1969) showed that oxidation of ferrous iron was independent of pH; however, under
neutral and alkaline conditions iron oxidation caused a large decrease in layer
charge and a smectite formed; under acid conditions the charge decrease due to the
oxidation of ferrous iron was balanced by the loss of octahedral Fe and Mg. and a
highly charged vermiculitic material was formed. Thus. in general, smectites would
85

Table 2-21
Structural Formulas for Soil Vermiculites. from Barshad and Kishk (1969)
~~

1 2 3 4
Octahedral
Al 1.44 1.12 0.33 -
Fe’+ 0.16 0.29 0.49 0.90
Ti 0.14 0.16 - -
Mg 0.27 0.75 1.26 1.22
Mn 0.05 0.03 0.08 0.03
H 0.30 0.19 0.19 0.04
1 2.36 2.65 2.22 2.17

Tetrahedral
Al 1.10 1.31 - -
Si 2.90 2.69 4.0 4.0

Interlayer
Na 0.43 0.44 0.68 0.76
K 0.37 0.25 - -

Layer Charge
Octahedral + 0.30 + 0.62 ~ 0.68 - 0.76
Tetrahedral - 1.10 -1.31 0.0 0.0

Total 0.80 0.69 0.68 0.76

C.E.C. meq/100g 212 181 168 196

Sample Soil Great Location Parent pH Mode of


No. Series Soil Material of Formation of
and Refer- Group Soil Vermiculite
ence No. Clay
1 Taiwan lateritic Eastern fine 5.7 a!teration
11A red earth Taiwan, grain of mica
China schist
2 Masterson podzolic Tehama Co.. mica. 5.3 alteration
58-52-22-3 California schist of mica
3 Aiken red. Lytonville, green- 6.0 alteration
138 podzolic California stone and
synthesis

4 Boomer brown Tehama Co.. basalt 5.7 synthesis


58-52-26-7 forest California

form in the more arid environments and vermiculite in the more humid environ-
ments.
There are few chemical analyses of pure clay vermiculite; however. Barshad and
Kishk (1969) made a detailed mineral and chemical study of 11 soils and calculated
a structural formula for the vermiculitic component (Table 2-21).
They found that the soil vermiculites could be placed into two groups. The first
group had at least one Al per four tetrahedral sites and Al in the octahedral sheet
86

(Samples 1 and 2). The charge deficiency in the tetrahedral sheet was partially
compensated for by a positive charge in the octahedral sheet ( > two cations per
three sites). These samples are similar to the coarser-grained vermiculites in com-
position. The second group (Samples 3 and 4) had no tetrahedral Al. the charge
deficiency arising from the octahedral sheet in which Mg and Fe’+ were the
dominant cations. Both groups were more dioctahedral than trioctahedral and had
similar layer charges (0.60 to 0.80). The vermiculite with Al in the tetrahedral sheet
is associated with mica and formed from the alteration of mica derived from acid
igneous rocks and their sedimentary derivatives. The samples in the group with no
A l in the tetrahedral sheet and dominant Mg and Fe3+ in the octahedral sheet are
associated with montmorillonite and probably formed authigenically from ions
provided by a basic igneous parent material. Barshad and Kishk suggested that this
group of vermiculites form a continuous series with montmorillonite, distinguishable
from the montmorillonite only by a determination of C.E.C.
The clay vermiculites developed by mild leaching action of pre-existing sheet
structures tend to inherit much of their octahedral and tetrahedral character. Clay
vermiculites formed by relatively intense weathering and by diagenetic alterations.
and approximate equilibrium with their soil environment, will have little. if any,
tetrahedral Al; the octahedral sheet can be quite variable in composition and
depends on the availability of Al, Fe, and Mg.
Quite often Al, Fe, and Mg hydroxides partially fill the interlayer position of the
“derived” vermiculites and decrease their exchange capacity and their ability to
contract completely to 10 A when heated or when treated with a potassium solution.
This material can usually be removed by treating the clay with a solution of sodium
citrate (Tamura, 1958). As the content of hydroxy interlayer material increases, the
expandable clay tends to assume the character of a chlorite. Thus, in the weathering
of a mica or illite i t is not uncommon to form discrete vermiculite-like, beidellite-like,
montmorillonite-like, and chlorite-like layers. These various layers can occur as
discrete packets or interstratified in a wide variety of proportions.
Roth et al. (1969) has demonstrated that clay vermiculites commonly have a
surface coating of positively charged Fe,O, (approximately 10 percent Al ,O, is also
present) which causes a decrease in C.E.C. This material can be removed (deferra-
tion) by reduction of free Fe,O, with Na,SzO, in the presence of Na citrate and
NaHCO,.
Additional information on vermiculites can be found on pages 161-170.

GLAUCON ITE-CELADONITE

Glauconites and celadonites are basically dioctahedral Fe-rich illi tes and I/S.
The term glauconite is used both as a mineral name and a “rock” name. The term
has been used for the past century and a half to refer to greenish-colored pellets
found in sedimentary rocks (for literature review see McRae, 1972). More recently
the term has been applied to brown-and-white-colored pellets, which, considering
the name is derived from the Greek word “glauconie”, meaning bluish or pale green,
x7

is probably an inappropriate use. Warshaw (1957) and Burst (1958), and subse-
quently many others (Thompson and Hower, 1975), demonstrated that though
many green pellets are monomineralic and consist of Fe-rich I/S. many are
composed of other physils such as smectite, chlorite (Fe), kaolinite, and mixtures o f
several physils.
In order to eliminate some of the confusion, Odin and Matter (1981) suggested
that green grains be called glaucony and the term glauconite be used only as a
mineral name. This seems like a reasonable suggestion. In addition to the typical
pellet form, glaucony also occurs as irregular films and patches replacing such
things as quartz, mica, calcite, volcanic rock fragments, phosphates. etc.
At one time it was believed that glauconite formed under marine conditions and
that celadonite formed in basalts under continental conditions. More recent studies
indicate low-temperature hydrothermal celadonite is abundant in deep-sea marine
basalts.

Structure
Glauconite and celadonite occur in the 1M and 1Md structural forms, though
high-resolution transmission electron microscopy studies indicate minor amounts of
2M mica is present in some glauconite (Amouric and Parron, 1985). X-ray data for
the better crystallized samples are given in Tabe 2-21. Both minerals contain
interstratified smectite interlayers. I t has been recommended that the names
glauconite and celadonite be applied only to those minerals containing less than 5%
interlayering. Celadonite is generally well organized and produces sharp basal and
hkl reflections (Fig. 2-31) and has a 1M structure. Glauconite has a broader basal
reflection and reduced hkl reflection and commonly a 1Md type structure (Fig.
2-31) (Buckley et d., 1978).
Glauconite layers occur interstratified with smectite layers and form a mixed-layer
G / S sequence similar to that of I/S. When the G/S contains > 30% smectite layers
the interlayering is random; between 15 and 30% smectite the ordering is allevar-
dite-like, GS; when < 15% smectite layers are present the ordering is GSGG.
equivalent to IMII (Thompson and Hower. 1975). In the older literature mixed-layer
G / S is referred to as glauconite, but the AIPEA Nomenclature Committee (Bailey,
1980) recommended that it be referred to as interstratified glauconite-smectite. The
smectite layers are commonly nontronite. Fig. 2-32 shows x-ray patterns of G / S
containing from 5 to 60% smectite layers.

Composition
The AIPEA Nomenclature Committee recommended that glauconite be defined
as an Fe-rich dioctahedral mica containing more than 0.2 atoms of tetrahedral
AI/0,,(OH)2 and having a d(060) > 1.510 A. Celadonite should contain less than
0.2 atoms of tetrahedral Al and have a d(060) < 1.510 A.The Committee (Bailey,
1986) later modified the definition to include the statement that the total charge of
the octahedral cations in glauconite is =. 5.3 (high R”) per O,,,(OH), and for
-
celadonite < 5.3 (high R2+). Celadonites, on an average, contain approximately
twice as much octahedral Mg as glauconite (Weaver and Pollard, 1973).
88

Celadonite

-
Glauconite

I I I I I I 1 I I I
38 36 34 32 30 28 26 24 22 20 lo 8 6

Fig. 2-31. XRD traces showing the characteristic peaks of glauconite and celadonite. From Buckley et al..
1978. Copyright 1978 Min. SOC.London.

Gluuconite: Numerous scientists (Hendricks and Ross, 1941; Smulikowski, 1954;


Borchert and Braun, 1963; Cimba Inikova, 1971; Weaver and Pollard, 1973;
Thompson and Hower, 1975; Buckley et ul., 1978; Odin and Mather, 1981; etc.)
have compiled data on the composition of glauconite and glauconitic material.
Table 2-23 contains analyses of glauconites with relatively few expanded layers.
Compositions are relatively uniform. A typical average formula based on 137
analyses of samples with few expanded layers is:

(Al,,,,Fe,3.~~Fe,Z.:,Mg,.4~)(si,.,,Al,.,,)O*,(OH)*K,.,,Na,,,Ca,.,,
Characteristically glauconites have about half of their occupied octahedral positions
filled with Fe3+,as compared with an average of about 10% for illite. The amount
of tetrahedral A1 and tetrahedral charge averages approximately half that of illite
(0.31 vs. 0.60). Though most glauconites (green marine pellets) contain more than
0.2 tetrahedral Al, some contain less. Approximately 20% of glauconites have more
tetrahedral charge than octahedral charge. Frequency distribution curves of the ions
in glauconite and illite show overlap for each ion. The minimum overlap is in
octahedral Al (Weaver and Pollard, 1973). The maximum amount of Al in the
octahedral sheet of glauconite is approximately 1.30 and the minimum amount in
illite is approximately 1.20. The two populations are nearly mutually exclusive and,
as such, are more characteristic of the two families of minerals than is the iron
content. Ferrous iron is almost as selective as Al.
89

.k 15%
RAN

Fig. 2-32. Comparison of X-ray powder diffraction patterns of oriented, glycol solvated. glauconite. with
computer calculated diffraction profiles. The tracings labeled (3294, GI 392, G3585, JH63-3. W89-6. and
G68A are actual patterns from glauconite samples, the profile immediately below each glauconite pattern
is a calculated pattern. N = the range in number of layers per crystallite used in calculating the patterns.
Interlayering type is either random (RAN), ordered (IM) or ordered with an IMll suplatice. Sample
G68A contains some 7 A material. From Thompson and Hower. 1975. Copyright 1975 The Clay Miner.
Soc.

For most samples the calculated octahedral cation population is larger than 2.00
(average 2.05). Leaching studies (Thompson and Hower, 1975) suggested the Al, Fe,
and Mg in excess of that necessary to fill 2.00 octahedral positions are present as
hydroxy complexes. Free iron oxides are present in some samples (Bentor and
Kastner, 1965).
As for the I/S physils, the amount of K present in glauconitic physils (G/S) is
proportional to the percentage of contracted layers (Manghnani and Hower, 1964;
Thompson and Hower, 1975). The K concentration between contracted glauconite
layers is less than that between the illite layers in I/S (Table 2-23).
According to Thompson and Hower (1975), the lower K concentration in
glauconite layers is due to the 1M structure and high Fe content of the octahedral
sheet. The orientation of the 0 - H axis is inclined at nearly right angles to the K +
ion. Thus, the K + is strongly bonded and less is needed to cause layer contraction.
It is also possible that the relatively small amount of tetrahedral rotation allows the
K + to be more deeply seated, and more strongly bonded, in the “hexagonal” holes
in the oxygen sheets.
Cation exchange capacities as determined by Manghnani and Hower (1964) show
a good linear relation to the relative proportion of expanded layers (Fig. 2-33). The
90

TABLE 2-22
X-ray data for glauconite and celadonite. After Bailey. 1980.
A. Glauconite-IM (Warshaw, 1957) B . Celadonite-l M (Wise and Eugster, 1964)
d (obs.) I hkl hkl d (oh.) r l (calc.) I
10.1 100 00 1 001 9.97 9.94 47
4.98 * 002 020 4.53 4.53 x5
4.53 80 020 1i i 4.35 4.36 42
4.35 20 1i i 021 4.14 4.123 37
4.12 10 02 1 112 3.635 3.638 XO
3.63 40 112 022 3.35 ‘ 3.349 60
3.33 60 003.022 003 3.318 I 3.314 70
3.09 40 112 112 3.087 3.0X1 xo
2.89 5 113 113 2.90 B 2.901 10
2.67 10 023 023 2.678 2.675 75
130.131 130 2.604 2.605 70
2.587 100
200 I 3T 2.580 2.581 100
2.396 60 132.201 132 2.402 2.404 75
2.263 20 040. 22i 040 2.264 2.265 18
2.213 10 220.041 04 1 2.209 2.208 25
2.154 20 133.202 133 2.150
2.148 31
1.994 20 B 005 202 2.125
1.817 5 224 005 2.092 1.988 10
1.715 10 31i. 24T 151 1.65 B 1.665 15
240, 312 060 1.509 1.510 60
1.66 30 B
310. 241
1.511 60 060.33T
1.495 10 30
1.307 30 260.400
170. 350
1.258 10
420
A. * Obtained only with oriented sample.
a = 5.234. b = 9.066. c = 10.16 A. p = 100.5
Specimen from Vidono Fm., Anzoategui. Venezuela.
B. I The exact location of these two peaks is difficult to obtain. B = broad peak
a = 5.23. b = 9.06. c =10.13 A, 0 =100°55’.
Specimen from Wind River area. Washington.

C.E.C. ranges from 5 to 12 meq/100 g for those glauconites having only 5 percent
expandable layers and increases to 30 to 40 meq/100 g for those having 40 to SO
percent expandable layers. The C.E.C. of the expanded layers alone ranges from 60
meq/100 g for those clays with a relatively high proportion of expandable layers to
90 meq/100 g for those with only a few expandable layers.
Celudonite: The octahedral sheets of celadonite contain on the average consider-
ably more Mg than glauconite (0.63 vs. 0.39 or 6.12% MgO vs. 3.58% MgO). Fe”
(0.21 vs. 0.20) and Al (0.49 vs. 0.44) values are similar, but octahedral Fe3+(0.72 vs.
1.01) averages less for celadonite than for glauconite. The Fe/Mg ratio in celadonite
(1.5) is less than half that of glauconites (3.1) (Weaver and Pollard, 1973).
91

Table 2-23
Chemical Compositions and Structural Formulasof Glauconites with < 10% Expanded Layers.

1 2 3 4 5
SiO, 53.13 56.98 48.5 53.59 50.69
A1 2 0 , 10.23 13.33 9.0 8.90 7.20
Fe20, 21.70 8.78 20.0 15.21 15.43
FeO 1.74 7.05 3.1 3.20 2.82
MgO 3.43 5.35 3.7 3.37 4.67
CaO 0.15 0.05 0.4 0.06 0.54
Na,O 0.20 0.23 1.5 0.21 0.58
K2O 8.50 8.21 6.1 7.09 7.43
TiO, 0.17 0.08 - 0.15 0.20
H,O+ 7.3 7.52 8.91
H20- 1.73 1.94
Total 99.56 100.84 99.6 101.03 100.21
Octahedral
Al 0.45 0.80 0.35 0.63 0.43
Fe3+ 1.11 0.44 1.12 0.82 0.87
Fe2+ 0.10 0.39 0.19 0.19 0.18
Mg 0.35 0.53 0.40 0.36 0.52
E 2.01 2.16 2.06 2.00 2.00
Tetrahedral
Si 3.62 3.76 3.58 3.87 3.79
Al 0.38 0.24 0.42 0.13 0.21
Interlayer
K 0.74 0.69 0.57 0.65 0.71
Na 0.21 0.03 0.08
0.03 0.02 0.03 - 0.04
Ca
R' +/R2 + 3.47 1.35 2.49 2.64 1.n6
1. Thompson and Hower. 1975: Cambrian, Gros Ventre Formation, Wyoming.
2. Thompson and Hower, 1975: Ordovician, Latorp, Sweden.
3. Schneider, 1927: Cambrian, Franconia Formation, Norwalk, Wisconsin.
4. Warshaw. 1957: Eocene, New Jersey.
5. Amouric and Parion. 1985: Paleocene, Ivory Coast.

An average formula based on 15 analyses of hydrothermally altered continental


basalt is:

Calculated layer charge is 0.88 with nearly 80% of this charge originating in the
octahedral sheet, compared with 55% for glauconite. Another average formula based
on 13 analyses of marine and continental samples (Buckley et al, 1978) is:

( 0.29 Fei.:,Fei.i2 Mg0.69) ( si3.96 0.04I') 0 (OH)?K 0.86 Na 0.02Ca 0.02

Celadonite is the only 2:l dioctahedral physil for which the octahedral R2+ions are
more abundant or nearly as abundant as the R3' ions. Typical compositions and
structural formulas are given in Table 2-24. Though published x-ray data are
92

limited, many reported “celadonites” have low K ,O values, suggesting they are
mixed-layer celadonite/smectite, probably celadonite/nontronite.
Glaucony occurs in rock ranging in age from Precambrian to Recent. Studies of
Recent sediments indicate it is formed on the outer portion of continental shelves at
depths between 60 and 500 m, where the sedimentation rate is low. Green pellets
have been studied for years and a variety of theories have been proposed to explain
their origin. A long-held theory proposed by Burst (1958) and later by Hower (1961)
is that glaucony forms from degraded 2:l layer lattice silicates by the replacement
of octahedral A1 by Fe and the simultaneous adsorption of K.
More recently Odin and Mather (1981) suggested that glaucony forms initially by

Table 2-24
Mean Number of K Atoms per O,,(OH), in the Glauconite Layers of G/S and Illite Layers of I/S.
After Thompson and Hower, 1975.

40 0.30 0.58
50 0.34 0.63
60 0.39 0.67
70 0.46 0.70
80 0.56 0.72
90 0.69 0.74
93

the crystal growth of glauconite smectite (high Fe mixed-layer G/S) in pores of


different substrates. With time the initial substrate is replaced or “disappears” and
the glauconitic smectite evolves into a glauconitic mica by a recrystallization
process.
For additional discussion on the occurrence and origin of glauconite see Chapter
VI (pp. 355-371, 387-400).
Celadonite has been considered to be an alteration product, usually hydrothermal
or low-grade metamorphic, of continental basalts. Deep sea drilling has shown the
celadonite is relatively abundant in deep sea basalts where it has formed by low
temperature hydrothermal alteration of the basalts (Chapter VII, pp. 355-363). In
many instances it is not clear whether the Fe illite should be classed as celadonite or
glauconite.

TALC-KEROLITE

Structure
Talc, Mg,Si40,0(OH),, is a trioctahedral (Mg), 2:l physil which has neutral, or
nearly neutral, layers and contains no interlayer material. Layers are held together
by van der Waals bonds and weak ionic attractions caused by minor substitutions in
the octahedral and tetrahedral sheets. Talc has a one-layer triclinic structure. The
001 spacing is of 9.3 to 9.4 A (Bailey, 1980); peaks are sharp and narrow. Kerolite is
similar to talc in composition and structure but has a highly random layer st:cking.
possibly contains some interlayer water, and has a basal spacing of about 9.6 A. The
x-ray reflections are considerably broader than those of talc (Fig. 2-34) (Brindley er
al., 1977). The thicker basal spacing is attributed to the disordered stacking which
does not permit the degree of close-packing of adjacent planes of oxygen ions found
in well-ordered talc (Rayner and Brown, 1973).

h~ .I . 1 1 1 . 1 . 1 1

II 1 003
TALC

132

001
06 3 3

... ,

I ’ 1 . 1 I I 1 I 1 .

10 20 30 40 50 60 70
26 d e g r e e s
Fig. 2-34. X-ray powder diffraction patterns of kerolite and talc. Cu-Ka radiation. Powders packed
gently into shallow cavities. From Brindley et al., 1977. Copyright 1977 Min. SOC.London.
94

Table 2-25
Electron Probe Microanalyses and Structural Formulas of Selected Non-Marine and Marine Celadonites.
1 2 3 4 5
58.84 52.58 56.05 47.23 52.28
1.69 5.28 6.88 2.87 0.82
14.81 14.61 10.30 12.67 18.99
5.88 3.37 3.98 2.81 3.02
6.98 6.35 4.95 9.00 6.33
0.07 0.68 0.22 0.39 0.16
0.15 0.31 0.05 0.00 0.00
8.93 7.69 10.29 8.57 9.43
0.00 0.20 0.04 0.00 0.00
93.35 91.07 92.76 83.54 91.03
Octahedral
Al 0.15 0.28 0.58 0.10 0.02
Fe” 0.81 0.78 0.56 0.77 1.08
Fe2+ 0.39 0.22 0.24 0.19 0.19
Mg 0.75 0.69 0.53 1.08 0.71
E 2.10 1.98 1.91 2.14 2.00
Tetrahedral
Si 4.01 3.83 4.01 3.83 3.95
Al 0.00 0.17 0.00 0.17 0.05
lnterlayer
K 0.83 0.73 0.94 0.88 0.91
Na 0.02 0.04 0.00 0.00 0.00
Ca 0.01 0.10 0.03 0.06 0.03
R’+/R~ + 0.84 1.16 1.30 0.69 1.22
1. Buckley et d..1978: Marine, vein in basalt.
2. Buckley el a / . , 1978: Marine, in crack in basalt.
3. Buckley et a / . , 1978: Continental. nodules, Faeroes.
4. Buckley et a / . , 1978: Continental, massive in basalt, Washoe Co., Nevada.
5. Buckley el d., 1978: Continental, massive, Omaru District. New Zealand.

Composition
Most talcs (Table 2-25) have a composition close to that of the ideal formula.
Commonly small amounts of Al (0.03 to 0.15/0,,(OH),) substitute for Si in the
tetrahedral sheet, giving the mineral a slight negative charge. Minor amounts of Al
and Fe can be present in the octahedral sheet (Deer et al., 1962). Padan (1984)
described an Al-rich talc (Table 2-25) from the Paradox salt which contained
approximately 5% Al,O,: the 001 peaks were broader than talc and the 001 spacing
was slightly larger, 9.5 A. Padan suggested the larger 001 spacing was due to the
presence of scattered brucite islands between the low charged layers. Kerolite (Table
2-25) has the composition of talc, except it contains 0.75 to 1.2 H,O/O,,,(OH),.
The H,O is believed to be adsorbed on the surface and between some layers;
however, kerolite does not expand or contract.
The talc minerals can form from ultrabasic rock by hydrothermal alteration or
weathering. It also forms during metamorphism of Mg-rich rocks, producing talc-rich
95

schists. More relevant is the low temperature origin of the talc minerals. Talc is a
common mineral in evaporite deposits. In some deposits i t is present only in the
anhydrite (Braitsch, 1971) and in others only in the halite beds (Padan. 1984). I t is
present in potash salts (Stewart, 1965; Bodine, 1978) and Yucatan dolomites
(Isphording. 1984). Some of the talc in evaporite deposits may be diagenetic, but
some is definitely authigenic (Padan, 1984). Kerolite has been found in lateritic soils
formed on ultrabasic rocks (Maksimovid, 1966; Wiewi6ra ef d.,1981) and in
lacustrine evaporites (Eberl et a/., 1982).
A number of mixed-layer talc minerals have been identified. The most common is
a talc/saponite (Table 2-25). I t has been described by Alietti (1958). Alietti and
Meisner (1980), and Veniale and Van der Marel (1969). In general these minerals
are associated with serpentine deposits and may or may not have a hydrothermal
origin. Mixed-layer kerolite/stevensite (Table 2-25), with 50 to 80% stevensite
layers, is abundant in the Amargosa Desert, where it precipitated in shallow,
spring-fed lakes or swamps following moderate evaporation (Eberl et d.. 1982;
Khoury et a/., 1982).
Additional information on the origin of talc and talc-like minerals can be found
on the following pages:
Soils: page 185
Lacustrine: pages 276-277
Marine: pages 323, 324, 341, 363
Evapori tes: pages 408-41 1

SEPIOLITE-PALY GORSKITE

Structure
Sepiolite and palygorskite ( = attapulgite) are “chain structure” fibrous or lath-like
physils. The laths are elongate along the X axis. with the z axis perpendicular to the
thin dimension of the lath (Fig. 2-35). Basically, the chain structure physils consist
of 2:l layers with limited growth along the y axis. The 2:l units are similar to those
in the sheet structure physils - an octahedral sheet sandwiched between two silica
tetrahedral sheets. The tetrahedral sheets extend for considerable distances in the a
and b directions; however, at periodic intervals, along the b axis, the tetrahedral
sheets invert, forming a checkerboard pattern (Fig. 2-35). Alternate squares, or
ribbons, are filled with water. The inversion is presumably due to strain induced in
the 2:l layers because of misfit between the octahedral and tetrahedral sheets or
distortion within the octahedral sheet (Weaver and Pollard, 1973). Structurally the
two chain structure silicates differ in the number of octahedral positions in the 2 : l
units.
Nagy and Bradley (1955) proposed a structure for sepiolite which contains nine
octahedral cation positions, of which eight are filled. A model proposed by Brauner
and Preisinger (1956) has only eight octahedral cation positions. The compositional
data favor the latter model (Bailey, 1980). Palygorskite has only five octahedral
cation positions, of which four are filled (Bradley, 1940).
96

SEPIOLITE

PALYQORSKITE

......... .... ............ .... .........

0....0............ 0....0.......
- 0H 2 0
........

b.18.0i

0 HYDROXYL 0 M g or Al
-
Q OH2 0 OXYGEN o SILICON

Fig. 2-35. [loo] projection of sepiolite and palygorskite structures. From Bailey, 1980. Copyright 1980
Miner. SOC.London.

A
The hollow channels (3.8 X 6.3 for palygorskite and 3.8 and 9.4 Afor sepiolite)
that alternate with the 2:l ribbons contain zeolitic water ( < 25OOC). In addition,
four bound water molecules (OH,) per formula unit occur at the edges of the 2:l
97

Table 2-26
Chemical Composition and Structural Formula of Talc and Talc-Like Minerals.

1 2 3 4 5
SiOz 63.90 58.63 50.29 54.31 63.15
A1 2 0 3 0.03 0.05 3.73 1.90 1.44
Fe203 0.21 0.10 1.14 - 0.69
FeO - - - 2.95 -
MgO 31.49 30.84 26.74 27.44 29.23
CaO 0.08 0.33 0.01 1.01 0.01
Na20 0.02 - 0.42 0.32 3.86
K2 0 0.01 - 0.05 0.86 1.18
Li ?O 0.38
TI02 0.10 0.05
Octahedral
A1 - - 0.05 - 0.07
Fe3+ 0.01 - 0.06 - 0.03
Fez+ - - - 0.17
Mg 2.95 3.09 2.95 2.85 2.72
Li -
E 2.96 3.09 3.06 3.07 2.91
Tetrahedral
Si 4.01 3.935 3.72 3.79 3.96
Al 0.002 0.004 0.28 0.21 0.04

lnterlayer
Na 0.002 - 0.060
K - - 0.005
Ca 0.005 0.024
Total 0.007 0.024 0.065 0.26 0.21

1. Brindley et d.,1977: Talc, Manchuria.


2. Brindley et a/., 1977: Kerolite. Goles Mt. laterite, Yugoslavia.
3. Padan, 1984: Al-talc in anhydrite. Paradox Basin Salt, Utah.
4. Alietti, 1958: Talc/saponite, hydrothermal alteration of serpentine. Taro Valley, Italy.
5. Eberl et al., 1982: Kerolite/stevensite (3:7), Arnargosa Desert, Utah.

ribbons. As they participate in the coordination of the octahedral cations they are
more tightly bonded than the zeolitic water. Some exchangeable cations are also
present in these channels, though the exchange capacity is not well known.
-
Sepiolite has the following unit cell dimensions: a = 5.28, b = 26.95, c = -
- 13.37 A, -
p = 90". The laths cleave along the diagonal planes where tetrahedral
inversion occurs. Thus, the major peak observed in x-ray patterns of oriented
samples is the 110 (or 01 of Bailey, 1980) reflection at 12 (Table 2-26).A
Palygorskite is more complex structurally and chemically than sepiolite. Unit cell
- -
dimensions are: a = 5.21, b = 17.9, c sin p = 12.7 A, O , = 90", 96", or 107". At
least one orthorhombic and three different monoclinic geometry unit cells have been
identified. The structure is similar to that of sepiolite except for the shorter b
dimension. The mean 110 (or 011) reflection occurs at 10.3 to 10.5 A (Table 2-26).
For more details on the structure see Bailey (1980).
98

Both minerals occur as short (0.25 to 1.0 pm) and long (occasionally > 10pm)
fibers (Fig. ). As yet, there has been no relation established between structure or
composition and fiber length.
Chemistry
Sepiolite: The ideal structural formula for sepiolite based on the Nagy-Bradley
model is (Si 1 2 )(Mg, )030(OH),(OH2)&H and )(Mg8)OdOH),(OH, ) 8 H , O
for the Brauner-Preisinger model. Table 2-27 shows the chemical compositions and
the structural formulas of selected sepiolite samples (based on the Brauner-Pre-
isinger model). Most sepiolites contain a minor amount (0.04 to 1.05) of Al'+
and/or Fe3+ substituting for Si4+ (11.96-10.95) in the tetrahedral sheet. The
magnesium-rich sepiolites have a relatively consistent composition. A1 203and
Fe,O, are commonly present in amounts less than the one percent; the MgO
content ranges from 21 to 25%. Mg fills 90 to 100% of occupied octahedral
positions. The analyses indicate sufficient cations to fill approximately eight (7.74 to
8.14) octahedral sites (Weaver and Pollard, 1973). This would fill all the sites
allotted by Brauner and Preisinger and leave one vacant site in the Nagy-Bradley
structure. The average octahedral cation charge is approximately 16.0, indicating
little or no octahedral charge. Sepiolites (50) have an average composition of 54.2%
SiO, and 24.2% MgO (Martin-Vivaldi and Cano-Ruis, 1955).
Although the Mg-rich variety is the most common, sepiolite-type minerals afford
a wide range of substitutions which are mostly accommodated in the octahedral
sheet. An aluminous sepiolite described by Rogers et al. (1956) has 19% of the
octahedral positions filled with Al. Xylolite, an iron-rich sepiolite, can have up to
9% of the tetrahedral sites and 39% of the filled octahedral sites filled with Fe3+.
Fahey et al. (1960) have described a variety of sepiolite which contains 8.16% N a 2 0
and only 16.015% MgO (loughlinite). Leaching experiments indicate sodium sub-
stitutes for magnesium.
Most of the cations in the octahedral positions are the large variety; Al, a smaller
cation, is relatively uncommon - this is in contrast to palygorskite, where A1
commonly fills half the octahedral positions. It appears that most of the exchange
capacity is due to the replacement of silicon by trivalent elements, though this is not
certain. The reported exchange capacity ranges between 20 and 45 meq/100 g.
Sepiolite most commonly crystallizes from Si- and Mg-rich solutions and is
commonly associated with carbonate and phosphate rocks. It frequently occurs
along with palygorskite and/or smecite. It is reported to have formed in lacustrine
environments of high basicity (Longchambon and Morgues, 1927; Millot, 1949;
McLean et al., 1972), in pluvial lakes (Parry and Reeves, 1968; Papke, 1972), in
normal marine environments (Millot, 1970), by hydrothermal alteration of serpen-
tine (Midgley, 1959), by hydrothermal activity in a deep marine environment
(Hathaway and Sachs, 1965), and in arid soils (Van der Heuvel, 1966).
Additional information on the origin of sepiolite can be found on the following
pages:
Soils: pages 182-185
Lacustrine: pages 275-278
99

Marine: pages 379-387


Brackish: pages 400-407
Palygorskite: Chemical analyses and structural formulas of palygorskite samples
are listed in Table 2-28. None of the samples approach that of the model formula

Table 2-27
X-ray powder data for sepiolite and palygorskite.
Sepiolite (Brindley. 1959) Palygorskite (Christ et a1.,1969)
hkl d(1) hkl d(calc) d(obs) I(obs)
110 12.03100) 110 10.43 10.54 100
130 7.47(10) 200 6.411 6.42 10
200 6.73(5) 130 5.416 5.418 7
040 6.73(5) 040 4.481 4.479 23
150 5.01(7) 121 4.244 4.258 15
060 4.498(25) 310 4.158 s 4.137 10
131 4.306(40)
330 4.022(7) 131 3.751 sh
260 3.750(30) 221 3.682 3.681 7
241 3.533(12) 240 3.673 sh
080 3.366(30) 150 3.453 b 3.444 5
331 3.196(35) 31 1 3.247 3.246 20
261 3.050(12) 400 3.206 3.205 18
370 2.932(4) 321 3.098 sh
081 2.825(7) 41 1 2.698 2.698 4
421 2.771(4) 25 1 2.681 sh
0,lO.O 2.691(20) 421 2.610 b
510 2.691(2) 440 2.607 b 2.601 10
44 1 2.617( 30) 06 1 2.590 b
281 2.61 7(30) 102 2.548 2.548 13b
5 30 2.586(NR) 161 2.539
112 2.560(55)
371 2.560(55)
191 2.560(55)
132 2.479(5)
202 2.449(25)
042 2.449(25)
1.11.0 2.406(15)
222 2.406(15)
46 1 2.406( 15)
062 2.263(30)
312 2.263(30)
2.10,l 2.263(30)
620 2.206(3)
570 2.206(3)
337 2.206(3)
640 2.125(7)
2,12.0 2.125(7)
4,100,O 2.125(7)
b = broad; sh = shoulder
100

Table 2-28
Chemical Compositions and Structural Formulas of Some Sepiolite Samples.
1 2 3 4 5
SiO, 53.98 52.50 56.1 52.43 45.82
A1 2 0 3 0.20 0.60 0.42 7.05
TiO, 0.05 0.05
Fe,O, 2.90 0.20 2.24 21.70
FeO 0.01 0.70 0.05 2.40 0.20
MnO 0.17
MgO 22.80 21.31 24.3 15.08 12.32
CaO 0.04 0.47 0.34 0.90 0.90
Na,O 0.58 1.13
K,O 0.16 0.11
NH,
H,O- 11.54 12.06 10.0 10.48 9.48
H,O+ 8.46 9.21 9.21 9.45 9.41
Tot a1 97.77 99.75 99.* 99.71 100.05

Octahedral
Al 0.05 0.06 1.37
Fe3+ 0.47 0.03 0.37 2.69
Fe2+ 0.13 0.44 0.04
Mg 7.60 7.14 7.22 4.90 4.20
Mn 0.03
c 7.65 7.14 7.81 7.18 6.96
Cation Charge 15.95 15.71 15.90 16.61
Tetrahedral
Al 0.16 0.04 0.48
Fe3+ 0.04 1.05
Si 12.09 11.80 11.96 11.52 10.95
Interlayer
Ca 0.01 0.11 0.15 0.22
Na 0.26
K 0.04
NH.4 0.43
* Includes 0.12% P20,, 0.27% CO,.
1. Post, 1978: Sepiolite, Amargosa Playa, Nevada.
2. Caillere, 1936: Sepiolite, Ampandrandava, Madagascar.
3. Hathaway and Sachs, 1965: Sepiolite, Mid-Atlantic Ridge.
4. Rogers et a/., 1956: Aluminous sepiolite, Tintinara, South Australia.
5. Caillere, 1936: Xylotile, Sterzing Tyrol.

for the half-unit cell: Si,Mg,0,,(OH),(OH,),4H20. The structural formulas were


calculated on the basis of 21 oxygens per half-unit-cell. The octahedral cation
population for most samples is close to 4.0 (Weaver and Pollard, 1973). The Ht
content of the octahedral sheet is so speculative that it is impossible to make a
reasonable calculation of the layer charge. The tetrahedral aluminum ranges from
0.01 to 0.69 per 8 tetrahedral positions, whch is similar to the range for smectites.
101

Octahedrally coordinated aluminum and total A1,0, are less than that found in
montmorillonites. The magnesium content of the octahedral sheet and MgO is 2 to 4
times as abundant as in montmorillonite. The iron contents are similar.
A unique feature of palygorskite is that for most samples the octahedral R’+/R’+
ratio is near 1.0. The R7+cations are predominantly Al and the R2+cations Mg. In
the sheet structure physils either R”+(small) or R2+(large) cations are dominant in
the octahedral sheet. This is not the case for most palygorskites. I t appears likely
that the presence of large and small cations in roughly equal proportions produces
enough lattice distortion that the tetrahedral sheets cannot easily adjust and are

Table 2-29
Chemical Compositions and Structural Formulas of Palygorskite Samples.
1 2 3 4 5
SiO, 53.64 52.35 57.19 57.17 69
8.76 15.44 12.13 13.25 13
Fe203 3.36 2.12 1.51 6.25 5
FeO 0.23 - - -

TiO, 0.60 - - 0.36


CaO 2.02 0.14 0.36 0.69 1
MgO 9.05 6.60 9.29 7.29 11
Na ,O - - 0.15 -

K2 0 - - 0.07 ~
1
H,O+ 10.89 12.00 - -

H2 0 - 9.12 10.32 - 15.04


Total 97.67 * 98.97 100.24 99.85
B.E.C. meq/100 g 21 36

Octahedral
Al 1.29 2.25 1 .X7 1.73 1.38
Fe-” 0.37 0.23 0.16 0.63 0.49
Fe2+ 0.03
Mg 1.96 1.44 1.91 1.45 1.87
E 3.65 3.92 3.94 3.81 3.74

Tetrahedral
Al 0.21 0.39 0.1 1 0.36 0.13
Si 7.79 7.61 7.89 7.64 7.87

In terlayer
Ca 0.32 0.32 0.11 0.20 0.02

Octahedral
AI/W 0.66 1.57 0.98 1.19
R3+/R2+ 0.83 1.72 1.06 1.63

* P,O,, K,O, Na,O. MnO total 2.40%.


1. Bradley, 1940: Attapulgus, Georgia. U.S.A.; short. perimarine
2. Stephen, 1954: Bakkasetter, Shetland.
3. Huggins el a/.,1962: Metaline Falls, Washington; long.
4. Weirsma, 1970: Jordan Valley, Israel; short.
5. Desprairies. 1981: Hydrothermal marine. Pacific Ocean.
102

forced to invert over an interval of five octahedral cation positions (Weaver and
Pollard, 1973).
The cation exchange capacities of the more pure samples, some smectite is
commonly present, are in the range of 5 to 30 meq/100 g. Ca is present in all
samples, and it appears that some of it is in an octahedral position (Weaver and
Pollard, 1973).
The literature on the occurrence and origin of palygorskite has been reviewed by
Weaver and Beck (1977), Singer (1979), and Callen (1984). In contrast to sepiolite,
which is primarily a Mg silicate, palygorskite contains an appreciable amount of Al.
This usually requires an Al-rich precursor (transformation), commonly smectite
(Weaver and Beck, 1977); however, it does form from solution (neoformation) in
soils and shallow water environments where Al does not have to be transported very
far.
The more extensive palygorskite deposits commonly occur in coastal areas,
perimarine environments, where it formed in slightly basic (pH 8 to 9) shallow
water, brackish lagoons, lakes, and bays (Weaver and Beck, 1977; Callen, 1984). I t
forms in alkaline lakes (Callen, 1977) and in soils in arid and semi-arid regions
(Singer, 1979). Detrital palygorskite (Weaver and Beck, 1977) and hydrothermal
palygorskite (Bonatti and Joensuu, 1968; Bowles et al., 1971; Desprairies, 1981;
numerous other Deep Sea Drilling Reports) are relatively abundant in Mesozoic and
Cenozoic marine sediments. Some authors (Couture, 1977; Kastner, 1981) believe
palygorskite forms authigenically in a normal marine environment. Additional
information on the origin of palygorshte can be found on the following pages:
Soils: pages 182-185
Lacustrine: page 276
Marine: pages 379-387
Brackish: pages 400-407
103

Chapter 111

SOILS AND WEATHERING

SOIL FORMATION

“Weathering is the response of materials within the lithosphere to conditions at


or near its contact with the atmosphere, the hydrosphere, and perhaps still more
importantly, the biosphere” (Reicke, 1945, modified by Keller, 1962). Weathering
processes are classed as either physical or chemical. Biological processes can be
considered as either physical or chemical.

Physical Weathering and Erosion

Physical weathering involves processes which cause in situ fragmentation or


comminutation. These include: expansion due to unloading, thermal expansion,
crystal growth, colloidal plucking, and the activity of plants and animals. Physical
weathering is considered to be a minor process compared to chemical weathering.
However, sedimentologists are primarily concerned with the dual processes of
physical weathering and erosion. Even though they are considered separately,
erosion can often be considered a form of physical weathering. Thus, the plucking
and grinding action of glaciers is erosion but also could be considered weathering.
Glaciers are effective in eroding, grinding and transporting unaltered physils.
Glaciers occur in cold regions where chemical weathering is at a minimum and they
are able to scour fresh rock from beneath the thin soil zone. Some of the fine
material in the glacial outwash is in turn eroded by the winds and redeposited as
loess in areas far removed from the site of the glacier. In deserts, wind is an
important agent of both weathering and erosion.
Even in regions with temperate climates, streams may transport more un-
weathered material than weathered soil material. The relative amounts of weathered
(chemical) and unweathered material carried by a stream depends on the thickness
of the soil profile, which in turn is related to climate and relief. Thus, in some
mountain streams in the northern Appalachian Mountains, where the soil is thin
and protected by lush vegetation, the physils (illite and chlorite) carried by the
streams are largely derived physically, by the stream, from the bedrock shales.
Farther south where the soil profile is much thicker, the physils (kaolinite) are
primarily derived from the upper soil zone. The type of bedrock is obviously a
factor. Shales, slates, and schists will supply relatively more unweathered physils
104

than will igneous rocks. In the latter rocks, chemical weathering is required to
produce the clay physils.
Based on statistical analysis and the crystallinity index of illite, Potter et al.
(1975) concluded that the “vast majority” of the physils transported by the
Mississippi River were derived from bedrock rather than from soils. Gibbs (1967)
showed most of the physils arriving at the mouth of the Amazon River were a
product of physical weathering processes in the Andes Mountains (p. 197). Physils
produced by chemical weathering are dominant only in the relatively small coastal
streams. The distribution and composition of detrital clays in ocean sediments
indicate that the majority were probably derived, relatively unweathered, from the
bedrock (Chapter V). This is particularly true in the high latitudes. In the region of
the humid tropics the transported physils can be predominantly chemical in origin,
but to a large extent this depends on relief and the extent of the river system.
Most rivers carry a continuous spectrum of weathered and unweathered physils.
The extent of chemical weathering is not always easy to determine. In some
instances a new physil is created, and at the other extreme, only a small amount of
interlayer material may be stripped from the bedrock physil, or the oxidation state
of the iron may be changed. K can be stripped from a mica or illite during
weathering only to be replaced (rejuvenated) in the marine environment. The end
product is similar to the original mica and illite, and evidence for it having been
chemically weathered can probably only be indicated by K-Ar analysis.
While we are on the subject of erosion, it should be pointed out that interpreting
the climatic significance of physils transported from a chemically weathered soil can
be difficult. Most soils have a weathering profile in which weathering effectiveness
decreases with depth. From top to bottom a soil could contain kaolinite, smectite,
I/S, and illite. As most rivers, at least in their upstream portion, erode to bedrock,
the entire soil profile is removed, not just the upper kaolinitic zone. This makes it
difficult to interpret the paleoclimatic significance of the detrital physil suite;
however, a complex, multicomponent physil suite commonly indicates weathering
and climatic conditions were moderate.

Chemical Weathering

Soil
Chemical weathering, which alters and produces physils, commonly creates a soil,
so it is necessary to discuss briefly the nature of soils. Soil has a wide variety of
definitions, depending upon the interest of the examiner. The most applicable
definition for the clay mineralogist is one proposed by Birkeland (1974):

“...a natural body consisting of layers or horizons of minerals and/or organic


constituents of variable thicknesses, which differ from the parent material in their
morphological, physical, chemical, and mineralogical properties and their biologi-
cal characteristics.”
105

C A

Bt

c
c

wid matorial before Thin solurn, organic Organic matter content Very acid In reaction,
(01 devebpment matter accumulation in is at a maximum Has severely weathered,
m A honzon, from which moderate clay and has lea8 organic
cattonates have been awumulatlon In Me B matter than mature
leached Minimal honzon and the solum stage. Clay
weatheflng and 16 acid Stage of accumulation in B
eluvlalion. maximum pmductwky horizon has lormed a
b r corn. clay pan. An E
horizon exislo.

Fig. 3-1. A summary of the stages in the development of soils in the central United States under tall-grass
vegetation. From Foth and Turk, 1984, Fundamentals of Soil Science. Copyright 1984 John Wiley &
Sons, Inc. Reprinted by permission of John Wiley & Sons, Inc.

Soil Profile
A soil profile consists of the vertical arrangement of all the horizons down to the
parent material (Fig. 3-1). The A, or dark surface, horizon is defined as the zone of
organic accumulation and eluviation. The zone is depleted of material that is carried
downward either in suspension or in solution by percolating waters. Organic matter
gradually accumulates and with time reaches a steady-state condition where the
gains equal the losses. Some of the organic material is converted to organic acids,
which play a major role in complexing ions and increasing their solubility and
mobility. Chemical weathering and the formation and alteration of physils are most
pronounced in this horizon. In soils with thin A horizons, a lighter-colored layer (E),
low in organic matter, may develop below the A horizon. The B horizon is the zone
of illuviation where material derived from the A and E horizons is deposited and
precipitated, e.g., physils, calcite (caliche), Fe and A1 oxides. Primary silicates may
be altered and new physils formed in B. The C horizon is a transitional zone of
partly altered rock or saprolite overlying the unaltered bedrock or (R) horizon.
Organic material is concentrated in the A horizon and to a lesser extent in the B
horizon. The agriculturalists use the term “soil” or “solum” to refer to these two
106

horizons and consider the C horizon to be virtually unaltered and essentially


“bedrock”. Chemical weathering extends into the C horizon (e.g., where the bedrock
(R) is a mica schist the C horizon commonly contains vermiculite), and I will use the
term “soil” in the broader context to include horizons A, B, and C.
The A and B horizons vary considerably in thickness but are generally relatively
thin compared to the C horizon. The total AB thickness is commonly less than one
meter. The A horizon is usually less than one meter and the B horizon less than two
meters. In well-leached tropical and subtropical areas the C horizon can be more
than 30 meters thick.
Immature or young soils are characterized by organic-matter accumulation in the
surface zone and little weathering or leaching. Only the A and C horizons are
present. When the B horizon has developed, the soil is considered to be mature.
With increased weathering, the A and B horizons become highly differentiated and
the soil is called an old soil (Fig. 3-1). Physil and oxide (Fe, Al, Ti) development is
greatest during the final two stages, with oxides commonly more abundant than
physils in old soils.
The crystallization of new physils and the alteration of existing physils are not
necessarily related to the maturity of a soil. Pre-existing physils can be transported
downward and accumulate between the A and C horizons, to form a B horizon,
without being significantly modified. For example, such a situation could occur
when the parent material was a kaolinitic sand.
In the early stages of weathering (immature soil), the parent rock will influence
the type of soil and physils that will form. With increasing maturity, climate and
drainage conditions become increasingly important and the same physils can form
from a wide variety of rock types.
The rate of chemical weathering and soil formation is largely controlled by
climate. Regardless of the climate, the rate is relatively slow. Glacially polished
rocks in cool temperate regions have weathered hardly at all in the past 10,000 to
20,000 years. At the other extreme, in a tropical climate volcanic glass from the
Krakatau explosion lost 5% silica and was partially converted to clay in 50 years. In
the temperate, arctic, and arid regions there has been very little mineral alteration
during the Holocene. What mineral alteration is present occurred during Pleistocene
or earlier, when the climate was generally more humid than it is now (Hunt, 1972).
Obviously, the rate of soil development and the extent of chemical weathering
depends on the rate of erosion.
Efforts have been made to determine the age of soils by obtaining radiocarbon
dates using organic and inorganic carbon. The significance of the dates is difficult to
determine because of the mobility of both materials and the solution and re-precipi-
tation of the CaCO,. Radiocarbon dates obtained on CaCO, in a desert soil near
Las Cruces, New Mexico, illustrate the range of dates in a soil profile. The whole
soil date from near the base of the B horizon is 5,725 years. Various carbonate
layers at the B-C interface have dates ranging from 4,575 to 18,300 years. The whole
soil of the C horizon has a date of 15,300 years and pebble coatings a date of 27,900
years. It was concluded that soil formation began about 30,000 years ago (Gile et
al., 1970).
107

In the United States superimposed soil profiles are the rule rather than the
exception (Hunt, 1972). The older profiles are usually more intensely weathered as
pre-Pleistocene climates were generally warmer and more humid. Due to climatic
changes the profiles may contain different physils. Most of the ancient soils formed
in pre-Wisconsinian time (70,000 yr. B.P.) - some as long ago as the Cretaceous
( - 10' yr. B.P.). In a typical soil sequence a pre-Wisconsinian soil is highly
weathered, 5 m thick, and contains an abundance of physils formed by chemical
weathering. This soil is overlain by a Wisconsinian soil 1 m thick, which in turn is
overlain by a Holocene soil less than 1 m thick. Though the younger two soils are
leached, no new physils were created; those present are inherited from the pre-
Wisconsinian soil.
In the southeastern United States saprolites, usually kaolinitic, formed during the
Cretaceous and early Tertiary are commonly over 30 m thick. These are overlain by
modern soils a few centimeters to a few meters thick.
We see from this brief review of soils that long periods of time are required for
physils to form in soils; superimposed soil profiles are common and may represent
as much as 100 m.y. of weathering. Thus, physils eroded from a complex of soil
profiles may have formed millions of years before the time of erosion and deposi-
tion. This complicates paleoclimatic interpretations based on detrital physil suites.
Much of the kaolinite transported to the Atlantic Ocean by rivers in the southeast
United States crystallized approximately 100 m.y. ago. Soil scientists tend to
emphasize the formation of physils from feldspars and other non-physil minerals,
which increases the clay content of the soil. However, even through n o new physils
are created, relatively short-term leaching can modify existing physils such as
chlorite, illite, and mica, generally increasing their expandability and decreasing
their size.

Soil Classification
The information in this section was obtained largely from Loughnan (1969),
Hunt (1972), Birkeland (1974), Wilding et al. (1983) and Foth (1984).
Soils are classified as zonal, intrazonal, or azonal. Zonal soils (great soil group)
are mature soils which are extensive, and their profiles correspond to specific
climatic and vegetation zones. Intrazonal soils also have mature profiles but are
local in extent and differ from the prevailing zonal soils because of local factors
such as rock type or relief. Azonal soils have poorly developed profiles and include
such things as alluvial deposits and sand dunes. Zonal soils are our chief concern.
Zonal soils are broadly divided into pedocals and pedalfers. The pedocals are
characterized by the presence of carbonates and form in the more arid regions ( < 60
cm rainfall). Carbonates have been leached from the pedalfers, and they contain
concentrations of A1 and Fe oxides; they develop in humid climates. Fig. 3-2 is a
generalized diagram showing the relation of the great soil groups to climate. The
differences between the various pedalfer soil groups is largely due to differences in
temperature, whereas differences among the pedocals is largely due to differences in
rainfall.
I
108

y cold We! cold


I

Gray-Wooded
Podzol

;ray Desert Brown Chestnut Chernozem


Brown Podzolic
Sierozem
Brunizem Gray-Brown Podzolic

-
Red Desert

Red-Yellow Podzolic
Reddish
Chestnut
Reddlsh
Prairie Yellowish Brown Latosolic

ry hot
Latasol
Wet hot
I
PEDOCALS PEDALFEAS

Fig. 3-2. A diagrammatic visualization of the relation between climatic conditions and the occurrence of
the great soil groups 1949 classification). From Foth and Turk, 1972, Fundamentals of Soil Science.
Copyright 1972 John Wiley & Sons. Reprinted by permission of John Wiley & Sons, Inc.

In 1960 the Soil Survey Staff (1960) of the U.S. Department of Agriculture
recommended a new soil classification based mainly on observable properties and
less on genetic concepts; this classification was further modified in 1975. Table 3-1
compares the old and new classifications and lists some physil associations. For a
more detailed, but no more definitive, compilation of the mineralogy of soil types,
see Wilding et al. (1983), Table 6.3.
Fig. 3-1 is a generalized cross-section showing the nature of soil profiles as a
function of climate and vegetation. The cross-section extends from the Hudson Bay
region (cold and wet) to the southwestern United States. The soils average progres-
sively younger towards the northeast (Hunt, 1972). Note that clay accumulation in
the B horizon is restricted to the acid, wet soils.
The soils in mountain areas can have a vertical zonation pattern similar to that
shown in Fig. 3-3. Thus, the soils on the mountains of the western United States can
be Desert or Brown soils at the base of the mountain and grade upward, with
increasing rainfall, into Chesnut and, higher, Podzolic soils. The summit can contain
some tundra-like soil. These various soils can have different ages. In the Rocky
Mountains the summits were covered by snow or ice during late Pleistocene time
and the soils are of Holocene age. The soils on the flanks are generally older but no
older than the Wisconsinian glaciation because most of the older soil below the ice
cap was removed by mass-wasting. Pediments and fans at the foot of the mountain
contain extensive pre-Wisconsinian soils (Hunt, 1972). Detrital sediment derived
from such mountains would be expected to contain a wide variety of physils.
109

Table 3-1
Dominant Clay Mineral Types in Soil Groups (after Jackson, 1964).
Order ’ Great Soil Group * Dominant Physil Minerals ’
Oxisol Laterite soils and Latosols Sesquioxides. gibbsite, kaolinite,
2 : l to 2:2 intergrades
U I tisol Red-Yellow Podzolic Kaolinite, halloysite. vermiculite,
2:l to 2:2 intergrades,
sesquioxides. gibbsite

Spodosol Podzol Sesquioxides, interstratified


layer silicates, 2:l to 2:2
intergrades. mica
Vertisol Grumusols Montmorillonite

Alfisol Gray-Brown Podzolic, Planosol Mica, montmorillonite, 2:l to 2:2


intergrades. chlorite, kaolinite
Mollisol Prairie, Chernozem, Chestnut, Montmorillonite, mica, vermiculite,
Rendzinas chlorite
Inceptisol Ando Allophane
Brown Forest, Sol Brun Acide Mica, interstratified layer
silicates

Aridisol Desert. Sierozem Mica. vermiculite. interstratified


layer silicates. chlorite
Entisol Azonal soils Highly variable
Tundra soils Mica
Humic Gley Montmorilloni te
’ New soil classification.
Older soil classification.
Each mineral given has been reported dominant in the clay of one or more soil profiles.

Zonal Soils
Soil scientists have fairly well established the characteristic physils present in the
great soil groups. In general, as can be seen in Table 3-1, the great soil groups do
not have distinctive physil suites. In part this is because a relatively high proportion
of inherited physils are are present in all but the most highly weathered soil profiles.
Entisols: Entisols are poorly developed soils with properties determined largely
by the parent material. The weak development is due to youthfulness or extreme
wetness or dryness, e.g. river alluvium. Chemical weathering is barely detectable.
Vertisols: Vertisols are dark colored soils that occur in tropical and subtropical
climates. Horizontal zonation is weakly developed because of “self mixing” result-
ing from the shrinking and swelling of the clay with drying and wetting. The
dominant physil is montmorillonite. The montmorillonite may be inherited or
formed by weathering basic igneous and metamorphic rocks in low relief areas
where rainfall is moderate and/or seasonal so that leaching is restricted. Some
kaolinite is always present and the amount is presumably proportioned to the
weathering intensity.
110

Dry, hot Wet, cold

Desert Brown Chestnut Chernozern Prairie Gray Brown Podzol Tundra


SOllS soils SOllS Soils Soils Podzolic Soils SOllS SOllS

Desert shrubs Short grass Tall grass Broadleaf forest Conifer forest Tundra

100th

Calcium carbonate accum Clay iron accumulate


Limit of soil water penetration Carbonates removed

Alkaline soils, 1 Neutral ~ Acid soils.


closed system ~ soils 1 open system
P Y

Fig. 3-3. Transect illustrating changes in soil profiles that accompany changes in vegetation and climate
between the tundra in northern Canada and the deserts in southwestern United States. At the 100th
meridian, the annual precipitation averages about 20 inches; there and to the west the soils are alkaline.
The easternmost grassland soils are about neutral; further east the soils are acid. From Hunt, 1972,
Geology of Soils. Copyright 1972 W.H. Freeman and Co. Reprinted with permission from W.H. Freeman
& co.

Mollisols: Mollisols are characterized by a dark colored surface horizon with


greater than 50% base saturation (mollic epipedon). These soils commonly border
desert regions and are grass covered. Rainfall is sufficiently limited to prevent
excessive leaching. Carbonates are present at a depth of 2 m. Many of the Mollisols
in the United States are developed on tills and loess. The 2:l physils are predomi-
nant and have been only slightly modified; in general, the expanded layers in illite,
I/S, and chlorite are increased.
Aridisols: Aridisols occur in dry regions, mostly deserts. The soil-forming processes
are similar to those of humid regions; however, the rate of development is much
slower. Pedogenic horizons are developed that have been either enriched or depleted
of substances by moving water. Physils, calcium carbonate, soluble salts, and
sometimes gypsum accumulate in the B horizon (argillic horizon). Winds pay a
major role in the development of Aridisols. Winds move dust from some areas and
deposit it in others.
The physils in Aridisols are believed to be largely inherited and to have formed
by Pleistocene weathering, though Walker (p. 157) has shown that at the present
time smectites are forming from the alteration of ferromagnesium minerals.
Montmorillonite is commonly the most abundant physil in the B horizon; illite and
mica are characteristic of the A horizon. The 10 A physils are believed to form
111

because of a high concentration of K f and NH: produced by plant cycling; the


high temperatures and desiccation could aid in the collapse of the expanded layers
(Nettleton et al., 1973).
Spodosols: Spodosols form in cool, temperate, humid regions. Most occur just
south of the tundras on glacial material. Typically, they have four major horizons: a
dark-colored organic surface horizon; a bleached eluvial horizon; a reddish, brow-
nish or black illuvial horizon enriched in amorphous material and a sandy C
horizon. The A and B horizons are very acid. The translocated material consists
mostly of poorly crystalline Fe and A1 oxides and Fe-Al-organo complexes. The
physils are a complex mixture containing transformation products of mica and
chlorite, primarily smectite and vermiculite with hydroxy interlayers.
AIfsols: Much of the world's farming is done on Alfisols. These soils are
characterized by prominent horizons of eluviation and illuviation. They form in a
humid, continental type climate that alternates from wet to dry. Physils are
translocated from the A and/or E horizon and accumulate in the B horizon. The
physils vary greatly, ranging from montmorillonite to kaolinite. Illite tends to be
altered to vermiculite with and without hydroxy interlayers.
Ultisols: Most Ultisols are more weathered than Alfisols. They lie almost entirely
between latitudes of 40'N and 30"s. They form where evapotranspiration rates
exceed precipitation and where precipitation exceeds the capacity of the soil to
retain water during some seasons (leaching). The horizons are well developed.
Kaolin and vermiculite and smectite with hydroxy interlayers are the major physils,
though a variety of weathered products are present in various locations.
Oxisols: The oxisols are the most weathered of all soils and are mainly in humid,
tropical areas. Where present outside the tropics they are believed to have formed
during past tropical weathering periods. Oxisols are generated where climatic
conditions are humid enough to remove the soluble weathering products and
produce a residual concentration of sesquioxides, gibbsite, and kaolinitic physils.
Physil translocation is insignificant and horizon differentiation is minor. Organic
materials are rapidly oxidized or destroyed by microbial action and little organic
material moves down the profiles.
Fe is precipitated near the surface (B horizon) as goethite and hematite. Varying
amounts of gibbsite and kaolinite are commonly present. The Fe oxides commonly
form concretions or pisolites. The true laterite (iron stone) hardens when dried. In
general, the Fe oxides are concentrated (as much as 50% Fe,O,) nearest the surface;
gibbsite is concentrated beneath the Fe zone and is underlain by a kaolinite-rich
zone. Various 2:l physils may underlie the kaolin zone. However, gibbsite may be
the predominant mineral in the near-surface horizon (bauxite).
Inceptisols: In Inceptisols the development of genetic horizons is just beginning,
but they are older than Entisols. They show little evidence of eluviation or
illuviation. They form under a wide variety of climatic conditions. They form on
geologically young sediments or landscapes and where environmental conditions
inhibit soil-forming processes. Due to the great variation in parent materials and
environmental conditions under which they form, Inceptisols have extremely diverse
mineralogies. Mixed physil suites and montmorillonite are the most abundant but
112

Poludrm Tvndri

7 2400
5 2100
E 1800

’......._.._._..____._.....
.
300
0 g o

DFresh rock 0 K.olini@


E l Little ehemiul alteration EllN203
1=3 illite-montmorillonitc Fs203 + N 2 O 3

M o i i t u m and temperature
9 both diminish to the north;
hence, weathering b dow,
g
(*

omanie matm decomposition


b dow, and either the slight High mokiurc promoles high High Lcmpratum aids rhemi- Heh moislun and temwm-
-g preipitition (palar desert) or Caching. relatively lox tcm. -1 weathering. but low moir lure rnult in rapid Weathering
g5 prmllrost (tundra) inhibit peratwe msult. in relatively tum i n h i b i t r v ~ b t d i o n , o m i n . and high rile of leaching or
3 I c a e h i n g o I m o b i l e bw rate or organic matter
mnslilucntr decomposition
IPmaltQr buildup. and leach-
ing or mobile eonrtiluents
mobile eonitituenls

-6 rnnptlwr Spodorol. AlfmoI. Molluol Aridid lntlrol. Oxuol


82

Fig. 3-4. Diagram of relative depth of weathering and weathering products as they relate to some
environmental factors in a transect from the equator into the north polar region. From Birkeland, 1974.
Pedology, Weathering and Geomorphological Research. Copyright 1974 Oxford University Press. Taken
mostly from Strakhov (1967).

allophane, halloysite and kaolinite can be abundant in soils developed on volcanic


material.

CLIMATE (AND OTHER FACTORS)

Introduction

The relation of physil type to climate is, at best, vague. Kaolin minerals tend to
develop preferentially in the humid tropics and illite, chlorite, and smectite to be
essentially unweathered in the arctic regions. Between these two extremes a wide
variety of physils can be found. Fig. 3-4 shows the accepted regional climate-soil-
physil relation.
Before proceeding, two major problems should be mentioned. First, it is often
difficult to determine whether physils are inherited or neoformed (or deposited from
the atmosphere). Second, weathered profiles which developed for thousands and
millions of years have probably been subjected to a variety of climatic conditions.
Climate is commonly considered to be the most important factor in determining
the properties of soils. Other important factors are parent material, topography,
organisms, and time (Jenny, 1941). Moisture and temperature are the two most
important factors influencing chemical weathering. Moisture is involved in most of
113

the physical, chemical, and biochemical processes that go on in a soil. In general, the
more moisture, the more intense the weathering and leaching and, to some extent,
the formation of physils. Temperature influences the reaction rate. Van Hoff‘s rule
states: “For every 10°C rise in temperature, the velocity of a chemical reaction
increases by a factor of two to three.” Weathering is at a minimum where the
climate is warm and dry, cold and dry, or cold and moist.
Though mean annual rainfall values are a general guide to weathering intensity, it
is important to know the seasonal distribution of the rainfall. Whether the precipita-
tion maximum coincides with the temperature maximum or minimum or whether
the rainfall is relatively uniform will influence the extent of leaching, growth of
plant material, and, to some extent, the type of physil that will be formed. When the
temperature and precipitation maxima coincide, particularly when both are high,
leaching is aggressive and kaolin minerals (along with Fe and A1 oxides and
hydroxides) will normally be formed as a weathering residue. When the maxima are
out of phase, mild leaching will occur in the wet period. Evaporation during the
warm, dry period can increase the concentration of ions in solution to the extent
that physils such as sepiolite and palygorskite will crystallize (as well as calcite,
dolomite, gypsum, etc.).

Clay Content

We are concerned with both the amount of clay and the type of physils formed
under varying conditions. Most soil studies report the clay content of soils as the
amount of < 2 pm material. In addition to physils, this material includes quartz,
carbonates, oxides, etc., though commonly the physils are the most abundant
component. In general, the amount of clay formed increases with increasing
temperature and/or moisture.
Two types of clay values are determined: the amount of clay present and the
amount of clay formed in the soil. When the parent rock is an igneous rock there is
no problem; when it is a sedimentary rock, loess, etc., there is a problem in
determining how much clay was inherited and how much was created withn the
soil. There is the added problem that much of the dissolved material can be leached
completely from the soil, rather than precipitating in the soil. Thls is particularly
true of the lateritic soils.
Using chemical techniques Barshad (1957) calculated how much clay was formed
on a series of soils that were selected to demonstrate the effectiveness of the various
factors involved in clay formation. Though these are just a few examples they
demonstrate the general principles.
Fig. 3-5 presents data obtained from a series of pedalfer soils. The soils were all
developed from granite, under conditions of similar rainfall but increasing tempera-
ture (from left to right). Not only does the amount of clay formed in the
near-surface horizons increase with increasing temperature, but the depth of
weathering, or clay formation, increases with temperature. Barshad concluded that
clay formation in the pedalfers is in the following order: Podzols < Gray-Brown
114

clay formation in grom/i00 gram parent material


I0 ZO 30 40 50 60
0
i I '

80

Great soils from


Granitic parent material

110 G. Br. 9 Gray Brown

Fig. 3-5. Clay formation in several of the Great Soil Groups developed from granitic parent material.
From Barshad, 1957. Reprinted with permission from Clays Clay Minerals, 2, 110-132. Copyright 1959
Pergamon Journals, Ltd.

Podzolic < Brown Earths < Prairies < Red and Yellow Latosols < Laterites (see Ta-
ble 3-1 and Fig. 3-2.)
Similarly, at constant temperature an increase in precipitation causes an increase in
clay formation.
In evaluating the effects of topography, Barshad found that more clay formed in
poorly-drained soils than in well-drained soils. This suggests that a higher propor-
tion of the more soluble material released by weathering, and perhaps some clay,
was leached from the soil in the well-drained areas. This observation does not
necessarily mean that more clay is formed in the low-relief environment; it may
merely indicate that more clay remains in the soil and more clay accumulates in the
downslope environment.
Numerous examples of slope studies, reviewed by Birkeland (1974), indicate the
clay content and the soil moisture content increased in a downslope direction. The
115

higher moisture content accelerates clay formation. In the tropics the distribution of
Fe- and Al-rich soils can be related to topographic position. A1 tends to accumulate
in the upland soils, whereas the more mobile Fe and Mn are transported downslope
and precipitated in the lower soils (Hamilton, 1964).
From the sedimentological standpoint the situation is a little different. In a
drainage basin, relief generally increases with elevation. Both high relief and high
elevation tend to cause a decrease in clay formation; however, the rate of erosion is
much higher and a much higher proportion of the clay is delivered to the deposi-
tional basin. The high erosion rate increases the amount of exposed fresh rock
available. As fresh rock should weather more rapidly, the amount of clay formed per
unit of time may be larger than that formed in the highly weathered lowland soils
where erosion is at a minimum. Physil weathering is also more active in the high
relief areas. For example, in the Amazon drainage basin 84%of the total amount of
dissolved salts ( < 50%) and suspended solids delivered to the Atlantic Ocean is
derived from the Andean Mountains, which comprise 12% of the drainage basin
(Gibbs, 1967).
The type of vegetation has an influence on the amount of clay formed in a soil,
The forested A horizons are relatively thin and the amount of organic material
decreases rapidly with depth, whereas grassland soils have relatively thick A
horizons and the organic content remains relatively high for a considerable depth.
The differences are due to the fact that forested soils obtain their organic input from
litterfall, whereas in grassland soils the input is from litterfall and roots. Barshad
(1957) found that more clay was formed in grassland than in forested soils and that
the differences extended throughout the profile. White and Riecker (1955) found
that the surface layers of grassland soils have a higher clay content (not necessarily
“formed”) than forested soils, but that forested soils had a high proportion of
translocated clay (to the B horizon). Both studies were of mid-continent soils.
The time factor has been difficult to quantify. Under most conditions the amount
of clay and the thickness of B horizon clay increase with time. As the soil gets older
it grows in depth with weathering extending into the C horizon (Barshad, 1957;
Birkeland, 1974).
During the early stages of weathering the composition of the parent material can
determine the type of physil that will form. Less is known about the influence of
parent material on the amount of clay material that will form. Factors other than
mineral composition of the parent rock can influence the amount of clay that will
form. Some of these factors are: texture, porosity and density, structure and fabric,
and degree of consolidation. Barshad (1957) found that in general basalts and
schists yield more clay than granites. In general, for igneous and metamorphic
rocks, basic and fine-grained rocks yield more clay than acid and coarse-grained
rocks.
From the sedimentologist’s point of view, another factor needs to be considered:
aggregation. Most of the data on the amount of clay ( c 2 pm material) in soils is
obtained by calculations based on chemical data or by thoroughly dispersing the
soil, particularly breaking aggregates. Much of the clay- and silt-sized material in
soils occurs in sand-sized (or coarser) aggregates, which are cemented by organic
116

material, Fe and Al oxides, and hydroxides. These aggregates are commonly firmly
bonded; when a soil is eroded by either wind or water, much of the clay (and
physils) and silt is actually transported as sand-sized aggregates. These aggregates
should be deposited in the high-energy environments along with the sand rather
than in the low-energy mud environments. In this respect they resemble fecal
pellets.
The percentage of aggregates in the A horizon depends on the amount of primary
clay- and silt-sized material and the availability of cementing material. The per-
centage of aggregates is low ( - 10%)in desert soils because of a small clay content.
As the rainfall becomes greater, clay formation increases and the percentage of
organic matter increases. Thus, Chernozem soils have about 50% of the total silt and
clay aggregated. The Podsols have a relatively low aggregate content (10 to 25%),
probably due to removal of Al and Fe from the A horizon. Where the humidity is
high, aggregation increases with increasing temperature. True laterites contain as
much as 95% Al- and/or Fe-rich aggregates. Dehydration of the hydrated oxides of
these elements produces very stable aggregates (Jenny, 1941).

Climate and Physil Types

First I will present some generalities on the relation of physils to climate and
then some specific examples and exceptions. Barshad (1966) analyzed a series
( > 400) of surface soil samples along a transect extending from Death Valley to the
northern border of California (Fig. 3-6). By collecting samples at various elevations
he was able to restrict the variation in mean annual temperature of soil formation to
50" to 60°F. The < 2 pm fraction was analyzed.
Except for illite, the distribution patterns of physils formed on acid and basic
igneous rocks are similar. Montmorillonite is present in soils developed in areas
where the precipitation is less than 40 inches and is most abundant where precipita-
tion is less than 10 inches. In the latter case it is usually associated with the
accumulation of CaCO,. Kaolinite-halloysite, vermiculite, and gibbsite are abun-
dant in areas where precipitation is greater than 40 inches (pH of soils is neutral to
slightly acid). Kaolinite-halloysite is the dominant clay physil in all soils formed
where the precipitation was greater than 20 inches. Presumably because of the
availability of K, illite was found only where the parent material was acid igneous
rocks. It occurs in the low precipitation region. Though illite is not present in the
basic igneous rock soils, montmorillonite is relatively more abundant over the
precipitation interval in which illite develops. This could indicate the development
of an Al-rich montmorillonite (beidellite).
In the soils formed from the acid igneous rocks, vermiculite is an alteration
product of micas; and in soils from areas with precipitation ranging from 35 to 45
inches, mixed-layer M/V often occurs.
Depth has little effect on the frequency distribution of physils in soils from areas
with less than 10 inches or more than 60 inches of precipitation; however, in the dry
area CaCO, and in the wet areas gibbsite are concentrated more in the subsoils than
117

CLAY MINERALS IN SURFACE LAYER OF RESIDUAL


SOILS FROM ACID IGNEOUS ROCKS
z? 80-
I-
z
W
KAOLlNlTE
c
z
2 60-
-I
4
a
W
t
WINTER
zI 40 - PRECIPITATION

G
-I
TEMPERATURE
V
50.F to 60.F
W
2 20-
l-
a
J
W
a
20 40 60 80
MEAN ANNUAL PRECIPITATION: INCHES

CLAY MINERALS IN SURFACE LAYER OF RESIDUAE


SOILS FROM BASIC IGNEOUS ROCKS

20 40 60 80

MEAN ANNUAL PRECIPITATION: INCHES


Fig. 3-6. The effect of precipitation on the frequency distribution of the clay minerals in the surface layer
of residual soils from acid and basic igneous rocks in California. From Barshad, 1966. Copyright 1966
Israel Prog. Sci. Translations.

the surface soils. In the dry soils the water remains in the upper portion and all the
products of dissolution tend to accumulate in the thin soil profile. In the soils from
areas of high precipitation water can permeate the soil to a great depth, and a
number of the products of dissolution are removed from the soil by the percolating
water. The end products are chemically simple kaolin minerals and oxides.
118

As a generality the intensity of chemical weathering, the alteration of inherited


physils, and the formation of new physils tends to decrease downward in a soil
profile. However, because of translocation, ascending solutions, and changing
climatic conditions, the expected trend is not always present.
In regions of moderate precipitation where illite or vermiculite are at a maximum
in the surface soil (acid igneous rocks), both decrease with depth. The decrease in
vermiculite is accompanied by an increase in illite and I/V. The decrease in illite is
accompanied by an increase in the kaolinite-halloysite content. This latter trend,
which is not the normal weathering trend, and the presence of some kaolinite-hal-
loysite in soils from the areas of low precipitation, suggest climatic conditions have
varied during the long period of soil development and that kaolinite-halloysite was
formed during an earlier period when the climate was wetter and/or the young soil
had a high permeability. Since the kaolin minerals are more resistant to weathering
than the other physils, they would tend to be preserved.
In Barshad’s California soil study, kaolin minerals increased from south to north
as the amount of rainfall increased. In the soils of the eastern United States the
kaolin minerals generally increase from north to south. Vermiculite, interstratified
physils, smectite, illite, and secondary chlorites are commonly dominant in the
northern soils. As the amount of rainfall ( - 40 inches) is similar throughout the
eastern region, the southward increase in kaolin is presumably due to the increase in
temperature, which causes an increase in the rate of chemical activity.
Several studies of the weathering of basaltic rock of the Hawaiian Islands show
the physil and oxide distribution as a function of rainfall is similar to that in Fig.
3-7 (Sherman, 1952). Montmorillonite is the predominant physil formed when the
rainfall is below about 40 inches; between 40 and 80 inches kaolinite is predomi-
nant; and above 80 inches Fe and A1 oxide hydrates are dominant. The process is
primarily one of desilication. The trend is similar to that in Fig. 3-6 and in many
other studies, but the absolute precipitation values at which the various minerals
form are quite variable. This reflects the importance of other factors such as
temperature, topography, distribution of rainfall, etc.
Van der Merwe and Weber (1963) studied a series of South African soil profiles
developed from granite under different climatic conditions. Summer temperatures
are generally high and those of the winters are medium. In the arid and semi-arid
regions mature soils, derived from granite under a rainfall of less than 50 cm (20
inches) - commonly less than 15 cm (6 inches) - in the summer rainfall region
(evaporation at a maximum), have illite as the principal physil ( < 2 pm fraction).
The 10 A peaks are relatively broad and spacings range from 10.1 to 10.3 A
(untreated). Most of the “illite” appears to be 1/S (or possibly I/Ch) with 5 to 20%
smectite layers. Minor kaolinite and smectite are present in some samples. One
profile shows a gradation with depth. The surface horizon contains a 10.1 A physil.
The underlying horizon contains a 10.3 A physil, and the lowest horizon contains
smectite and I/S with a relatively high smectite content. At least in this profile it
appears a smectite phase developed first and was converted to I/S, with an upward
increase in illite layers. This is probably related to the upward migration of K-rich
water during periods of evaporation. p H values range from slightly acid to slightly
basic.
119

-z
t
2 60
u (HYDRATES)
GIBBSITE
IA
0 DIASPORE
BOEHMITE
LIMONITE
I BAUXITE

ANNUAL RAINFALL IN INCHES


Fig. 3-7. Progressive types of clay development in Hawaiian soils under a continuously wet climate. From
Shermann, 1952b. Copyright 1969 Elsevier Pub. Co.

Where the summer rainfall is 90 cm (35 inches) and higher, kaolinite is the only
physil at all levels of the profiles; gibbsite, goethte, and hemate are relatively
abundant. pH values range from 5 to 6. The crystallinity (peak sharpness) increases
downward, indicating the kaolinite is being destroyed in the upper horizons.
Soils under a summer rainfall of 50 to 80 cm (31 inches) have an intermediate
composition with poorly crystallized kaolinite increasing at the expense of I/S illite.
The authors noted that under low rainfall conditions there was slightly more
chemical weathering when the rain occurred in the winter (less evaporation) than
when it occurred in the summer. However, in the high rainfall regions, summer
rainfalls produced more chemical weathering, more nearly pure kaolinite physil
suite, than winter rainfalls.
The relation of the type of physil formed from granitic rocks to the amount of
rainfall in South Africa differs somewhat from that in California. The main
difference occurs in the arid environments where “illite” is the dominant physil in
South Africa and montmorillonite in California. Some factor other than rainfall
presumably accounts for the difference, i.e., periodicity of rainfall, topography, etc.
As noted, regional variations in soil physil suites can be related to rainfall. The
same type of associations can be developed in a small area, actually in a hand
specimen. Microenvironments or microclimates can produce seemingly anomalous
chemical changes.
Grenada is a small volcanic island in the Caribbean composed predominantly of
basic volcanic lavas and tuffs. Over a distance of 5 km, a flat lying coastal area gives
way to mountainous areas where peaks rise to a height of 840 m. Rainfall ranges
from over 160 inches (410 cm) in the steep mountain areas to less than 50 inches
(128 cm) in the coastal areas. Analyses of the soils indicate that where the rainfall is
120

greater than 80 inches (205 cm) only kaolinite, along with gibbsite and Fe oxides, is
present; where the rainfall is between 60 inches (154 cm) and 80 inches both
kaolinite and smectite are present; where the rainfall is less than 60 inches smectite
is the only physil (Beaver and Dumbleton, 1966).
The relation between rainfall and physil type on Granada is similar to that which
occurs in other areas though the boundary values are different. On Grenada
smectite forms at the relatively high rainfall value of 60 inches and higher. This
illustrates the importance of topography or relief. In the low relief coastal areas of
Granada drainage is restricted, leaching is inhibited (water-logged), and the rela-
tively soluble cations remain in the soil, combining to form smectite.
Mohr and Van Baren (1954) described several similar microenvironmental situa-
tions in the tropic soils of Java. Though lateritic soils and kaolinite are the
predominant weathering products formed from volcanic material, where drainage is
restricted by low relief and/or fine grain size (volcanic ash) smectite forms.
An example of chemical weathering in a temperate climate is provided in a short
review by Paquet and Millot (1972) of weathering in the Mediterranean region. The
Mediterranean climate is a warm temperate climate with dry summers. In parts of
Morocco, Lebanon, Greece and southern France, where the mean annual rainfall is
higher than 80 cm and the podzoilic soils are well-drained, illites and chlorites are
hydrolyzed, first forming mixed-layer Ch/V and I/V and eventually vermiculite
(degradation). As usual, chlorite weathers more easily than illite. In Israel, under
similar conditions, inherited montmorillonite is partially altered to kaolinite (Yaalon
et al., 1966).
When the drainage and rainfall is less (average - 50 cm/year) but still signifi-
cant, illites are converted to I/S. The formation of smectite layers rather than
vermiculite is believed to be due to fixation of Si by the smectite layers. The Si
concentration in the pore waters is relatively high. The process is considered to be
one of degradation and aggradation.
In the semi-arid or arid lowlands (rainfall < 30 cm/year) solutions brought in
during the humid season are concentrated during the dry season. Ions enter the
open illite (illite + I/S) structure and the illite becomes better crystallized. Presuma-
bly, some of the smectite layers contract by fixation of K (aggradation).
In the poorly drained lowlands where the rainfall ranges from 30 to 50 cm and
the seasons are alternately wet and dry, the ions carried in from the highlands
during the wet season are concentrated in the dry season to the extent montmoril-
lonite precipitates (neoformation). Under similar, but more arid conditions (rainfall
< 30 cm), thick caliche develops and palygorskite, commonly 100% of the clay
fraction, is neoformed. Palygorskite alters to montmorillonite when rainfall exceeds
30 cm.
Neoformed montmorillonite and palygorskite are to some extent characteristic of
alternating wet and dry climates.
Though chemical weathering is generally considered to be minimal in cold
climates, Reynolds (1971) has shown that it can be relatively intense in Alpine
environments (temperate, glacial). His study was conducted in the vicinity of the
South Cascade Glacier in northwestern Washington at elevations between ap-
121

proximately 1800 and 2500 m. Slopes range between 30 and 90‘. Perennial snow and
ice cover about 50% of the watershed. Analyses of the lithosols (mass of imperfectly
weathered rock fragments; found primarily on steeply sloping land) showed the
phlogopite in the phlogopite schists was altered to vermiculite and, to a lesser
extent, mixed-layer mica-vermiculite and trioctahedral smectite. The metadiorite
was altered to lithosols rich in gibbsite, up to 46% Al,O,. No kaolinite was detected.
Total annual precipitation averages - 400 cm, mostly in the form of snow;
however, during the spring and summer the snows melt and release an equivalent
amount of water. The water is highly charged with chemically corrosive CO,. The
pH of the waters in gibbsite lithosols (bauxite) ranges between 6.5 and 7.2. This is
the optimum pH for bauxite formation, in contrast to kaolinite, because at these
hydrogen-ion concentrations the solubility ratio of Al(OH),/SiO, is at a minimum.
Further, though air temperatures are low, rock surface temperatures are as high as
40°C. Mechanical weathering and erosion continually produce fresh rock surfaces
and increase the surface area. Thus, there is abundant runoff, an effective pH, a
reasonably warm temperature, and a continuous supply of exposed fresh rock - all
the ingredients necessary for chemical weathering.
Based on the K concentration of runoff waters, Reynolds calculated that ap-
proximately 100 to 200 T/km2/yr of vermiculite (clay to sand size) was formed
from biotite. This is much higher than the average rate of physil production for
North America. One reason for the high weathering rate is the abundance of easily
weathered phlogopite. Nevertheless, the study demonstrates chemical weathering
can be effective in alpine environments. The physils are easily eroded and extensive
soil profiles are not developed. The situation in the South Cascade Glacier area is
analogous to that in the Amazon River drainage basin, where the vast majority of
the physils and dissolved solids are derived from the Andes Mountains.
As shown by Reynolds’ (1971) study of the conversion of biotite to vermiculite,
Jackson’s (1964) study of the great soil groups (Table 3-1), and many other studies,
the conversion of bedrock micas and chlorites to expandable, mixed-layer, and
intergrade (soil chlorites) physils is characteristic of chemical weathering processes
in temperate environments. The studies discussed in the preceding pages largely
describe the conversion of granites (feldspars) and volcanics (glass and feldspar) to
soil physils. Most of these physils are neoformed. Other bedrocks, such as shale,
slate, schst, loess, till, etc. contain an abundance of physils, and the soils formed on
these rocks can contain inherited, modified, and neoformed physils. In arid and cold
climates these bedrock physils are relatively unmodified; in humid climates they are
commonly completely altered to kaolin and sesquioxides. It is in temperate climates
where these “intermediate” physils are usually best developed. Chemical weathering
is active enough to leach ions from the physils but does not completely destroy the
character of the primary layer silicate.
122

CHEMICAL WEATHERING PROCESSES

Inorganic

Introduction
Fresh rocks and minerals are seldom in equilibrium with near-surface waters,
temperatures, and pressures. In the presence of water these rocks and minerals are
chemically modified towards a phase that is more stable in the near-surface
environment. Commonly, the product is a physil. During weathering much of the
newly-formed or altered physils are eroded before an equilibrium mineral assemb-
lage is developed. We are concerned primarily with the processes that lead to the
modification (transformation) or formation (neoformation) of physils. These include
hydrolysis, dissolution, hydration, oxidation, chelation, and the activities of primi-
tive plants (fungi, lichen, and moss).
Rainwater pH values range from about 3.0 to 9.8 and are generally slightly acid.
Pure H,O equilibrated with atmospheric CO, should have a p H of 5.7 and contain 1
ppm of dissolved CO,. pH values lower than 5.7 are usually due to industrial
contamination; higher values are due to the presence of sea aerosol. The chemical
weathering effectiveness of rain is largely due to the H + ion, which is present in
amounts of to moles per liter.
The acidity of rainwater should increase after it has entered the soil; however,
reaction with minerals tends to cause an increase in pH. The upper part of the soil
has a very large bacterial population. Bacterial activity causes the organic material
to be oxidized, releasing CO,. Soil gases above the zone of water saturation
commonly have CO, contents 10 to 40 times that of the free atmosphere. The CO,
+
reacts with H,O to form carbonic acid and H + ions: H,O CO, S H,CO, %
HCO, + H + 4 2Hf + C O : - . In the humid tropics bacterial oxidation is so rapid
that little organic material persists to move down into the soil. As oxidation occurs
at the surface most of the CO, is released to the atmosphere rather than to the soil
water (Garrels and Mackenzie, 1971).
Respiration by plant roots can produce very low pH ( - 2) waters adjacent to the
root. One rye plant was found to have 385 miles (2550 square feet of surface) of root
and 6600 miles (4320 square feet of surface) of root hairs (Foth and Turk, 1972).
The contribution to soil acidity can be considerable.
Other sources of H + are organic acids, exchangeable H, and the hydrolysis of
AP+.
Al’+ + H,O S Al(OH)’++ H + (very acid conditions)
or
A13++ 3H,O S Al(OH)3 + 3H+ (less acid conditions)

Experimental
The major sources of clay physils are the aluminum-silicate minerals and rocks. A
freshly fractured feldspar or other AlSi mineral has a surface dominated by oxygen
123

Fig. 3-8. Diagrammatic representation of the experimental feldspar weathering model of Wollast (1967).
From Berner, 1971, Principles of Chemical Sedimentology, McGraw-Hill.

ions. Si, Al, Fe, Mg, Ca, K, and Na cations normally occur beneath the surface
oxygens, but some will be relatively accessible to surface water. The fractured
surface is disordered and higher in energy than the interior of the crystal (Keller,
1978). When the surface comes in contact with H 2 0 the H + ions will displace the
alkali and alkaline earth cations (hydrolysis) and, for feldspars, form a thin film of
“acid feldspars” (Fig.3-8). (Keller, 1978; Wollast, 1967; Garrels and Howard, 1959).
The initial reaction is simply:
Hdq + K sparsurfxe --* H sparurfxe + Kzq
The rate of this reaction is a function of pH. At pH values above 5, dissolved A1
precipitates as Al(OH), (Wollast, 1967). The dissolved silica (H4Si04)concentra-
tion builds up to a certain level and then reacts with the Al(OH), to form kaolinite.
When a continued supply of fresh water is available, the silica and alkali ions will be
flushed from the system and the relatively insoluble A1 will remain as Al(OH),.
When the flow rate is low and the dissolved ions remain in the vicinity of the parent
mineral, montmorillonite (or illite) forms (Berner, 1971). The latter assumes Mg2+
and K + are present, derived from another mineral. These reactions can be repre-
sented as:
albite gibbsite
+
H + NaAlSi,O, + 7H,O --* Al(OH), + N a + + 3H4Si04,, (high flow rate)

2 H + + 2NaAlSi,O, + 9 H 2 0 -+

kaolini te
A12Si2(OH), + 2Na+ + 4H4SiOOdq (moderate flow rate)

mon tmorilloni te
M g 2 + + 3NaA1Si,08 + 4 H 2 0 --* 2NaO0,+ All 5Mgo,Si4010(OH)2+ 2 Na + + H,SiO,,,
(stagnant)
124

sdutuvl

H H
'
0
\ti 0; H\
H 0
HS w
H
H
30 - K* x
6 HO
K"

s',
OH
OH- H HO OH
H OH'
:H
0

Fig. 3-9. Scheme of an orthoclase surface reacting with water at various p H s . Conditions in acid soils
might approximate that depicted for pH less than 5, those for near-neutral soils by that for pH 6, and
those for quite basic soils by that for pH near 9. From Birkeland, 1974, Pedology, Weathering, and
Geomorphoplogical Research. Copyright 1974 Oxford University Press.

Hydrolysis of aluminum silicates increases the pH of the neutral water to 8 to 11


(Keller, 1962). If this water is not exchanged or additional H + or carbonic acid
added, the environmental situation is one in which, depending on the ions present,
smectite, illite, chlorite, sepiolite, or palygorskite can crystallize. Fig.3-9 summarizes
125

the reactions occurring at the interface of K-feldspar and water with various pH
values.
Over the pH range of most natural water silica is more soluble than A1 or Fe3+
and chemical weathering is sometimes referred to as desilication. However, the
presence of weak organic acids such as humic and fulvic can reverse the process.
Various organic materials can complex A1 and Fe and effectively remove them from
the system. Numerous recent studies suggest organic acids and organisms may play
a major role in chemical weathering.
Buxton (1968) proposed a weathering index based on the SiO,/Al,O, ratio of the
bulk soil. He found that fresh Si-A1 rocks commonly have values - 4.0-4.5; with
increasing chemical weathering, in humid areas, the ratio systematically decreased to
2.0 (kaolinite).
Various studies (Johnson et af.,1968; Nesbitt et af., 1980; Nesbitt and Young,
1984) of igneous rocks indicate Na and Ca (from plagioclase) removal rates are
greater than the removal rate of K (K-feldspar and mica). Thus, some of the K and
Mg is more likely to be incorporated in pedogenic physils than are Na and Ca. With
continued weathering all four elements will eventually be removed from the soil.
Some Si is removed during these preliminary stages of weathering and, ideally,
continues to be removed after most or all of the Na, Ca, K, and Mg are gone.
Fig. 3-10 shows the solubility of Si and A1 as a function of pH. At pH values less
than 9, dissolved Si is present as monomeric silicic acid, H4Si04 (Si surrounded by
four OH’S). At values above 9, H4Si04 dissociates to H,SiO; and H2Si02-. As
most natural waters, other than some alkaline lakes, have pH values lower than 9,
H4Si04 is the complex commonly involved in the neoformation and diagenesis of
physils. The solubility of Si in distilled water in contact with quartz at room
temperature is approximately 10 ppm (Krauskopf, 1967); for amorphous silica the
value is 108 ppm (Greenberg and Price, 1957).
The species of A1 in aqueous solution, calculated from stability data, range from
A13+at pH 3 and lower, through Al,(OH);+ at a maximum concentration at pH
4.5, Al(OH)’+ at a maximum at pH 4.7, Al(0H): at a maximum at pH 6.0, to
Al(0H); at pH 8.5 and higher. As the OH to A1 ratio increases above 1.0, the
monomeric hydrated A13+, in six-fold coordination, polymerizes. The polymerized
Al-hydroxide eventually forms gibbsite, at a pH of 6.7 (Huang and Keller, 1972).
Concentrations of dissolved A1 range from about 100 ppm A13+ in higher acid
ground waters to about 20 ppb at intermediate pH values. High concentrations exist
as Al-OH complexes at high pH (Keller, 1978). Concentrations are higher in the
presence of complexing organic acids. Although A1 is inherently mobile, the
availability of the precipitating anions - OH-, PO:-, and Si0:- - causes A1 to be
relatively immobile.
Depending on the proper pH conditions, when both Si and Al, or Mg, are present
in solutions they tend to coprecipitate. Under neutral to mildly acid conditions
allophane or kaolin usually forms. At higher pH values the presence of other
cations, such as K, Na, and Mg, allows the more complex physils to crystallize.
The mechanism by which physils crystallize from solution is not well known. One
concept is that colloidal SiO, and Al(OH), are mutually attracted and coprecipi-
126

' 0
I I I I 1 I
0 I 2 3 4 6 8 7 8 s 10 I1
PH

A1203 dato f r o m C o r r r n s
Slop d a t a f r o m Krauskopf
Fig. 3-10. Solubility of A1,0, and amorphous Si0, in water. From Keller (1963). Reprinted with
permission from Clays Clay Miner. Proc. 12th Conf., 129-152. Copyright 1963 Pergamon Journals, Ltd.

tate. The surface charge of SiO, is negative at any pH above 2, owing to the loss of
surface protons. Conversely, all simple oxides have positive surface charges in acid
solutions and negative charges in alkaline solutions. Thus, at pH values higher than
the isoelectric point (pH where charge is zero) both the silica and other oxides and
hydroxides have a negative charge and repulse each other. Electrostatic attraction
can only occur at pH values lower than the isoelectric point. The isoelectric points
of various A1 oxides and hydroxides occur at pH 7.7 or lower. This explains why
kaolins form from acid to neutral solutions (Siffert, 1978).
In contrast, montmorillonite is formed above pH 7.5. In order for this to happen
the aluminous material must contain some Mg in order to maintain a positive
charge at p H values above 7.5. Brucite has a positive charge that is not influenced
by pH. The coprecipitation of Mg with A1 hydroxide allows the complex to keep a
positive charge at relatively high pH values and to combine with the negative
charged silica. Laboratory experiments indicate Mg is necessary for the synthesis of
smectite, illite (I/S), and chlorite (Harder, 1972, 1977; Siffert, 1978). The laboratory
studies indicate crystalline material forms best from very dilute solutions under-
saturated with respect to amorphous silica. At higher SiO, concentrations the
precipitates remain amorphous.
127

2 OH-Si-OH
I
I
- I
OH-Si-0-Si-OH
I
I
I
+H,O

In the presence of appropriate metal cations %(OH), polymerizes to form a


tetrahedral sheet.
The formation of the octahedral sheet is related to the formation of hydroxides.
Cations in solution attract the negative end of water molecules and become
hydrated [Al(H,0)6]3f, [Mg(H20)6]2'. As the pH of the solution and hydroxyl ion
concentration increases, the hydroxyls progressively replace the molecules of water
present in the layer of hydration:

[Al(H20)6]3++OH + [Al(H20)5(OH)]2++H,O

... + [A1(H,0),(OH)3] + H,O


Due to the small size of A13+ and the high charge density it is difficult for the
hydroxyl ions to replace the inner three water molecules. Thus, the initial precipitate
of the neutralized A1 is not the dehydrated hydroxide. For larger cations, such as Zn
and Mg, the water molecules are more weakly bonded and can be completely
replaced by hydroxyls relatively easily to form the hydroxide precipitates; however,
a stable, complex cation [Zn(H,O),OH]+ forms before (lower pH) the precipitation
of Zn(OH),. Further, when Si is present it always coprecipitates with the hydroxide.
The process is similar to that involved in the formation of basic salts.
Basic salts, such as oxychlorides, (RZ+OH)Cl,have a structure similar to Mg
hydroxide (brucite). In brucite a plane of hydroxyl ions occurs on each side of a
plane of Mgz+ ions, so that the hydroxyls have an octahedral arrangement around
each Mg2+ ion. A triad of hydroxyl ions occurs above each Mg2+. In oxychlorides,
where half the OH- ions are replaced by C1- ions, the structure consists of a plane
of metal ions with a parallel plane of OH ions on one side and a plane of C1 ions on
the other. Physils are essentially basic salts with silica acting as an anion (Siffert,
1967). Note that the complex ions (two types of anions) necessary for the formation
of basic salts exist only in the pH range between the formation of hydrated ions and
the precipitation of the hydroxide.
Siffert suggests that Si(OH), can combine with Al-hydroxy ions to form the two
128

types of monomers shown in the following equations:

Ho\
HOFSi-OH
HO
+ [AI(OH)]’+ - + H+

followed by

HOo ~\ S
/H , I-O--A~-OH~+ + HO-SSiTOH
/
OH

1 HO 1 OH

- Ho\ .
HO7Si-O-Al-O-Si\OH
HO I
OH
/
OH

OH
+ Hi

or

Ho,
HO/SI-OH
HO
,
+ [AI(OH),]+ - Ho, ,
HO/Sl-O-Al,
HO
/
OH
OH
+ H’

Depending on the Si concentration and pH, either a 2:l (high pH) or 1:l (low pH)
type of “monomer” can precipitate. These elements apparently polymerize to form
sheet silicate structures. The H + ions in the hydroxyls attached to the Si ions are
removed as H,O:

H,O - Ho\,
HOFSi-0
HO
As the water of hydration is relatively weakly bonded to Mg2+, Fe2+,and Fe”,
i t is relatively easy to form basic salts of these cations. Removing the water of
hydration surrounding the Al’+ ion is more difficult, and Siffert found it necessary
to form oxalate ions of Al’+, [Al(C204)3]’-,in order to grow kaolinite. The oxalate
also imparts a coordination number of 6 to Al. A number of A1 complexing acids,
including fulvic, have been used to synthesize kaolinite at low temperatures (Siffert,
1978; Iglesia Fernandez and Martin Vivaldi, 1978). It is likely that organic acids
play the same role in nature.
Another technique used in the study of chemical weathering is the construction
of stability diagrams (Fig. 3-11). These diagrams are calculated from data on free
energies of formation and the composition of the solid phases, neither of which is
very accurately known. However, the diagrams indicate the approximate solution
compositions in which various minerals are most stable and indicate what minerals
should precipitate from solutions of varying compositions. Because the concentra-
tion of dissolved A1 is usually very small, it is assumed to be retained entirely in
solid phases (inert). The vertical dashed lines are solubility lines for quartz ( - 4) and
amorphous silica. Only H4Si0, concentrations that plot on the quartz line are in
equilibrium with quartz. Solution compositions that plot to the right of the quartz
line are supersaturated with respect to quartz; compositions that plot to the left are
129

K-feldspar

I
I
I
I
I
I
I Kaolinite
I
I
I
I
I

-5
,-4
I

LW a u 4 s b o 4

ao ' ' A ILLITE,GIBBSITE.OR BOTH


' '
C MIXED-LAYER, KAOLINITE.OR BOTH
D M0NTMORILLONITE.KAOLINITE.W BOTH
-
X ILLITE.MICROUINE.OR BOTH
1x1 Y YIXEDLAYER.MlCROCLINE,OR BOTH -
Z MONTMORILLONITE.MICROCLINE, OR
-

LOG %i02(aq)

Fig. 3-11. Stability relationships among some minerals in the system K20-A120,-Si0,-H,0 at 25OC.
Upper. From Drever, 1982, The Geochemistry of Natural Waters. Reprinted by permission of Prentice-
Hall. Lower. From Aagaard and Helgeson, 1983. Copyright 1983 The Clay Miner. SOC.

undersaturated. Quartz is kinetically unreactive at low temperatures and the solubil-


ity of amorphous silica is generally the upper limit for H,SiO, activities in natural
water (Drever, 1982).
130

Because the stability diagrams can take into account only a limited number of
cations, they must be used with caution, For example, Fig.3-11 shows only minerals
in the system K,O - AI,O, - SiO, - H,O. If a solution contained Na in addition to
the other ions, then a different group of minerals might be stable (Drever, 1982).
However, the diagrams do show the approximate stability fields for minerals
precipitated from solutions of varying composition. They also indicate the sequence
in which minerals would crystallize as the water chemistry changes. The diagrams
indicate which minerals can coexist in equilibrium. Minerals adjacent to one another
in the diagram (kaolinite and gibbsite) can be in equilibrium, but those which do
not have common boundaries (gibbsite and K-feldspar) cannot.
The stability diagrams can be used to provide information on the origin of
physils. When soil water-chemistry data are plotted on a diagram, they indicate
which mineral is in equilibrium with the solution. If x-ray analyses indicate the same
mineral is relatively abundant in the soil, it can be concluded the mineral probably
was neoformed in the water in which it now occurs. If the minerals in the soil are
different from those indicated from the water data plot, it is reasonable to conclude
the minerals are inherited. In the latter case, with time, the waters will eventually
change composition so they are in equilibrium with the soil minerals or, if the
composition of the water is controlled by a “remote” source, it is likely the minerals
will eventually alter to an assemblage in equilibrium with the water.
Waters from most rock types plot in the stability field of kaolinite (Na,O -
A1,0, - SiO, - H,O system (Bricker and Garrels, 1965). Either this indicates that
all soil physils will eventually be kaolinite or they chose the wrong chemical
parameters.

Natural
SEM and TEM studies of weathered feldspars indicate the formation of physils is
a more complex process than that suggested by laboratory studies. Various studies
of weathered feldspar crystals showed that crystallographic controlled solution pits
form on the surface of feldspar; however, Eggleton and Buseck (1980) studied
microcline from a humid, temperate climate (New South Wales, Australia) and
observed that during the early stages of weathering physils form in small vacuoles
within the body of the feldspar. It is these vacuoles that give weathered feldspar a
turbid appearance in t h n sections. The vacuoles tend to develop in high strain
regions (boundary between lattices of different geometry). Water diffuses along the
structural defects, gradually expanding withn the crystal. This water is essentially
static and is in equilibrium with the feldspar and physils, either illite or I/S.
In some vacuoles they were able to establish that the primary precipitate was an
A
amorphous material with a ring texture (about 250 in diameter). A progression
was observed from very unstable, almost structureless rings, to more stable arcuate
A
laths with 10 lattice planes, to a well-crystallized 10 A physil. Many of the physil
flakes were only one or two layers thick.
A different sequence of events was deduced from a study of weathered feldspars
from a humid, tropical environment (southern Brazil) (Tazaki and Fyfe, 1986). In
this instance the highly weathered K-feldspar has been altered to halloysite (7 A)
131

and gibbsite. The slightly weathered K-feldspar contains a variety of “primitive clay
precursors” on the surface of the feldspar. Auger depth profiles indicate that Si and
Fe (and presumably H,O) increase and A1 and K decrease from the interior to the
surface of the feldspar. The outer primitive clay zone is about 300 A thick. The
suggested sequence is: The precipitation of hydrated Fe-oxide on the unaltered
feldspar surface; growth of mohair-like fibers which curl into circular forms
(150-200 A diameter) which have some structural order; development of capsules
and squat cylinders; finally, tubular halloysite (7 A). The fibers, circles, and
cylinders have some structural order, but it is not well defined and is called
primitive clay. The primitive clay is composed largely of Si and Fe, with little or no
Al. Appreciable A1 was not detected until the stage of crystalline halloysite forma-
tion. No evidence was found to indicate the mechanism by which A1 eventually
replaced the Fe.
It is interesting to note that amorphous material (11-12%) in mildly weathered
soils from Quebec, Canada, is composed primarily of mixed silica-iron hydroxides.
The ratio of SiO, to Fe,O, ranged from 1.9 to 2.5 (McKyes et al., 1974). It was
proposed that silica was chemisorbed on ferric hydroxide.
On the other hand, a study of weathered feldspars from a saprolite zone (bauxite
laterite profiles) developed from a granite (45 km southeast of Perth, western
Australia) showed the feldspars altered to halloysite, kaolinite, and gibbsite with no
evidence of a noncrystalline precursor (Anand et al., 1985). Eswaran and Bin (1978)
found that, in a soil developed on a Malaysia granite, feldspars altered to kaolinite
or gibbsite directly without any intermediate phase; however, when feldspar altered
to halloysite there was an intermediate phase consisting of globules of amorphous
gel. They also observed that in the same horizon one feldspar crystal may alter to
kaolinite while another may alter to gibbsite, depending on the absence or presence
of a void adjacent to the feldspar.
Eggleton and Boland (1982) describe the conversion, during weathering (Wara-
tah, Tasmania) of orthopyroxene to talc and smectite by topotactic transformation
in the solid state with no intervening non crystalline phase. TEM pictures of
iron-beam thinned material show that initially the orthopyroxene alters to a 9 A
layer silicate, structurally coherent with the parent material (Fig. 3-12). In this
process the (100) of the pyroxene (17.27 A) becomes the (001) of the 9 A layer
silicate (talc). The I-beam chains of the pyroxene shift b/4 parallel to the original
(100) plane (Fig. 3-12). Columns of vacancies develop where strips of various chain
widths terminate (A in Fig. 3-12). These channels are believed to be sites of high
speed diffusion. Growth of talc can be aclueved by exchanging the vacancy for the
adjacent 3-chain strip of I-beams. There is no change in the coordination of the
tetrahedral and octahedral cations, no noncrystalline intermediate phase, and
minimum bond-breakage. Transformation is entirely in the solid state, merely
requiring the release of Mg and the incorporation of H (to produce OH). During
more advanced stages of weathering, the orthopyroxene and/or talc alter to
smectite and minor chlorite.
Numerous studies have described the procedures by which micas are altered to
kaolinite. Typical mica weathering sequences in the Cenozoic weathering crust,
132

Fig. 3-12. Upper: Layer silicate (9 A) and alteration of orthopyroxene. The enlarged inset shows two
channels formed at the termination of 3-chain ‘zippers’. Lower: Diagram showing the I-beam structure of
part of the region shown in the upper figure. The chains shift to form sheets. From Eggleton and Boland,
1982. Copyright 1982 The Clay Miner. SOC.Courtesy T. Eggleton.

developed on granite and gneisses in Lower Silesia, have been described by Stoch
and Sikora (1976). In areas where the K+/H+ ratio is low, biotite and muscovite,
and presumably secondary illite, can pass directly into kaolinite. Mica books with
interlayers of kaolinite are common in kaolinite deposits. As K + is removed from
the interlayer position, tetrahedral A1 apparently moves from the tetrahedral sheet
133

to the interlayer position to form a new octahedral layer. In the initial stages water
molecules may participate in the octahedral coordination. Fe and Mg are removed
from the original octahedral sheet and replaced by A1 obtained by the solution of
feldspars. Eventually every other octahedral sheet inverts to form a 1:l structure.
Inversion is probably facilitated by the decrease in isomorphous substitution and
layer charge.
Where weathering is more gradual, the sequence for muscovite appears to be:
muscovite + M/S (or M/V) + smectite + kaolinite

The gradual alteration of biotite is more complex:


biotite + Fe dioctahedral mica + illite + kaolinite
The biotite and dioctahedral micas have the following composition:

( Fe?.i2 Fe,2,:9Mg0.79) (si2.87A11 .11)OIO (0H)2K0.86Na0.05


+

-71Fei,&2Mg0.15) (si3.46A10.54 )OlO (oH)2K0.58Na 0.03

It is of interest that a trioctahedral mica can convert to an A1 dioctahedral mica.


This type of alteration apparently occurs where the concentrations of K + and A13+
are high, such as in the vicinity of K-feldspars. Thus, K remains in the interlayer
position, and the relatively mobile Fe2+and Mg2+are removed from the octahedral
layer and are replaced by the relatively immobile A13+. The Fe2+ oxidizes to Fe3+
and precipitates as Fe oxides. The Mg is apparently removed from the system and
becomes available, downstream, for the formation of Mg physils.
Apparently no one mechanism is going to explain the procedures by which
physils form from A1 silicate minerals. In this regard the role of microenvironments
in determining the nature and distribution of authigenic physils in soils should be
mentioned. In their study of weathered feldspars Anant et al. (1985) observed small
scale variations in the physil suite (kaolinite, halloysite, and gibbsite). Tardy et al.
(1973), in their studies of granite, found that plagioclase altered to montmorillonite
and vermiculite in narrow fissures, but kaolinite formed in well-drained pores. The
same situation often prevails on a megascale.
In a study of the initial stages of weathering of a granite (France), Meunier and
Velde (1979) found the secondary physils were related to mineral composition and
mineral position rather than depth in the profile. Illite formed at the boundaries
between muscovite and orthoclase. Micas and orthoclase altered to vermiculite and
beidellite. The fractures, where water movement was more rapid, contain kaolinite
and Fe-oxide. With time everything alters to kaolinite and oxides.
These microenvironmental studies indicate that even in climates where weather-
ing is intense 2:l physils can form under “protected” conditions, where water
movement is restricted. With time the conduits for moving water prevail and, under
most conditions, most primary and secondary minerals are converted to kaolin and
Fe and A1 oxides-hydroxides. If erosion is relatively rapid 2:l minerals could be
produced and eroded for long periods of time.
134

Oxidation
Oxidation involves the loss of electrons from an ion, causing an increase in the
positive charge. Oxidation of F e z + to Fe3+ is the most important inorganic
oxidation process that occurs during weathering. The oxidation of Fe in a mineral
increases the positive charge and forces other cations to leave the crystal lattice so
that neutrality can be maintained. This disrupts the lattice and makes the mineral
more readily subject to faulting and chemical solution. The role of the Fe oxidation
process during the conversion of biotite to vermiculite has been studied in consider-
able detail, and it appears that Fe oxidation is not a major factor in the vermiculiti-
zation process.

Organic

Water soluble anions of organic acids of both h g h and low molecular weight are
important complexing agents and are abundant in most soils. Oxalate, acetate, and
formate are commonly present in minor amounts and most of the complexing is
done by the more abundant humic and fulvic acids. These complexing or chelating
agents are formed primarily by biological processes and excreted by lichens.
Chelation basically involves the exchange of H + in the organic molecule by other
ions, commonly A1 and Fe. The chelate is stable at the p H condition in which the
complexing occurs. The chelates may move to B horizon where changing pH
conditions cause them to precipitate or to be adsorbed on physils. Or the chelates
can be flushed from the soil, eventually finding their way to rivers and the ocean.
The vast majority of Fe in many river waters is complexed with fulvic acid rather
than being in true solution.
Huang and Keller (1970) determined the solubilities of five aluminum silicates in
H 20,C0,-charged water, and organic acids (acetic, aspartic, salicylic, and tartaric).
The organic acids are representative of components of humic acid and lignitic acids.
Si, Al, Fe, Ca, and Mg were more soluble in organic acids than in the other two
media. On a relative basis A1 was more soluble than Si. Periodic sampling of water
percolating through unconsolidated volcanic ash indicated that the presence of
humic acids increased the concentration of A1 and Fe by several orders of magni-
tude over what it would be in the absence of organic complexing material. Na, K,
Ca, and Mg values also correlated with the dissolved organic carbon concentration
(Antweiler and Drever, 1983). Though A1 is commonly considered to be relatively
insoluble and remains behind in the residue, it is evident that it can be mobilized
when organic acids are present in the water. Also, Fe in solution tends to form
oxy-hydroxyl compounds and settle from the solution; however, organically com-
plexed Fe can be transported for great distances.
Fast dissolution rates were observed when a 0.025% solution of fulvic acid flowed
through columns filled with biotite and chlorite (Fig. 3-13) (Kodama et al., 1983).
All cations were dissolved, but there was a preferential dissolution of Fe and Mg
from the octahedral sheet. The biotite was considerably more soluble than the
chlorite. Extrapolation to natural soil (Podzolic) conditions indicates biotite would
135

BIOTITE
FLOW RATE . 1 mVh

4-

3-

0 a0 40 lo0
Days
Fig. 3-13. The dissolution by 0.025% FA solution of Si, Fe, Al, Mg, and K from biotite in an open
system; flow rate: 1 mL/h. Amounts of elements dissolved are expressed in milligrams per gram of
mineral. From Kodama et al., (1983). Copyright 1983 Can. Jour. Soil Sci.

be dissolved in 3024 yr and chlorite in 9240 yr, relatively short times. The biotite
apparently went through an expansion phase during dissolution.
Organic acids can apparently play a direct role in the formation of clay physils.
Fulvic acid has been used to promote the formation of kaolinite in solutions
containing Si and Al. Apparently the fulvic acid chelates A13+ ions, maintaining
them in octahedral coordination and facilitating the crystallization of kaolinite
(Linares and Huertas, 1971; La Iglesia Fermandez and Martin Vivaldi, 1971). One
of the problems with forming kaolinite at low temperatures has been how to prevent
the A13' ions from assuming tetrahedral rather than octahedral coordination.
In addition to their role in weathering, soil organic material can be bonded to the
physils, forming physil-organic complexes and commonly producing aggregates.
Most of the organic matter in soils is present as negatively charged polymeric
materials (polyanions). The polyanionic humic and fulvic acids are only weakly
adsorbed by negatively charged physil mineral surfaces. Adsorption of humic and
fulvic acids by physils is commonly due to the formation of cation bridges.
Polyvalent cations act as a bridge between the anionic groups of the polymer and
the mineral surface. The process involves the replacement of a water molecule from
136

the hydration shell of the exchangeable cation by an oxygen of an anionic group of


the organic polymer. Charge neutrality at the surface is maintained by the ion
formerly part of the organic group entering the exchange suite of the physil. Humic
and fulvic material adsorbed only by bridge linkage can easily be displaced by
leaching the soil with the salt of a monovalent metal (Greenland, 1971; Theng,
1979).
At pH values below 8, A1 and Fe hydroxides normally have a positive charge and
can bond organic anions by coulombic attraction. This type of bonding can also
occur when the A1 and Fe hydroxides are adsorbed on the surface of the physils.
Adsorption at the edges of the clay lattices is due to the same mechanism. The
organic anions bonded by coulombic force can be displaced by raising the pH to 8
or 9, which neutralizes the charge on the hydroxides.
Under wet conditions the physil-organic bonds can be broken fairly easily;
however, once the complex has been dried, hydrogen and Van der Waals bonding
develops and it is difficult to desorb the organic material.
Another type of adsorption process is called ligand exchange. In this type of
adsorption the anion penetrates the coordination shell of an Fe or A1 atom in the
surface of the hydroxide. The anion is thus incorporated with the surface hydroxyl
layer. The organic anions are difficult to displace from t h s complex. Ligand
exchange apparently is responsible for the strong adsorption of humified materials
by amorphous Si-A1 complexes (allophane). The adsorption capacities of the various
physils are generally proportional to their surface areas. At comparable pH and
solution concentrations, the adsorption of humic acid decreases in the order:
montmorillonite > vermiculite > kaolinite > chlorite > biotite > muscovite. Organic
matter adsorbed on allophane and physils is more resistant to microbial decomposi-
tion than non-adsorbed organic material.
Physils can apparently adsorb more humic acid when in sea water than in fresh
water (Rashd et al., 1972). This presumably, in part, accounts for the organic-rich
floccules formed in estuaries.
In the natural system some organic materials can also be adsorbed in the
interlayer space of expandable physils. Experiments indicate intercalation only
occurs when the p H of the solution is below 4. At low pH the organic polymers exist
as more or less uncharged species. Under these conditions humic acid is relatively
insoluble, so most studies have been conducted with fulvic acid (for review see
Theng, 1979). Na+-montmorillonite expands to - 18 A when exposed to a low pH
fulvic acid solution, suggesting a double layer of fulvic acid in the interlayer space.
However, the amount of expansion depends on the type of exchangeable cation;
basal spacings range from about 15 A to 19.2 A. Theng (1979) believes the bonding
mechanism is primarily ion-dipole interaction. Several types of bonding may be
operative. For example, at the low pH values ( - 2.5) at which many of the
experiments are conducted, A13+ is likely leached from the physil lattice and
adsorbed on exchange sites as a polyhydroxide where ligand exchange bonding
could occur.
It is often difficult to determine if fulvic acid is present on external physil
surfaces or in the interlayer position, or both. Differential thermal analysis studies
137

indicate the combustion of interlayer fulvic acid occurs at a higher temperature


( - 670°C) than for adsorbed fulvic acid (300" to 550°C). Pure fulvic acid has
exotherms at 330°C (shoulder) and 450°C (Kodama and Schnitzer, 1969).
Tan and McCreery (1975) were unable to get humic acid to move into the
interlayer space of montmorillonite and concluded the humic compounds, with a
molecular weight of 1,000 to 30,000, were too large to fit into the interlayer space.
As interlayer organic complexes only form at pH < 4 and fairly high solute
concentrations, they are not likely to be present in normal soil. Such complexes have
been found in several acid soils.
Though organic acids are effective in altering minerals to kaolin, muds and rocks
with a high content of organic detritus do not have a high content of kaolinite, nor
do most black shales. In the organic-rich ( - 30%) marsh muds of the southeastern
United States, the physils are composed predominantly of montmorillonite or
kaolinite depending on the physil suite of the closest river. There is no evidence that
kaolinite is forming in the marsh muds.
Timofeev and Bogulyubova (1975) studied the physils in Holocene peat bogs in
the Rioni intermontane trough, adjacent to the Black Sea. They observed that with
increasing organic content, up to 80%,the montmorillonite expanded slightly more
when treated with glycerol, chlorite converted to vermiculite via a Ch/V stage, and
in some cores kaolinite increased slightly (in other cores the reverse trend was
observed). Muscovite was generally unaffected; however, it was less abundant and
presumably altered to an expanding physil in reed type bogs. They found that reed
ash of living reeds contained 20 to 27% potassium, as compared to alder ash, whch
contained 6 to 9%. Presumably the strong demand for K by the reeds was
instrumental in stripping K from the mica.
Pliocene to Quaternary sapropels in the eastern Mediterranean contain as much
as 16.7% organic carbon. Physil degradation increases with increasing organic
content. Palygorskite and smectite are the least stable physils, with the latter
apparently forming mixed-layer Ch/S and I/S. With increasing organic content,
illite and chlorite are altered to I/V, I/S, Ch/V, and Ch/S. Kaolinite is least
affected, and though nearly all of the clay-sized physils may be destroyed there is no
evidence of the formation of kaolinite (Sigl et al., 1978).

Microbial Alteration

Microbial life - such as bacteria, algae, fungi, lichen, and moss - play a major
role in the chemical weathering of minerals, particularly fresh rock. Microfauna
rapidly inhabit fresh, exposed rock surfaces and participate in the earliest stages of
soil formation. Varnish encrustations on bare surfaces of glaciated rock have been
found to contain up to one million microorganisms per gram of material.
Large-scale microbial mineral transformation can be brought about by direct
enzymic oxidations or reductions. The role of bacteria in concentrating such
elements as S, Fe, and Mn has been well documented (Silverman and Ehrlich, 1964).
Their role in the formation or degradation of silicate minerals is less well known.
138

One experiment indicated that when the iron-oxidizing bacterium Thiobacillus


ferrooxiduns was added to solutions of ferrous sulfate (pH 3.0) and montmorillonite
and muscovite, and incubated for 451 days, jarosite was formed. Approximately
20% of the montmorillonite was destroyed and the muscovite was partially con-
verted to an expanded phase (Ivarson et al., 1980). As the final pH was 2.0, it is not
clear how much of the degradation was due to the pH and how much was due to the
direct action of the bacteria. Nevertheless, the bacteria supplied the energy for the
reactions.
Wall et al. (1974) mixed sewage effluent with a calcareous clay and monitored
changes over a five-week period. The microbial population increased continuously,
the mixture went anaerobic, and the p H increased slightly from 7.3 and 7.5. The
relative percentages of illite, I/S, and vermiculite in the clay decreased and the
relative percentages of quartz and chlorite increased. They suggested the former
three minerals were being decomposed, quartz remained constant, and some chlorite
was possibly being created by the precipitation of hydroxy A1 in the vermiculite
interlayer space.
Boyle et al. (1967) found that when a biotite flake was placed in a glucose
solution inoculated with Aspergillus niger, light colored alteration bands developed
around the edges of the flake and gradually moved inward. X-ray analyses indicated
no vermiculite was formed. Chemical analyses of the liquid indicated octahedral
cations were removed in approximately the same proportion as they occur in the
octahedral layer, though slightly larger amounts of Fe and A1 were extracted.
Organic acids and HCl produced the same effect, edge dissolution, as the bacteria.
A fragile matrix of amorphous material, largely silica, remained. On the other hand,
in solutions of NaCl and CaC1, the K was replaced by Na and Ca and a
dark-colored rim of vermiculite developed on the biotite. Chromatographic analysis
of the A . niger culture indicated the bacteria had altered the glucose to oxalic and
citric acids. Presumably the destruction of the biotite is due to the chelating ability
of the organic acids, particularly the oxalic acid. Fe is a required nutrient for many
bacteria. They create chelators to extract relaively insoluble Fe in low Fe environ-
ments.
Examination of biotite and hornblende grains collected from the sublittoral sand
zones at a number of estuaries along the Connecticut coast (Frankel, 1977) indi-
cated they contained an abundance of diatoms, blue-green algae, and bacteria (Fig.
3-14). They were concentrated in depressions on the surface, along the junction of
the treads and risers produced by stepped cleavage, and at the edge of biotite flakes,
primarlily in protected sites. For biotite the microorganisms were most concentrated
inside the flakes in the protected wedge-shaped spaces between folia. These flakes
were so brittle that they broke into small pieces and split along cleavages when
touched. The physical disintegration was apparently caused by the stresses created
by the growth and reproduction of the organisms that penetrated into the cleavage
and fractures. The mineralogy and chemistry was not investigated, but it is likely
that bacterial acids produced significant chemical degradation.
A study of the weathering of feldspars in granite (Parfenova and Yarilova, 1962)
showed that during the earliest stages of weathering feldspar surfaces become
139

Fig. 3-14. Blue-green algae (popcorn balls) and bacteria on biotite flake. Courtesy L. Frankel.

covered with a film of clayey products and with lamellae of sericate. These films
contain many algae, bacteria, and fungi. Montmorillonite was observed covering the
filaments of the blue-green algae.
When a variety of igneous rocks were added to nutrient solutions inoculated with
the fungus Penicillium simplicissimum, isolated from weathered basalt, large amounts
140

of ions were solubilized. After only seven days’ incubation as much as 31% of the Si,
11%of the Al, 64% of the Fe, and 59% of the Mg were solubilized. The basic rocks
were considerably more soluble than the acid rocks. During growth the p H of the
medium decreased from 6.8 to < 3.5 as citric acid was produced. As the same
amount of solubilization was achieved in spent (fungi dead) acid medium, Silverman
and Munoz (1970) concluded the dissolution was due to biogenic acid.
When the fungi Rhizoctonia sp. were grown for two weeks in a nutrient solution
containing various micas, biotite and phlogopite were completely converted to
vermiculite; only a small amount of muscovite was converted to vermiculite. Any
acids produced were neutralized by the buffered nutrient solution. The conversion
was accomplished because the fungi served as K+-sinks, keeping soluble K + at a low
level. Na+ in the solution replaced the interlayer K + (Weed et al., 1969).
Lichens consist of fungi containing symbiotic algae. They can rapidly colonize
fresh rock surfaces and increase the rate of mechanical and chemical weathering.
When ground, dry lichen (10 species) and distilled water were mixed with fine
granite, muscovite, and glauconite, Schatz (1963) found variable but appreciable
alteration (measured by changes in light transmission of the supernatant solutions)
of the rock and minerals. In some cases reaction started within a few hours. He
concluded the weathering was due to chelation by lichen acids. In most experiments
the pH of the solution was increased from - 5 to 6-7, presumably due to the
solubilization of alkalis and alkaline earths.
Lichens develop a wide variety of strong chelating acids (also ammonia) which
are called collectively lichen acids. These acids are readily available from both living
and dead hyphae. Lichen acids may average 2 to 5% of the dry weight of the lichen
and can be as high as 35%. The function of the lichen acids is, presumably, to
supply the symbiants with needed elements.
Jackson and Keller (1970) found that lichen on recent lava flows in Hawaii
accelerated markedly the rate of weathering. Si, Al, Fe, Ti, and Ca were extracted
from the lava. The AI:Si ratio in the lichen was higher than that of the rock,
presumably due to chelation by the lichen acids.
Studies reviewed by Parfenova and Yarilova (1962) described the role of lichen in
both physical and chemical weathering. When moist lichens dry out, the resulting
contraction detaches fine fragments of rock which are absorbed in the lichen’s body,
where they are readily dissolved by lichen acids. Micas are a source of K, and the
hyphae grow along the cleavage planes until the mica becomes comminuted and
cloudy, eventually altering to kaolinite. The Fe is removed from biotite, leaving opal
flakes. Similarly, the Fe in chlorite is oxidized and removed. Ca-feldspars are
attacked more readily than Na-K feldspars. The hyphae penetrate the Ca-feldspar,
concentrating Ca, along cleavage planes and break the mineral into minute parts.
K-feldspars weather much less than Na-feldspar. The sericitization of Na-feldspars
is believed to be due to the concentration of K in the residue of lichen. In the
transition from fresh lichen residues to decayed residues, the molecular ratio of
SiO,/AI ,O, changes, decreasing from 10.3 to 4.4. X-ray investigations suggest the
residue contained montmorillonite and illite. Unfortunately, the identification was
inconclusive.
141

Weaver (1967) reviewed the literature on the concentration of K by various


organisms. The ash of various primitive organisms contains from 20 to 40% K 2 0 . I
suggested that the residue of these organisms, both marine and terrestrial, may have
provided K-rich solutions which allowed illite (or I/S) to form during the Pre-
cambrian and early Paleozoic.
Numerous studies have demonstrated that K can be stripped from biotite
relatively easily. One example shows the ability of wheat to convert biotite to
vermiculite (Cecconi et al., 1975). They grew wheat in biotite-containing cultures to
which nutrient solutions both containing K and free of K were added until ripening.
Periodic analyses of the biotite indicated that through the tillering, boot, and
flowering stages K was systematically extracted from the biotite. In the experiment
where there was no K in the nutrient the biotite altered to vermiculite; where K was
present in the nutrient the biotite altered primarily to a mixed-layer B/V. At the
ripening stage, where plant growth presumably decreased, some K was readsorbed
by the vermiculite layers. In the K-deficient experiment vermiculite was converted
to B/V; where K was present in the nutrient the B/V that had developed reverted
to biotite. The CEC decreased from 203 meq/100 g for the original biotite to a low
of 160 for the ripening samples. Oxidation of Fe*+ to Fe3+ accounted for some of
the reduction; but other processes, such as absorption of H + or release of OH-,
were involved. The layer charge remained high enough so that when treated with
NH; (acts the same as K + ) the expanded layers contracted to 10 A.
Experiments conducted by Spyridakis et al. (1967) demonstrated the effectiveness
of higher-order plants in chemical weathering. They planted tree seedlings in a
mixture of biotite and quartz. After a period of only 13 months as much as 50% of
the biotite was converted to kaolinite and much of the remainder to vermiculitic
material. There was also an appreciable decrease in particle size. The order of
kaolinization effected by the different tree species is: white cedar > hemlock > white
pine > white spruce > red oak > hard maple. In a planting (Monterey pine) where
organic material (root sloughings) and more nutrients were added to the biotite-
quartz medium, only vermiculite and B/V developed.
The initial pH of the cultures ranged from 5.2 to 5.4; the final p H ranged from
4.2 to 4.7. H 3 0 + from the roots exchanged with ions in the biotite. The uptake of
some of these ions by tree roots maintained a low ionic solution activity. The culture
medium was subjected to fairly rapid wetting and drying cycles. These conditions
favored the formation of kaolinite from the decomposed biotite. In the case where
only vermiculite was formed, the ionic solution activity was believed to be relatively
high due to the presence of metal cations in the root sloughings that were added to
the culture.
In the southeastern United States temperatures and rainfall are relatively high
and organic matter is decomposed relatively fast. Buffering capacity and ionic
concentrations are low and kaolinite forms in the red and yellow podzols. In the
northern podzols organic matter accumulates and little kaolinization occurs.
In addition to the nutrient ions, many plants concentrate large amounts of SO,.
Horsetail, sedge, reeds, and bamboo contain about 10% SiO, (dry weight). Grasses,
tropical lumber trees, and some temperate hardwoods are silica accumulators. At
142

least some of this material is returned to the soil as opal (Lovering, T. S., 1959).
Some of the plants also accumulate appreciable amounts of Fe.
The Si sequence, soil quartz + Si in solution ( - 6 ppm) + plant + opal + soil
opal + Si in solution (110 ppm), could increase the concentration of silica in
solution and influence the type of physil that would form.
It is evident from the studies on the weathering of biotite, one of the most easily
weathered minerals, that chemical and biochemical weathering are complex and that
a given mineral can be altered to a variety of end products. In the case of biotite it
can be altered to vermiculite, smectite, illite, kaolinite, chloritic material, or simply
destroyed. Under marine conditions it can be altered to glauconite.
Though it has been established that biochemical weathering is a major factor in
the solubilization of minerals and the formation of physils, it has not received the
study it deserves by clay mineralogists. Considerable information is available on the
destruction of primary minerals, but relatively little is known, with certainty, about
the physils which form as a result of biochemical activity. This is a promising area
for research.
In addition to the apparent extensive role played by microorganisms in the soil
environment, various organisms can alter and aid in the formation of physils in the
marine and lacustrine environments. Most notably is the formation of glauconite
and chamosite (pp. 386-396) and perhaps other physils (p. 296) by marine organisms.
Indirectly, or perhaps directly, diatoms contribute to the formation of a number of
physils (p. 338) by supplying Si for their formation.

PEDOGENIC FORMATION OF PHYSILS

We have examined some of the processes and factors that are involved in
chemical weathering and the formation and distribution of physils and soils. This
section is intended to be more geologically oriented. After a geologist determines the
physil suite of a set of samples, he/she first must try to establish if the physils are
detrital, authigenic, or diagenetic (or any combination). He/she then attempts,
among other things, to deduce the emvironmental and paleoclimate in the source
area and/or at the site of deposition. Finally, the physil data are integrated with
other geologic data and the paleogeography is synthesized.

Kaolin and Related Materials

Tropical Climate
Kaolin minerals, bauxites, and laterites are generally most extensively developed
in humid tropical climates. However, kaolin is not always the most abundant physil
in tropical soils and it can be a major physil in soils formed in a temperate climate.
Allophane and imogolite, which convert to halloysite and kaolinite with time, are
the most important components ( > 5091) of soils derived from volcanic ash and, to a
lesser extent, basalt (Ando, Andepts, and Humic Allophane soils). The pH levels
143

over which allophane and imogolite form usually range from 5 to 7. Andepts
(Inceptisols formed on pyroclastic material) occur from the cold subhumid to the
humid tropics and desert and semidesert regions (Wada, 1977). At least where
volcanic ash is the source rock there appears to be no climatic control of the
alteration product, though the amount of alteration should bear some relation to
climate. There appears to be no evidence that allophane ever alters to a 2:l physil,
though under moderate leaching and high pH conditions volcanic ash alters directly
to smectite or vermiculite.
Allophane or amorphous material forms from material other than volcanics and
though it forms in temperate regions it is generally more abundant in tropical soils
(Van Houten, 1972).
The young volcanic ash from the Paricutin volcano of Mexico has partially
altered to spheroidal halloysite (Dixon and McKee, 1974). A similar halloysite is
present in soils (near-neutral pH) developed on recent basalt of Central Cameroon
(Siefferrnann and Millot, 1968). Spherical halloysite also formed (33%) in 8,000 to
9,000years in a volcanic ash near Mount Aso, Japan (Aomine and Miyauchi, 1963).
Younger volcanic soils contain allophane. These soils are considered to be examples
of very young soils containing one of the easiest-formed physils, yet it still takes on
the order of 10,000 years or more for halloysite to develop from volcanic glass. This
is approximately equivalent to all of Holocene time. Older profiles developed on
either acid or basic igneous rocks typically consist primarily of kaolinite and quartz
in the lower and middle sections with gibbsite and goethite being more abundant
near the surface. In the A horizon, 40 to 60% of the initial rock may have been
removed by the circulating fluids (Lelong, 1969; Mohr and Van Baren, 1954). The
kaolinite-gibbsite distribution in a profile is often irregular, in part due to resilica-
tion and the formation of kaolinite from gibbsite.
If appreciable water is available, not necessarily rainfall, and can move freely
(relief, porosity, permeability) virtually any rock or aluminosilicate mineral will alter
to kaolin-gibbsite-Fe oxides suite. The relative amounts of the various phases
depends largely on the water flux. Percolation is more effective than surface flow. In
the humid tropical climates gibbsite tends to develop in the highlands and kaolinite
in the lowlands.
Tamura et al. (1955)describe soils formed from basalt in high rainfall areas of
Hawaii where hematite (48%) is dominant in the A, horizon; in the B horizon
gibbsite (31%), hematite (25%), and amorphous aluminosilicate (22%) are present in
nearly equal amounts. Gibbsite (48%) is dominant in the C horizon; Fe oxides total
14% and amorphous aluminosilicates 28%. In some Hawaiian soils TiO, is a major
component. In humid tropical climates with a pronounced dry season, kaolinite can
predominate in all soils. In semi-arid regions with long dry spells Fe-rich montmoril-
lonite develops at the bottom of soil profiles and in downslope regions where ions
are concentrated by evaporation and permeability is low (Tardy et al., 1973).
Herbillon et al. (1981)found that under the hot and dry climatic conditions at
the bottom of the African h f t Valley, Rusizi Valley in Burundi, Central Africa, the
basalts had weathered to smectitic soils. The flatter and depressed parts of the
landscape contain black smectitic soils (Vertisols). On the steeper slope areas
144

f
7.0
10.0 hJ
17.7

H0r.O- 50m
VWt?- 15m

Fig. 3-15. Relation of physils to slope in the savanna region of Kenya. The solid lines show the slope and
X-ray patterns illustrate the changes in physils with depth and location on slope. Kaolinite (7A) is
dominant in the steeper slope areas (left) and montmorillonite (17.7 glycerol) in the flatter areas. From
Kantor and Schwertmann, 1974, Jour. Soil Sci. Copyright 1974 Blackwell Sci. Pub.

(Ultisols), the smectite is transformed to kaolinite via a mixed-layer


kaolinite/smectite stage. The released Fe formed Fe-oxides which gives the soil a
red color. The extent of the transformation of smectite to kaolinite increases as the
steepness of the slope increases (up to a point) and from bottom to top within such
soil profiles. Mixed-layer K/S is common in the poorly drained soils in the humid
and subhumid parts of Nigeria (Okusami, 1985).
Kantar and Schwertmann (1974) describe a similar kaolinite-montmorillonite
pattern in the savanna region of Kenya (Fig. 3-15) where the rainfall averages
around 100 cm/year with 3 to 4 dry months. They concluded that the smectite is the
first weathering product of the underlying basic igneous rocks. However, they note
that the smectites concentrated in large flat basins in the Sudan, Ethiopia, and parts
of Kenya are detrital and derived from adjacent basaltic terrain. The pedogenic
smectite developed from basalt is a ferriferous member of the montmorillonite-be-
145

Fig. 3-16. Soil profile develop on granite in tropical Malaysia. From Eswaran and Bin, 1978. Reproduced
from Soil Science Society of America Journal, v. 42, No. 1, 149-153 by permission of the Soil Science
Society of America, Inc.

idellite series and is apparently typical of the relatively Fe-rich smectites formed
from basal ts.
Poncelet and Brindley (1967) were able to easily convert montmorillonite to
kaolinite by precipitating hydroxy-A1 between the layers, forming a “dioctahedral
chlorite” and subjecting the material to hydrothermal temperatures around 200”.
Presumably a similar reaction could occur at lower temperatures over a longer
period of time. Acid waters partially dissolve the octahedral sheet and some of the
A1 is precipitated in the interlayer regions.
Numerous studies of basalt and volcanic ash on Hawaii have demonstrated the
close relation between rainfall and intensity of weathering (Sherman, 1952; Tamura
et al., 1953; Bates, 1962). Gibbsite and Fe-oxides are predominant in soils where the
rainfall is in excess of -
200 cm; smectite is dominant where rainfall is less than
- 100 cm; kaolinite and halloysite are abundant in the intermediate rainfall range.
Allophane is apparently formed both when basaltic material alters to halloysite and
when kaolinite and halloysite alter to gibbsite.
Fig. 3-16 shows the distribution of minerals in a soil profile developed on granite
in tropical Malaysia. This sequence shows an upward increase in kaolinite and
gibbsite at the expense of halloysite. Gibbsite becomes relatively more abundant in
the coarser size fractions (Eswaran and Bin, 1978).
Though, in general, in relatively high rainfall areas the composition of the parent
rock does not determine the type of physil that will develop during weathering,
Clemency (1975) describes an outcrop from near Sgo Paulo, Brazil, where a basalt
146

dike penetrating a schist or gneiss has altered to Fe-rich smectite, whereas the
schist-gneiss weathered to kaolinite. He suggests the two physils formed as a result
of a permeability difference. In Hawaii, Bates (1962) found plagioclase altering to
halloysite and adjacent olivine grains altering to montmorillonite.
The weathering of biotite has been studied under a great variety of condtions and
the tropics are no exception. Whereas in the temperate regions biotite frequently
alters to vermiculite and/or montmorillonite, in the humid tropics it tends to
weather directly to kaolinite or halloysite. They are often pseudomorphics of the
biotite. Acid leaching rapidly removes interlayer K and octahedral Mg and Fe. The
Fe is commonly precipitated nearby as oxides, but the K and Mg are usually flushed
from the immediate area of the biotite (DeKiempe and Tardy, 1967; Ojamega, 1973;
Eswaran and Heng, 1976). Muller and Bocquier (1985) found that the kaolinites
replacing micas in the tropics are Fe-kaolinite and with increased weathering
(upward) become more Fe-rich and less well crystallized.
Solubility studies indicate that gibbsite should be unstable relative to kaolinite in
most natural ground waters, which have a relatively high concentration of H,SiO,.
Resilicification of gibbsite, in bauxite deposits, to kaolinite has been described by a
number of people (for references see DangiC, 1985). Curtis and Spears (1971)
suggest t h s is a common phenomenon. Whereas kaolinite is relatively abundant in
ancient sediments, gibbsite is rare. They suggest the detrital gibbsite usually ends up
in an environment where the silica content is high enough to convert it to kaolinite.
Changes in water chemistry with time allow the in situ conversion of bauxite to
kaolinite.
The dominant physils and associated minerals in the humid and subhumid
tropical soils are kaolinite-halloysite, smectite, gibbsite, and Fe-oxides. Illite, chlo-
rite, and the chain structure physils are generally absent or present in minor
amounts; however, these physils can be abundant in arid tropical (desert) soils.
Chemically the most significant difference between the two groups is the relative
abundance of readily soluble Mg in the arid physil suite.
Though kaolinite forms under temperate conditions, thick oxisolar latosal pro-
files dominated by kaolinite and quartz form today on quartz-rich plutonic rocks
only in near-equatorial areas (20-30" north and south) with a mean annual
temperature around 25" and a minimum annual precipitation of more than 100 cm
(Singer, 1980).

Temperate Climate
Though kaolinite is considered to be characteristic of the humid tropics, it forms
readily in the humid temperate regions, particularly in the warm temperate or
subtropical regions. One problem with determining the significance of soil physils in
the temperate region is that many of the soils were formed millions of years ago
when the climate was generally warmer. In general, precipitation was significantly
greater in the Paleogene than now. Precipitation decreased near the end of the
Eocene to a level similar to that of the present (Frakes, 1979).
The distribution of kaolinite-rich soils and clays in the relatively young temperate
zone sediments in general reflect the Cenozoic climatic pattern. Well-developed
147

quartz-kaolinite oxisol profiles of Paleogene age have been identified in Baja


California (Abbot et a/.,1976) and Poland (Stoch and Sikora, 1976). Paleogene
laterites are present on the Siberian platform (Epshteyn, 1978) and Ireland (Frakes,
1979). Extensive, and probably the most, kaolinite and bauxite deposits were
developed during the Cretaceous and Paleogene; many of these deposits are in
Europe and United States (see Chapter IX). The thick, kaolinite-rich saprolites of
the southern Appalachian Mountains probably developed during the Paleogene.
Paleogene laterite and bauxite profiles, indicative of tropical to subtropical weather-
ing, extended to at least 45" paleolatitude in both hemispheres (Frakes, 1979).
To repeat, in the present-day temperate zones it can be difficult to determine
which physils are the stable phase. In part this is due to changing climatic
conditions with time and the youthfulness of the soil profiles developed on Pleisto-
cene glacial detritus. Soil water analyses from temperate regions indicate kaolin
should be the physil in equilibrium with the water in most soils (Garrels and
MacKenzie, 1971). Thus, one line of reasoning is that, assuming some water
movement, the ratio kaolinite/total physil suite is a measure of the soil's age,
assuming physils are present in the parent rock. If water conditions are appropriate
(slightly acid and moving), kaolin should be the stable physil formed from igneous
rocks in temperate regions; due to lower temperatures, the weathering rate should
be lower than in the tropics.
Another process that operates at different levels in the tropical and temperate
zones is the destruction of organic material in the soil. Temperatures of 25°C or
above promotes destruction of macroflora by microfauna (Keller, 1964). As dis-
cussed earlier organic matter can be a major factor in increasing the mobility of Fe
and A1 and decreasing the mobility of Si.
We can get an idea of the formation and amount of kaolinite in the temperate
zone soils by examining north to south weathering trends along the east coast of
North America. In general the kaolinite content of soils increases from north to
south. Kaolinite, and associated A1 and Fe compounds, is the dominant physil in
many of the soils of the southeastern United States, where the climate ranges from
subtropical to warm temperate. However, as in the tropics, montmorillonite is
abundant in some soils. Kaolinite is the most abundant physil in both the Coastal
Plain and Piedmont soils of the southeastern United States, as far north as Virginia.
Most of the soils are Ultisols. Most of these soils contain appreciable amounts of
vermiculite and/or Ch/V and many contain gibbsite (Southern Cooperative Series
Bull. 61, U.S. Dept. Agri., 1959; Fiskell and Perkins, 1959). The relative abundance
of vermiculitic material tends to distinguish these soils from the kaolinite-rich
humid tropical soils. Also, gibbsite is usually a minor component. Most pH values
are between 4 and 6. Locally, montmorillonite rich soils develop on chalk (pH 7 to
8). As in the tropics the nature of the parent rock is generally of little consequence.
Kaolinite develops from montmorillonite present in Recent and Cenozoic sedi-
ments, on basic and acid igneous rocks, and metamorphics.
Over large areas of central Florida (Altschuler et al., 1963) and southern Florida
(Weaver) and probably coastal Alabama and Mississippi, montmorillonitic parent
material has been altered to kaolinite, often via a mixed-layer K/S stage. The
148

Fig. 3-17. Alteration of Sparta Granite, central Georgia. (A) Tuffs of halloysite along microfractures (?)
in feldspar. (B) Enlarged view of halloysite. (C) Halloysite and books of kaolinite in the lower part of the
saprolite over Sparta Granite. (D) A vein filling of halloysite. Bar A = 5 pm; B,C,D = 1 pm. From Keller,
1978. Copyright 1978 The Clay Miner. Soc.Photographs Courtesy W.D. Keller.

process is apparently one of desilication as silicified fossils and carbonate rocks are
commonly present.
Fig. 3-17 shows the systematic alteration of feldspars in the Sparta granite
(central Georgia), to kaolin (Keller, 1978). During the early stages of weathering
tuffs of halloysite (or elongate minerals) grow from dissolved feldspar Si and Al.
Eventually grass-like mats of halloysite develop. Above the transitional granite
149

alteration zone, but near the base of the overlying saprolite, both halloysite and
platy kaolinite are present. Several meters upward in the saprolite, books of
kaolinite predominate. The saprolite, called “weathering crusts” by European
geologists, form over long periods of time, commonly below the ground-water table.
In the Sparta example, much of the saprolite may have formed during Mesozoic
time. Keller (1978) suggests that the halloysite forms under “ high-land conditions”
where the movement of ground water is rapid and the chemical activity HOH is
relatively high; halloysite can later alter to kaolinite, or the kaolinite crystallized
directly from solution, under constant water, more nearly equilibrium conditions
(“low-lying terrain”). Recent weathering of the saprolite has produced some corro-
sion of the kaolinite books and a general decrease in grain size (Keller, 1977).
In eastern Alabama, Ashland and Opelika Plateaus, saprolite developed on
granite-gneiss shows the effects of more intense weathering than in the Sparta area.
Gibbsite is a major component, up to 308, of the upland, well drained soils, and
kaolinite of the valley soils with restricted drainage. In the well-drained regions
plagioclase alters directly to gibbsite and orthoclase to kaolinite. In the poorly-
drained region (high Si waters) both feldspars alter to kaolinite. Biotite alters to
kaolinite via a clay mica, I/V stage (Clark, 1973).
In the Kings Mountain Belt near Gaffney, South Carolina, the only physil in
saprolite formed on biotite gneiss is an Fe-rich kaolin, pseudomorphic after biotite
(Weaver, 1978). The physil contains 8% Fe,O, and may be a form of halloysite:
- (si1.88A10.12 ) (A11.74 Fei.10
Mn26t,)O5 (OH),*
In a study of residual kaolin derived from feldspathic rocks in the Southern
Appalachians, Sand (1956) noted that kaolinite was dominant in Georgia and
Virginia samples, with halloysite (10 A) increasing in abundance towards the middle
of the region (North Carolina). He observed that where conditions were favorable,
Blue Ridge Province of North Carolina, halloysite (10 A) formed by the weathering
of all types of feldspar. Mica altered to vermicular kaolinite. To the north, south
and east of this area K-feldspar commonly altered to “books” of mica which in turn
altered to small books of kaolinite. Most of the weathering apparently occurred
during the Cenozoic.
Sinex (1975) studied saprolite profiles developed on three different rock-types in
a small area of the Blue Ridge (Ashe County), North Carolina. The initial physil
formed (saprock) differed for each rock-type, but the “final” product was kaolinite
in all cases:
-
Amphibolite Montmorillonite + Kaolinite
Gneiss + Vermiculite -+ Kaolinite
Ultramafic + Chlorite 4 Kaolinite
Kaolinite is the predominant physil of Piedmont soils as far north as Pennsyl-
vania (Johnson, 1970). Varying amounts of gibbsite, vermiculite, illite, and
montmorillonite are present in some samples.
There is little doubt that most of the kaolinite in the Appalachian Region was
formed during the Tertiary when climates were, in general, warmer and more humid
150

I
Kaolinite

.I'
1

1.5

10

Srnectite

Fig. 3-18. X-ray patterns and TEM pictures of altered Miocene clay section from central Georgia.
Smectites altered to kaolinite. Amorphous material (subspherical, dehydrated particles in lower TEM) is
present in the early stages of alteration. Thickness and idiomorphism of kaolinite plates increases
upward. White bar = 0.1 pm.
151

than today; however, in the southeast kaolinite is still forming, primarily from
plagioclase and montmorillonite.
Analyses of thin soil zones (not saprolite) and friable weathered “rotten” rock
developed on granitic rock in the Atlanta, Georgia, area indicates kaolinite (occa-
sionally halloysite or dickite) is almost always the dominant physil; appreciable
gibbsite is present in most samples. Varying amounts of mica, vermiculite, and I/V
are present in some samples. The kaolinite formed primarily from plagioclase and
vermiculitic material from mica. Analyses of a large number of Coastal Plain wells
(Weaver and Beck, 1977) that started in Miocene and younger sediments indicated
that the montmorillonite, which was the predominant physil in the Tertiary sandy
clays and clayey sands, was systematically altered to kaolinite, commonly via a
K/M phase (Fig. 3-18). In many wells kaolinite is the only or predominant clay to a
depth of around 7 meters and gradually decreases to a depth of around 17 meters.
Based on the age of the parent rock the conversion of montmorillonite has been
going on since the late Miocene ( - 10 m.y. B.P.) and is likely continuing at the
present time.
The kaolin content of soils decreases to the north of the southeastern United
States, where the climate is cooler, to the west (dryer), and to the northwest (dryer
and cooler). In addition, large areas outside the southeast are covered with loess and
glacial deposits and contain relatively young soils. Studies of soil profiles in these
areas commonly show a slight increase in kaolinite ( < 10%)in the A horizon and in
older soils (Novak et al., 1971). Most of these soils have an acid pH and if preserved
long enough should develop a hgh kaolin content. Presumably it will take apprecia-
bly longer than in the southeast, where it is warmer and wetter.
In the temperate regions rainfall and relief can be the major factors that influence
the formation of soil kaolin. In northwestern United States in the high rainfall area
west of the northern Sierra Nevada and the Cascade Mountain Ranges kaolinite and
halloysite are commonly the dominant soil physils developed on a wide variety of
parent materials but primarily igneous acid and basic igneous rocks (Barshad, 1966),
tuff and pumice (Robertson, 1963), Wisconsin and older Tills (Birkeland, 1974), and
volcanics (Taskey et al., 1978). In soils developed on acid igneous rock vermiculite
can be as abundant as kaolin. The abundance of kaolin decreases to the south and
east where the climate is dryer and warmer.
Japan provides another example of extensive kaolin formation under temperate
conditions. The climate ranges from warm temperate to cold humid (snow); rainfall
is in excess of 100 cm/yr. Kaolin minerals are the most abundant and widely
distributed constituents of Japanese soils (Sudo and Shimoda, 1978). Kaolin miner-
als form from volcanics (along with allophane and imogolite), granites and sedimen-
tary rock. Hydrothermal kaolin minerals are relatively abundant. It is of importance
to note that hydrothermal activity is commonly extensive in volcanic provinces and
relatively large volumes of physils, usually not kaolinite, are formed from these
warm waters. Also, from the standpoint of relation of marine physil suite to
pedogenic physils and climate, it should be noted that the marine muds surrounding
the Japanese islands contain only minor amounts of kaolin.
Mixed-layer kaolinite-smectite is widespread in altered volcanic material in Japan
152

(Sudo and Shimoda, 1978). In addition to numerous occurrences in North America,


it has been found in Africa and Scotland (Wilson et al., 1972) and is widespread in
Australian soils (Norrish and Pickering, 1983). It appears to be a common alteration
product, produced during acid weathering of smectite, that has not always been
identified.
Thick weathering crusts in which kaolin minerals (plus gibbsite and Fe oxides)
are the predominant physils are common in such temperate climates as New South
Wales (southeastern Australia) (Loughnan, 1969) and Western Australia (Gilkes et
al., 1980) and Europe (Millot, 1970; Storr, 1975; Stoch and Sikora, 1976). These
crusts have developed on a wide variety of rocks including basalts, granites, and
sediments. Much of this weathering occurred during the Cenozoic, or older, and has
continued, probably with diminished intensity, up to the present. The distribution of
commercial Cenozoic kaolin deposits indicate they occur as far north as 40 to 55"
paleolatitude (Weaver and Beck, 1977). As the global climate was considerably
warmer during the Cenozoic than at present, this is not surprising. Large kaolin
deposits are indicative of warm, humid climates but are not necessarily located near
the paleoequator. For example, kaolin is the dominant mineral in 60% of the soil
samples from Australia and is a major component of 90%; however, samples from
higher rainfall regions are over-represented (Norrish and Pickering, 1983).

Discussion
Though kaolinite, bauxite, and laterite deposits are considered to have formed
under humid tropical and subtropical conditions, it is apparent that at least
kaolinite can form in cool temperate climates under conditions of moderate (100 to
200 cm) rainfall. However, for appreciable thicknesses of weathering crusts to form,
long periods of leaching are required. Thus, in southeast United States, southern
Australia, and parts of Europe extensive leaching has persisted, with varying degrees
of effectiveness, for tens of millions of years, and thick kaolinitic and lateritic
profiles have developed. Theoretically, thick lateritic crust could develop in Japan in
the future but it could take a long period of time. Less time is required in warmer
regions to form kaolin, and it is probably why the larger kaolin and bauxite deposits
are restricted to the warmer climates.
In addition to a large water flux over a long period of time, to form thick
kaolinite deposits, it is necessary to have a relatively smooth terrain and quiet
tectonic conditions so that chemical weathering can be more effective than erosion.
As kaolin minerals form more readily from volcanic glass and feldspar (granitic
rocks), parent material is a factor. Though virtually any rock type can alter to a
kaolinitic crust, granitic and volcanic rock weather more rapidly than most sedimen-
tary rocks.
A concern of the geologist is, what is the significance of a high concentration of
kaolin in a shale sequence? Let's look at the recent marine sediments (Fig. 5-18).
Kaolin physils are abundant in the soils of Japan and the southeastern United
States, but there is no evidence of this in the surrounding marine sediments. In
Japan the kaolin soils have formed primarily on young volcanic rocks and are
presumably relatively thin. Thus, their contribution to the marine physil suite is
153

relatively small. In the southeastern United States large volumes of kaolin are
delivered to the coast by the major rivers (p. 203); however, much of it is trapped in
the coastal marshes and near-shore area (Fig. 4-41). In the continental shelf
sediments there is little evidence of a kaolin-rich source area. The distribution in this
area appears to be controlled by strong marine currents and a rising sea level.
There are four major areas with high (> 25-308) concentrations of kaolinite.
Two occur in the equatorial region ( - 30"N to 30"s) of the Atlantic Ocean and the
other two occur off the east and west coasts of Australia. The Atlantic occurrence
fits the model to some extent. High kaolin concentrations are offshore of a humid
tropic region; however, relatively high kaolin concentrations also occur seaward of
the Sahara Desert and are believed to have been transported by Saharan winds (p.
315). Most of Australia has a warm temperate or desert climate (tropical along the
northern edges). Thick kaolin and lateritic crust, formed during the Cenozoic, occur
in both the desert (west) and temperate (east) regions. The highest kaolin concentra-
tions occur offshore from the Western Australia Desert. These kaolins are ap-
parently largely wind transported (Griffin ef al., 1968), as are at least some of those
off the eastern coast (Windom, 1976).
Abundant kaolinite in marine sediments is likely to indicate that a humid and
relatively warm climate existed for a considerable time in the source area. However,
on the basis of the physils alone, it would be difficult to determine if the
warm-humid climate existed at a time penecontemporaneous with the age of the
marine sediments, or if the kaolin reflected climatic conditions tens of millions of
years earlier. Of course, it is also possible, as in the present Arctic Ocean, that the
marine kaolin could have been derived from rocks, rather than soils, hundreds of
millions of years old.
A comparison of the Australian region with the southeastern United States and
Japan indicates that wind transported physils are more likely to reflect the composi-
tion of the soils than river transported physils. Commonly, a high proportion of
river-borne physils are derived from bedrock and the C horizon.

Smectite

Smectites of various types can form from essentially any mineral and rock that
kaolin minerals can form from. Montmorillonites-beidellites are by far the most
abundant smectites in soils. As noted previously, for smectite to form it is necessary
to retain (or add) many of the ions that are flushed out of the system during the
formation of kaolin. Specifically it is necessary to retain much of the silica
(SiO,/Al,O, kaolinite = 1.2, montmorillonite = 2.7), some Mg and/or Fe, and an
exchangeable cation, Na, Ca, Mg, though H can serve as the exchangeable cation.
The climatic and topographic conditions necessary for the formation of smectite are
basically the opposite of those that favor the formation of kaolinite. Thus, low relief
and/or low permeability, low rainfall and/or low water flux, and low temperature
favor the formation of smectites.
Montmorillonite can form by the alteration (transformation) of volcanic glass,
154

feldspars, micas, various FeMg silicates, and silication of detrital physils. It can also
precipitate directly from solution (neoformed). Thus, montmorillonite can form in
basins where seepage supplements the sediments as a source of solutes; in the B
horizon of soils where the water and its solute load is arrested by a “hanging water
table” and the water is removed by evaporation and phyto-transpiration; in Verti-
sols where the montmorillonite provides a large water-holding capacity and shrin-
kage and swelling engulfs rock fragments from the C horizon as a supplemental
source of cations (Vertisols may be thought of as a “super B” horizon, able
periodically to engulf the A horizon and part of the C horizon); in protected areas
(leaching shadow) in soils containing basic rock fragments, including loess. In
temperate climates montmorillonite commonly forms in the B horizon but can form
in the A horizon of some Spodosols (Podzols). In tropical and subtropical climates
montmorillonite forms mostly in drainageways and basins (Jackson, 1965). Where
evaporation is extensive Mg-rich smectites can form.
Smectite-rich, largely montmorillonitic soils that have a high shrink-swell poten-
tial, which requires distinct wet and dry seasons, are called Vertisols (formerly
Grumusols). Moderately weathered soils with significant quantities of smectite have
basic pH values. A world map of soil physils (Gradusov, 1974) indicates that nearly
half the soils in the temperate zone contain more than 50% montmorillonite (and
I/S). Several large areas rich in soil montmorillonite occur in the tropic zone; one
occurs in northwest India and is primarily due to the weathering of basalt (Deccan
Trap). High concentrations of montmorillonite in the Nile Valley and associated
lacustrine sediments are believed to be due to the river transport of montmorillonite
formed on basic volcanic rocks at the headwaters of the Nile. Much of this material
was apparently deposited during the Pleistocene (Borchardt, 1977).
Montmorillonite (and I/S with a high smectite content) is abundant in sedimen-
tary rocks, particularly those of Mesozoic and Cenozoic age. Commonly it is
difficult to determine if the montmorillonite is detrital or if it altered in place from
transported volcanic material. If it is detrital, as in the Mississippi Delta, then the
problem is to determine if it formed under continental conditions, soil and lacustrine,
or was derived from older rocks. In the case of the Mississippi Delta physils, where
montmorillonite ( - 80% of < 2 p m fraction) has been deposited for the past 30
m.y. and longer, the detrital montmorillonite was derived both from Cretaceous and
older Cenozoic rocks, both marine and non-marine, and Pleistocene loess. Soils in
the western part of the drainage basin mostly contain more than 50% montmoril-
lonite, some of which is inherited.
Even for the recent Mississippi Delta deposits it is difficult to determine the
original source of the montmorillonite. Presumably, the ultimate source was prim-
arily volcanic debris created by the extensive Cretaceous to recent volcanic activity
in the western United States. This volcanic material was altered to montmorillonite
under both continental and marine conditions and tens of millions of years later
transported to the delta.
In depocenters closer to volcanic provinces (northwest United States and south-
west Canada) thick sequences of volcanic debris can be deposited and altered after
burial.
155

It is commonly assumed that most montmorillonite was derived from volcanic


rocks. While this is probably true, there are other major sources such as micas,
feldspars and FeMg minerals.

Tropical Climate
The origin and occurrence of smectite in the tropics was discussed, to some
extent, in the section on kaolinite. In general, basic volcanic rock alters to smectite
in areas where rainfall is low and/or the slope is low and drainage restricted.
Extensive vertisols occur in large flat basins in Sudan, Ethiopia, and Kenya. The
montmorillonite is derived from the weathering of basalt. In some basins the
montmorillonite is detrital and in others it forms by in situ weathering and is
concentrated in the lower part of the profile (Kantor and Schwertmann, 1974). The
smectite in the Kenya Vertisol and in other Vertisols typically has a high Fe
content:
( .28 Fei.zl
FeL?.& Mg0.24)(si3.80 0.20 1’10 0.35

Smectite-rich vertisols are common in Chad, north-central Africa (tropical with


wet and dry seasons) (Paquet, 1970). Those formed on basic rock such as basalt and
amphibole gneiss are ferric beidellites,
Mg0.Z8Ti 0.16 ) (si3.28 A10.72 )OlO (OH),
.O Fe0,77 9

whereas those developed on granites (feldspar) and shales and clays tend to have
less Fe and tetrahedral Al:
.33Fe0.49Mg0.21Ti0.06)(si3.59A1 0.41 )OIO (OH),.
The latter two smectites more closely resemble beidellite than montmorillonite.
Smectites developed on the basalts of the Deccan Plateau, India, also have a high
content of tetrahedral A1 and octahedral Fe and more closely resemble beidellite
than montmorillonite (Singh and Murti, 1975).
(A11.17’Fe~.&Mg0.25)(Si3.4, A10.58 )OIO (OH)ZX0.52

These calcareous vertisols developed on low relief basaltic flows under tropical
conditions. Though rainfall is fairly heavy (50 to 100 cm), it is concentrated in the
four monsoon months and the climate is predominantly dry (Roy and Barde, 1962).
In the same region smectite-rich vertisols develop on granites containing Ca and
CaNa feldspars.
During weathering in a poor drainage environment, the soils developed from
ultrabasic rocks (serpentine), Sabah, Borneo, incorporated much of the original Fe
in nontronite. Much of the Mg was lost from the soil, though some was retained in
chlorite (Eswaran and Sys, 1972). Nontronite is the predominant physil developed
from ultrabasic rocks (olivine pyroxenite) from the western Ivory Coast (Paquet et
al., 1982). There is considerable localized variation in composition. Tetrahedral A1
ranges from 0.03 to 0.56 and octahedral Fe from 1.00 to 2.04. Much of the Mg is
present in talc. A typical nontronite formula is:
(A10.06Fe~.~6M~0.Z9)(si~.59A10.41)OIO (OH),
156

Though it is more typical of temperate regions, Flexor et al. (1975) found that in
some humid tropical regions of Brasil (Bahia and ParB) illites and micas were
altered to montmorillonite via mixed-layer and vermiculite phases. Similarly, in
Madagascar, under restricted drainage conditions biotite, chlorite, and talc in gneiss
and schists alter to montmorillonite (Paquet, 1970) and in Zaire chlorite schists alter
to vermiculitic material which alters to nontronite in the clay fraction (Herbillon
and Makumbi, 1975):

Chlorite

( .33Fei.;7 Fei.kMg2.36Ti0.01Mn 0.03)(si 2.90 .I0 )OIO (OH),.

Nontronite

(A 1 0 , 7 6 F e ~ . & T i 0 . 0 9 M ~ 0 . 1 6 ) (si3.51 0.49 I o l O 0.17Mg0.06Na0.07K 0.01

It is apparent that smectites are relatively wide-spread in the tropics even where
the rainfall is high; however, extensive deposits are largely restricted to areas
containing relatively flat-lying basaltic rocks. It forms readily in localized low,
depressed areas where rainfall is restricted. Thus, in India the rivers draining the
basaltic Deccan Traps are transporting a smectite-rich physil suite to the ocean. In
contrast, the Niger River drains a tropical region, where variations in relief and
drainage determine that montmorillonite and kaolin can form side-by-side, and
carries a physil suite containing both montmorillonite and kaolinite, with minor
illite (p. 255). Most of the smectites formed in the tropical soils tend to be Fe-rich
beidellites rather than true montmorillonites. Fe is apparently preserved in these
tropical soils where complexing organic acids are relatively sparse. From the
standpoint of soils as a source material, the amount of clay formed can be as
important as the mineral composition. Vertisols have a particularly high clay
content, commonly 50 to 70% < 2 pm, which consists largely of smectite.

Temperate Climate
In temperate regions smectite is neoformed from basic volcanic material as it is
in the tropics. More so than in the tropics, smectite forms by the transformation of
other 2:l sheet silicates. As shown in Fig. 3-6, smectite is the predominant physil
forming from both basic and acid igneous rock of the west coast of the United
States (California), in areas where the rainfall is < 25 cm.
Ismail (1975) found biotite in granite from the Mojave Desert, southern Cali-
fornia, altered to smectite and to a lesser extent vermiculite. The pH values
increased from 7.2 to 8.6 with depth as the amount of smectite increased, particu-
157

larly in the finer sizes. In another example, Mojave smectite is believed to have
formed pedogenically from ions derived from feldspar (Southard and Boettinger,
1987).
In the alkaline (pH > 7.8) soils of the Sonoran Desert (northern Mexico,
southern California, and Arizona) montmorillonite is the predominant physil in
most samples (B horizon, 43 samples) but is usually accompanied by significant
amounts of illite and small amounts of kaolinite. Illite, and to a lesser extent
kaolinite, is the predominant physil in about 20% of the samples. Alluvial samples
contain more kaolinite than the soil samples, presumably reflecting contributions
from the more intensely weathered highland source areas (Walker and Honea,
1969). In an earlier study (Walker et al., 1967) of arkosic and red beds from the
Sonoran Desert, it was shown that hornblende was altered by interstitial solution,
under arid conditions, to montmorillonite (2 to 13% Fe,03). The Fe is progressively
leached from the montmorillonite, with increasing distance from the hornblende
core, and precipitated as hematite.
The amount of soil physils formed under arid conditions is considerably less than
that produced in humid climates.
In the Pacific Coast Range of southern California quartz diorites weather under a
mediterranean climate to a depth of more than 30 m. In the foothills, where
permeability is low but drainage is good, biotite, andesine, and hornblende alter to a
beidellite-montmorillonite (Nettleton et al., 1970). Where the permeability is higher
kaolinite tends to be more abundant. The smectite formed from the andesine
feldspar contains considerably less Fe and tetrahedral A1 than those formed in
tropical soil.
) (si3.72 A10.28 )OlO
.31Fe0.30Mg0.37Ti0.04 i.03

C.E.C. 119 meq/100 g.


In the Western Cascades of Oregon, where the rainfall averages 200 cm/year, the
fine-grained basaltic and andesitic colluvial material alters to halloysite (10 A) and
allophane, but the coarser andesite cobbles alter to smectite. Smectite forms from
material supplied by the dissolution of glass and plagioclase. In the initial stages of
formation the smectite grows as - 1 pm spherical aggregates of very small plates.
These spheres merge to form clay films or sheets. Smectite presumably forms
because of the restrictive drainage of the microenvironment of the andesitic clasts
(Glasmann, 1982). Soils developed on the Mazama pumice deposits of central
Oregon consist primarily of amorphous material and varying amounts of smectite,
chloritic physils, vermiculite, and micaceous material. Though the environmental
conditions are such that the kaolin minerals should form, it is believed that the
vesicular structure of the pumice inhibits the flow of water and ions enough to allow
2:l physils to form (Chichester et al., 1969). In the drier region of central and
eastern Washington the altered palagonite rinds on the Columbia River pillow
basalt contain nontronite (Summers, 1976).
Smectite is apparently the major physil in most of the soils of the midcontinent
temperate region of North America (Griffin, 1962;Forman and Bryon, 1965). Many
of the smectite-rich soils in the central and northern part are developed on till and
158

loess deposits and the smectite is largely inherited; however, studies of till profiles in
Indiana (Murray and Leininger, 1956; Droste et a/., 1960) indicate that both
chlorite and illite decrease upward as smectite increases, indicating some of the
smectite is pedogenic. In other profiles chlorite and illite alter only to vermiculite.
Smectite is the dominant physil, usually > 50%, in the major rivers draining this
area (Griffin, 1962; Potter et al., 1975). In the warm temperate humid south-central
region smectite is the dominant physil in the Houston series soils extending from
central Texas to Alabama. These soils commonly contain more than 60% < 2 pm,
with montmorillonite comprising 90% of the physil suite (Kunze and Templin,
1956). The montmorillonite is inherited from Upper Cretaceous and Lower Ceno-
zoic marls and chalks. There is little evidence of weathering in this low-relief
Coastal Plain region other than the removal of calcite.
Chemical analyses (Sawhney and Jackson, 1958) of the temperate zone soil
smectites afford the following structural formula:
Houston Black Clay, Texas:

( .34Fe&Mg0.28) (si3.73A10.u )OIO (OH)2'0.46


C.E.C. 115 meq/100 g
Peorian Loess, Wisconsin:

.?.4Fe%Mg0.29) (si3.93A10.07 )OlO (OH)2'0.37


C.E.C. 98 meq/100 g
The loess smectite is a fairly typical montmorillonite, whereas the Houston
material, with a relatively high tetrahedral charge, is a beidellitic montmorillonite.
Both have a relatively high Fe content.
Though smectites are abundant in midcontinent soils from central Canada to the
Gulf of Mexico, most of it is inherited. Some idea of the age of the smectite can be
obtained from K-Ar apparent ages obtained from the < 0.2 p m fraction of Pliocene
muds from the Mississippi Delta (Weaver and Wampler, 1970). Ages before burial
diagenesis are 160 f 60 m.y. The K-Ar data suggest much of the smectite in the
upper midcontinent soils was derived from Mesozoic sediments; this was a time of
extensive volcanism. To the west, off the coast of Texas, the <0.1 p m smectite
fractions from Middle Miocene clays have K-Ar ages of 55 k 3 m.y. (Aronson and
Hower, 1976). Subtracting 15 m.y. for age since deposition gives an age of 40 m.y.
(Eocene) for the detrital smectite. This is the approximate age of the smectite-rich
carbonates on the Texas Coastal Plain.
Relatively little soil smectite is present in eastern North America; it is being
formed rather than inherited. Many studies have shown that where chemical
weathering is reasonably mild or slow the following sequence can commonly be
observed:
mica-illite -j I/V -+ vermiculite -+ V/S --* smectite (beidellitic)
Chlorite produces a similar sequence:
chlorite -+ Ch/S -+ smectite
159

or
chlorite + Ch/swelling chlorite - Ch/V + vermiculite -+ smectite
The weathering sequence moves from left to right with increasing intensity of
weathering, increase in age, and from bottom to the top of a soil profile. Under
some conditions A1 and Fe hydroxides can precipitate in the interlayer space of the
expanded layers producing chloritic layers or soil chlorite.
Most young soil profiles developed on rocks containing an abundance of sheet
silicates contain a spectrum of expanded physils. Particularly in the cool temperate
regions of the eastern half of North America, smectite formed by the alteration of
chlorite and biotite, and to a lesser extent illite-muscovite, can be fairly abundant in
the A and B horizons (Droste et al., 1960; Whitting and Jackson, 1955; Brown and
Jackson, 1958; Malcolm et al., 1969; Forman and Brydon, 1965; Weaver, 1958;
Ross et al., 1983). In northeast Canada, in soils developed on a variety of rocks,
primarily Paleozoic sediments, illite commonly alters to smectite and to a lesser
extent vermiculite (McKague and Brydon, 1970).
In some instances (Brown and Jackson, 1958; Weaver, 1958; Egashira et al.,
1981) the alteration product of illite and muscovite have properties of both smectite
(expand to 17 A with ethylene glycol) and vermiculite (contract to 10 A when
treated with K). These boundary-line physils can be classed as high-charge smectite
or low-charge vermiculite. As they are mostly derived from illite-muscovite they
probably have a beidellite composition, high tetrahedral charge, which would favor
the fixation of K.
The formation of expanded layers from sheet silicates involves the removal of
interlayer K from micas and the interlayer hydroxyl sheets from chlorites. This
normally occurs under acid conditions and in the case of chlorite is favored by the
presence of organic acids which complex the A1 and Fe and aid in its removal
(Malcolm, 1969). The removal of interlayer material is commonly accompanied by
the oxidation of ferrous Fe. If the content of octahedral Fez+ is high, as in biotite,
oxidation can cause a decrease in layer charge and the formation of a smectite. If
the Fe3+is removed the charge can remain high and a vermiculitic material created.
For illites and muscovites that have a relatively low content of Fez+ the layer
charge is only slightly decreased during the early stage of weathering, and the initial
expanded layers are vermiculitic; with additional leaching, and/or the migration of
H + into the lattice the vermiculite can be altered to produce a complex mixture
and/or mixed-layer of vermiculite and smectite. Eventually a relatively high charged
montmorillonite-beidellite is produced. The relatively low tetrahedral charge of
pedogenic smectite suggests that in addition to altering the octahedral sheet there
has apparently been some silication of the tetrahedral sheet or the formation of new
tetrahedral sheets, which complex with hydroxy Al-Fe to create montmorillonite
(Jackson, 1965).
From the standpoint of source material, it is unlikely that smectite produced by
the weathering of layer silicates would be the dominant ( > 50-70%) physil in a
detrital physil suite. During erosion physils are normally removed from the whole
soil profile. If smectite is abundant in the A horizon, vermiculite, I/V, I/S, and
160

illite-micas are usually abundant in the B and C horizons. Some kaolinite may be
present in A and chlorite in C . A large portion of the physils in Pennsylvanian age
shales of the United States were presumably derived from immature soil profiles of
this type. The shales typically contain 50% illite and a mixture of I/S, smectite,
vermiculite, kaolinite, and chlorite. If weathering is extensive enough so that
smectite is produced in the lower horizons, the smectite in the upper horizon is
usually altered to kaolinite; however, under conditions of poor drainage thick
sections of smectite can be produced (Vertisol). This is more likely to occur when
volcanics are the principal parent rock.
Though the major constituents of the volcanic ash soils of Japan are allophane
and kaolin (Sudo and Shmoda, 1978), smectitic physils are relatively common
(Masui and Shoji, 1969). Montmorillonite is the dominant physil in the younger
soils. With increased weathering vermiculite increases, as does A1 interlayering.
Smectites of various types are major pedogenic physils in soils of Ireland and
Scotland (warm temperate with moderate rainfall). Basaltic material has been
weathered to produce saponite and mixed-layer swelling chlorite/saponite (Smith,
1959), montmorillonite with some vermiculite and illite layers (Bain and Russell,
1980), and a smectite intermediate between a dioctahedral and trioctahedral smec-
tite (Curtin and Smillie, 1981). The latter physil weathered from hydrothermal
saponite. The suggested formula is:
( .07 FeC%Fe,.':l Mg0.82)(si3.33 A10.67 )OIO (0H)2X0.40
Hornblende in a biotite-hornblende rock has altered to a mixed-layer swelling
chlorite/saponite (Wilson and Farmer, 1970); the plagioclase in granitic boulders
has altered to montmorillonite (Wilson et al., 1971); the chlorite in glacial till has
altered to chlorite/smectite and smectite.
Farther north, in Scandinavia, northern Russia, and northern Canada smectite
and mixed-layer physils are reported to be a major component of soils (Gradusov,
1974). As the expandable physils are more abundant in the A horizon than the C
horizon they presumably are pedogenic in origin.
In the subhumid Mediterranean region mineral degradation is moderate. Com-
monly illite is altered to I/S, rather than smectite, in areas where the rainfall is
between 50 and 100 cm. In lowland areas where the rainfall is between 30 and 50
cm, ions brought in during the humid season are concentrated during the dry season
and authigenic montmorillonite crystallizes. Where the rainfall is greater than 30 cm
palygorskite and sepiolite, present in many of the Cenozoic carbonates, alter to
montmorillonite (Paquet and Millot, 1972).
Montmorillonite is abundant in the semiarid region of the eastern Mediterranean
region, where it is inherited from carbonate rocks. In the subhumid region the
montmorillonite alters to kaolinite in the surface horizon of the well drained areas.
Basalt alters to montmorillonite in both the semiarid and subhumid environments
but in the latter can be partially altered to kaolinite (Yaalon et al., 1966).
Discussion
So, what does it mean when you discover that the physil suite of a shale has a
high smectite content? It can mean that the paleoclimate was cool, or arid, or warm
161

and humid with low relief, the soil in the source area was relatively young or had
low permeability, the source area contained smectitic clays or shales tens of millions
of years old, the source rock was volcanic, plutonic, metamorphic, or shale.
If the physil suite is more than - 70% smectite, the source rock was probably
volcanic. If the shale contains appreciable kaolinite, it suggests, assuming the ideal
case of one homogeneous source rock, that the source rock was volcanic and the soil
profile was relatively mature or there was considerable variation in relief, and/or
elevation, or rainfall in the source area. If the shale contains appreciable illite, I/S,
vermiculite, chlorite and kaolinite, or any three, the source rock was probably shale,
slate, and/or schist, and the soil was moderately young (or there was two source
areas, i.e., Mississippi River).
A high smectite content could indicate volcanic detritus was carried into the
depositional basin by water or wind and altered to smectite on the sea bottom or the
paleoclimate in the area including the depositional basin was alternately wet and
dry, probably warm, with smectite precipitating from solution during the dry
season, using ions that were taken into solution during the wet season.
Take your pick. Many parameters must be considered when speculating on the
geologic significance of physils.
Pedogenic soil smectites have a wide range of composition but in general have a
higher content of tetrahedral A1 and octahedral Fe than the typical bentonite
montmorillonite. They are slightly to very beidellitic and/or slightly to very
nontronitic. Chemically they should be fairly distinct, but they have compositions
similar to many marine hydrothermal smectites.

Vermiculite

Vermiculite occurs in variable amounts in all the major soil groups but is most
common in soils of temperate and subtropical climates. It is not a major component,
except locally. Soil vermiculites are almost always the alteration products of mica
and chlorite though Barshad and Kishk (1969), Barshad (1966), Ildefonse et al.
(1979), and others described vermiculitic physils that grew from solution. The
literature on the formation of vermiculite in soils has been reviewed by Norrish
(1972) and Douglas (1977).
Vermiculites can be either dioctahedral (derived from muscovite and illite) or
trioctahedral (derived from biotite, phlogopite, and chlorite) depending on the
composition of the source mineral. With continued weathering biotite can be altered
to the dioctahedral form, but the charge is usually lowered to the extent that it
becomes a smectite. In soils, dioctahedral vermiculite is more common than trioc-
tahedral vermiculite (Jackson, 1959) probably because of the relative stability of the
Al-rich muscovite and the abundance of dioctahedral illite.
Due to the nature of their origin “vermiculites” commonly contain some inter-
layers of the source physil and are mixed-layer Mi/V, I/V, Ch/V. Also, because of
the size of the original flakes vermiculitic material is commonly of sand and silt size
during the early stages of weathering. With increased weathering it fractures into
162

clay size material. With decreasing size the proportion of vermiculite layers de-
creases as hydroxy-Al layers develop in the interlayer region (Rich, 1968). This
process is discussed in the following section.
In the laboratory trioctahedral micas can easily be altered to vermiculite by
leaching with a solution of MgCl, (Barshad, 1948) or a variety of other cations,
particularly Ba2+ and Sr2+ (Norrish and Pickering, 1983). To remove the more
strongly bonded K from muscovite and illite, ions such as alkyl-ammonium or
sodium tetraphenylboron are required. The exchange is diffusion controlled and
proceeds inwards from the edge in an orderly manner. Very small amounts of K in
solution will stop the reaction. Thus, leaching must be efficient (Bassett, 1963).
The major factor controlling the rate of K exchange appears to be the proximity
of the proton of the octahedral (OH) to the interlayer K ion. Where the (OH) bond
is normal to the mica sheet, as in the trioctahedral micas, the repulsive forces allow
the K to be easily exchanged. Where the (OH) bond is inclined, due to the vacant
octahedral position in dioctahedral mica-illite, the K is more firmly bonded and
difficult to exchange (Bassett, 1960). As octahedral Fe and Mg are removed from
biotite, the layers become more dioctahedral in character and the K is more difficult
to remove.
During the conversion of biotite to vermiculite most of the Fe2+ is oxidized to
Fe3+,whch should significantly lower the layer charge; however, the layer charge is
not decreased to that extent. Under acid conditions some octahedral Fe and Mg are
removed, creating dioctahedral sites, and the high charge is maintained (Ismail,
1969). It is also suggested that concomitant with the oxidation of Fe there is a loss
of protons from octahedral hydroxyls, thus maintaining much of the layer charge.
Eventually there is some loss of Fe and Mg (Veith and Jackson, 1974):
[ (Fe2i)2Mg4(OH)4] tf [(Fe3+)2Mg404(OH)202]+ 2e- + 2H+ (1)
and
[(Fe2+)5Mg04(OH)4]-+ [(Fe3t)4(OH)4] + 5e- + Mg2+ (2)
Reaction (l), produced in the laboratory, caused no significant change in layer
charge; thus, a loss of interlayer charge is not necessary for the exchange of K.
Laboratory weathering of dioctahedral micas (Cook and Rich, 1963) indicates
that the resulting products can vary depending on the chemical used and the
particular mica used. Acid solutions cause A1 to be removed from the 2:l layer and
precipitated as hydroxyl groups in the interlayer space, thus limiting the increase in
C.E.C. Na citrate and molten Li nitrate are effective in removing K without
dissolving Al. Treatments with Li nitrate show that some muscovites produce
vermiculite (one layer glycerol) and others smectite (two layers glycerol). Though
there is little Fe2+in muscovite, there is a decrease in layer charge and C.E.C. when
it is converted to vermiculite. The initial K content of muscovite generally ranges
from 210 to 220 meq/100g. The charge on the expanded layers derived from this
material, in the laboratory, ranges from 156 to 184 meq/100g. The charge decrease
is believed to be due to the incorporation of protons in the 2 : l lattice (Leonard and
Weed, 1967).
163

During weathering micas commonly develop a regular mixed-layer I/V or I/S


phase. Norrish (1972) suggests that when a K + is replaced by a hydrated N a + the
site of the interlayer positive charge is located at a greater distance from the
octahedral hydroxyls. This allows the hydroxyl protons on either side of the
octahedral sheet to move towards the Na+ and slightly away from the K + on the
other side of the 2:l layer. This increases the bond strength of the K + and makes it
less likely to be exchanged than the K + from an unweathered layer. The resulting
product is a regular interstratification of mica and expanded layers. Calculations
(Giese, 1971) indicate that when K is removed from 2M muscovite, th OH band
angle with the cleavage plane increases from 18" to 53". However, random or zonal
development of expanded layers is also common.
The layer charge of both biotite and muscovite is near 1.0 per O,,(OH), and
that of vermiculites ranges from 0.6 to 0.9. For most samples there is some
reduction in layer charge. There is a considerable difference of opinion as to the
cause of the charge reduction. In general, for natural vermiculites, the reduction in
layer charge correlates with an increase in the amount of octahedral Fe3+ (Norrish,
1972); however, in some samples as much as 50% of the charge reduction cannot be
accounted for by Fe oxidation. Protons may be added or (OH) ions lost from the
silicate sheet. Laboratory results are conflicting. The reduction in layer charge of
vermiculitized muscovite and illite can be due to Fe oxidation, but the reduction
may be only apparent and reflect the deposition of hydroxy-A1 groups in the
interlayer space (Cook and Rich, 1963).
Based on laboratory experiments, Robert (1973) concluded that whether a mica
alters to vermiculite or smectite depends on both the total charge of the mica and
the amount of tetrahedral charge. Biotite and muscovite are difficult to convert to
smectite because even though the octahedral charge can be decreased the tetrahedral
charge is high enough to keep the product from expanding beyond 14-15 A
(vermiculite). To lower the charge to the smectite range, 0.2 to 0.6 per O,,(OH),,
the tetrahedral layer would have to be altered. On the other hand, lousy illites
(Fithian and Grundite), which are actually I/S and have a low total and low
tetrahedral charge, can be altered to smectite by oxidizing the small amount ( - 0.1
per O,,(OH),) of Fe2+ in the octahedral sheet.
Note also that, contrary to what one might expect, K is more difficult to remove
from both dioctahedral and trioctahedral micas in the fine, clay-sized fractions than
in the coarse fractions; also, less total K is exchanged in the fine fractions. This
presumably accounts for the stability of illites relative to coarser grained micas.
There is no adequate explanation for this phenomenon; but, as Norrish and
Pickering (1983) point out, in 1M - and presumably 1Md - micas and illites, the
OH dipole is parallel to the sheet structure and there is little interaction between the
H+ and K+. Therefore, the K + would be more strongly bonded than in 2M micas,
which are generally coarser. However, when 2M muscovite is ground to finer sizes
the same effect is observed.
During the process of mica weathering to vermiculite the expansions causes
disruption at both the edge and surface (Rich, 1964; Raman and Jackson, 1964). A
wedge-shaped pattern develops along the periphery of the flake when hydrated
164

MICA CLEAVAGE AT “FRAYED EDGE”

=K’ 0 * qo :. * YQZ’

~
4
I
1 AVERAGE: 390 mYg and 5.9% $0 c

(ILLUSTRATING SURFACE CONSTANCY)

cations and water replace K (Fig. 3-19). The layers expand to 14-15 A where
replacement occurs and tapers to 10 A towards the center of the flake where the K
has not been replaced. Small humps, blisters, and buckling occurs on the surface.
Rolling of the layers can occur along the flake margins.
In the natural system K is most commonly replaced by H + , Mg2+, Ca2+,A13+,
and probably organic cations. The Mg and A1 is often obtained by partial solution
of the mica lattice. The latter reaction can occur in distilled water (Serstevens et a/.,
1978).
Because of the high layer charge the cation exchange capacity of vermiculite is
considerably higher, - 140 to 200 meq/lOOg, than that of smectites. The “accepted”
average value for trioctahedral vermiculites is 159 meq/100g of which 154 meq is
from interlayer sites and 5 meq from external surfaces (Alexiades and Jackson,
1965). Dioctahedral vermiculites should have a higher exchange capacity; however,
they are more likely to have hydroxy-A1 blocking the exchange sites.
Because of the high layer charge vermiculites are able to “fix” K and collapse to
10 A. Fixation refers to the ability of K to replace the hydrated interlayer cations in
vermiculite and pull the 2:l layers together, forcing out the interlayer water as well
as the cation. The 14-15 A layer contracts to 10 A and the K is bound or fixed in
the interlayer space so that it cannot be replaced by NH:’. Fixation and contrac-
tion are not always directly related. Some fixation can occur without appreciable
contraction (x-ray), and contraction can occur where the K is not well fixed
(Weaver, 1958). But in general, fixation leads to contraction and vice versa.
Due to the decrease in layer charge, the presence of organics and hydroxy
material in the interlayer space, and the physical disruption of the flakes, K release
and K fixation are not fully reversible. Vermiculites have a selectivity for K, but it
varies with the extent of weathering, pH, K concentration, and counter-ion con-
centration (Rich, 1964).
165

K fixation is of particular importance to the soil scientist as the availability of K


to plants is determined by how strongly the K is bonded. The geologist's primary
concern is whether a particular vermiculite is capable of extracting K form sea water
and reverting to a 10 A physil. This is possible (Weaver, 1958), but its volumetric
significance has not been determined. However, eastern rivers draining the Appa-
lachians commonly contain vermiculite but it is usually not observed in the
estuarine physils.
Most any x-ray pattern of a mildly weathered shale will confirm that the chlorite
has altered to a mixed-layer Ch/V or Ch/S. Presumably the hydroxy-Mg or brucite
sheet has been preferentially removed from between some of the 2:l layers. This
weathering process appears to be more involved than removing K from mica.
Acid leaching experiments (Ross, 1969) indicate there is no preferential dissolu-
tion of the hydroxide sheet in true chlorites. Makumbi and Herbillon (1972) were
able to produce a regular mixed-layer Ch/V from chlorite by treating it with Na
dithionite or Na sulfite solution. The process involved oxidizing the Fe and
removing it. Ross and Kodama (1974)produced complete vermiculite from chlorite
by first heating the chlorite (sheridanite) to 610°C to dehydroxylate the hydroxide
sheet and oxidize the Fe2+.The sample was then shaken in a solution of 0.2 NHCl
and 0.2 NNaCl to dissolve the dehydroxylated hydroxide sheet. The composition of
the octahedral cation suite changes as follows:

(A1,,,Fei.+,Fe~.:,Mg4.,2) (A10.32Fei,12Mg2.49)
Chlorite Vermiculite
There was little change in the tetrahedral sheet. They concluded that the hydroxide
sheet must be structurally disturbed before it can be selectively removed and that
this could occur during metamorphism. Oxidation of Fe2+is also believed to play a
major role in initiating structural disorder.
Later they (Ross and Kodama, 1976) were able to produce vermiculite from
diabantite (IIb, high Fe chlorite) but only regular Ch/V from brunsvigite (IIb,
intermediate Fe chlorite) when the Fe2+was oxidized by treating the chlorites for
four months in saturated bromine water on a steambath. A similar treatment of
sheridanite (low Fe) produced no significant amount of vermiculite layers. It is
apparent that the oxidation of octahedral Fe2+ is an important process in the
transformation of chlorite to vermiculite; the amount and distribution of the Fe will
influence the amount, distribution, and type of expanded layers produced.
Actually the regularly interstratified Ch/V produced by Ross and Kodama
(1976) was Ch/S. The expanded layer easily expanded to approximately 17.7 A
when the sample was treated with glycerol, indicating the layers are more like
smectite than vermiculite. Most of the regular interlayered Ch/V described in the
literature is Ch/S. The natural weathering product of the chlorite (in metabasalt)
used by Ross and Kodama was a true Ch/V, i.e., no expansion with glycerol (glycol
would have been a better test). It is likely that, in the natural weathering conditions,
cations were removed from the octahedral sheet of the 2:1 layer, preserving much of
the negative tetrahedral charge.
166

Hayes (1970) has suggested that the weathering sequence is:


IIb chlorite + regular Ch/V + random Ch/V -+ Ia vermiculite
However, laboratory experiments indicate this sequence is idealized and the se-
quence is influenced by mineral composition; i.e., for high Fe chlorite it is possible
to go directly to vermiculite.
Herbillon and Makumbi (1975) suggest that alternate layers in some IIB chlorites
may have different compositions (high A1 vs. high Mg) which would favor the
development of regular Ch/S or Ch/V.
Commonly the weathering evolution of vermiculite is towards an intergraded
material or a smectite. Vermiculites derived from biotite and phlogopite tend to
alter to nontronitic or saponitic smectites, whereas those derived from muscovite
and illite tend to form a beidellitic smectite.

Occurrence of Trioctahedral Vermiculite


Though vermiculite is commonly reported as a soil component, it is usually a
minor phase and not studied in detail. Douglas (1977) lists references for 31 papers
describing the occurrence of soil vermiculite. It has been found in most areas of
North America and in other countries, ranging from Australia to Japan.
It would be safe to say that vermiculite can be found anywhere biotite is found,
though it is more likely to be preserved in cold and temperate than tropical climates.
Biotite (and phlogopite) has been observed to weather to vermiculite in Antarctica
(Wilson and Nadeau, 1984) and in the alpine zone of the northern Cascades,
Washington (Reynalds, 1971). Vermiculite and/or vermiculitic mixed-layer physils
derived from both biotite and chlorite are common components of glacial deposits
where the soils are relatively young (Whittig and Jackson, 1955; Fry et al., 1960;
Borchardt et al., 1966; Mitchell, 1955, Scotland; Gjems, 1969, Norway).
Weathering profiles in which vermiculitic material is dominant develop on
biotite-rich metamorphic rocks (Wilson, 1970; Eroschev-Shak, 1970). Wilson (1970)
observed the following sequence (in the > 75 p m fraction):
Biotite -+ B/V (1:l) + Ch/V + Vermiculite
C horizon B horizon
In the B horizon Fe and Mg hydroxy sheets developed between about half the
layers to form chlorite layers. Nearer the surface these hydroxy sheets were leached
out, particularly in the finer sizes.
Whereas mixed-layer Mi/V and vermiculite are present in sand and silt fractions,
smectite may be the dominant physil in the clay fraction. Jackson and Sridhar
(1974) describe a weathered phlogopite deposit from Kansas which had vermiculite
in the coarse fraction and saponite in the clay fraction.
Chlorite weathers as readily as biotite and when both are present as in some
glacial tills and loess, both will alter toward vermiculite. More often chlorite occurs
with illite (shales) or muscovite-phengite (greenschist) and forms a trioctahedral
vermiculite before there is appreciable weathering of the more chemically stable
dioctahedral micas.
167

Chlorites are relatively abundant in slates. The chlorite in the Mesozoic slates of
the western foothills in the Sierra Nevada (central California into Oregon) alters
early to a regular mixed-layer Ch/V (Post and Janke, 1974). I have observed a
similar alteration of the Paleozoic slates of the Appalachan region. In both
localities the hydration occurs as the slate changes in color from gray to light brown
but remains a rock. Further, alteration in the soil zone leads to the formation of
vermiculite and saponite.
Johnson (1964) found that the chlorite (silt-size) in a metamorphosed basalt from
eastern Pennsylvania altered, during weathering, to a regular mixed-layer Ch/V.
Based on a study of New Zealand soils, Churchman (1980) suggested that with time,
increasing precipitation, and/or increasing temperature the weathering of chlorite
proceeded as follows:
chlorite + interlayered hydrous micas + chlorite/swelling chlorite
+ Ch/V -+ decomposition

He found no discrete vermiculite and suggested Ch/V was destroyed rather than
altered to vermiculite. The nature and origin of the hydrous mica phase (I/S) is
puzzling, but I have noticed a similar weathering phase. Apparently in some
situations K is released from associated mica-illite and fixed in the early formed
vermiculite layers.
In the humid tropical region of Zaire (4-month dry season), the weathering
sequence is only slightly more advanced than in the temperate regions (Herbillon
and Makumbi, 1975). Soil profiles developed on chlorite schist show that chlorite, in
the sand fraction, undergoes the following weathering reaction:
chlorite --f Ch/V (regular 1:l)
In the silt and clay fraction the sequence is:
chlorite + Ch/V (regular 1:1) + Ch/V or Ch/S (irregular)
+ nontronite + kaolinite (and K/S?)
The first stage of alteration involves the oxidation of Fe2+ and the removal of some
Fe and Mg. This is accompanied by a loss of some tetrahedral A1 as dioctahedral
nontronite forms. Near the surface this A1 may be deposited in the interlayer space
to form K/S.
It should be noted that in some soils chlorite simply disappears and is presuma-
bly decomposed to an amorphous phase rather than transformed into another
physil. This is believed to be due to acid conditions (McKeague and Brydon, 1970),
and fulvic acid (Bain, 1977).
Though soil vermiculite and mixed-layer vermiculitic physils are largely formed
by the limited alteration of layer silicates, numerous studies indicate these physils
can be neoformed. Vermiculite books up to 150 pm in size are present in a soil
develop on a metagabbro (Loire-Atlantique, France) (Ildefonse et al., 1979). No
layer minerals are present and vermiculite crystallized from ions derived largely
from the dissolution of actinolite. The “ vermiculite” expands to about 16 A when
treated with ethylene glycol, has a relatively low octahedral occupancy (2.5-2.7),
168

low octahedral Mg, and low layer charge (0.6-0.7). Near the surface the material
weathers to an Fe-rich smectite. A regular mixed-layer Ch/V formed from the
weathering of a similar rock in Maryland (Rabenhorst et al., 1982). Smectite and
random Ch/V occurrs in the < 0.2 p m fraction.
Both di- and trioctahedral vermiculite have been found as a weathering product
of hornblende (Massif Central, France) (Proust, 1981):

Trioctahedral

Dioctahedral

Occurrence of Dioctuhedrul Vermiculite


In soils dioctahedral vermiculite is more common than the trioctahedral variety
(Jackson, 1959). In part this reflects the relative stability of muscovite and biotite,
but to a large extent it reflects the relative abundance of illite as a source mineral.
Under the acid weathering conditions in which much of the dioctahedral vermiculite
is formed, A1 is commonly released from the 2:l lattice and deposited as Al-hydroxy
groups in the interlayer position to produce a chloritic layer. This material is
discussed in the next section.
Prouse et al. (1986)described a process by which trioctahedral Mg chlorite was
converted to dioctahedral vermiculite (schist, Limoges, France). The weathering
sequence is: chlorite + ordered Ch/V + vermiculite. Mg was preferentially ex-
tracted from the chlorite hydroxide sheet and FeZC,along with Mg from the 2 : l
layer. The Fe3+ content remained unchanged. Tetrahedral A1 decreased and oc-
tahedral A1 increased as the Mg-rich trioctahedral 2:l layer was converted to an
Al-rich dioctahedral sheet.
Dioctahedral vermiculite and/or M/V are abundant and can be the major physil
throughout the Appalachian Region from Canada to Alabama, in soils developed on
muscovite schist (Fig. 3-20).They are also abundant as detrital material in many of
the Atlantic Coastal Plain soils (Hathaway, 1955; Rich, 1956; Rich and Cook
(1963). The physil suite in most of the river draining the Appalachian Mountains
contains a detectable amount of dioctahedral vermiculite. This material is not found
in the coastal marine sediments and has presumably adsorbed K and reverted to the
mica form. As a relatively high weathering intensity is required to form dioctahedral
vermiculite-rich soils, such soils commonly contain appreciable kaolinite.
Dioctahedral vermiculite occurs throughout the temperate region (largely in
Podzols); only a few references will be listed: central United States (Jackson, 1959);
169

100;
1 4 6 A Depth.
inches
3-8

8-13

13-32

44-56

~ 20 5,M-
Fig. 3-20. Weathering sequence of muscovite in silt loam, Fluvanna Co.,Virginia. Muscovite (10 A) is
increasingly converted to vermiculite (14 A) with decreasing depth and decreasing size. Intermediate
peaks are M/V. From Rich, 1956. Copyright 1956 The Clay Miner. SOC.

western United States (Barshad and Kishk, 1969); Canada (Kodama and Brydon,
1968); England (Brown, 1953); Europe (Millot, 1970; Paquet and Millot, 1972);
Lebanon (Paquet, 1970); New Zealand (Churchman, 1980); Australia (Jones et af.,
1964).
As in the formation of trioctahedral vermiculite, it is likely that muscovite-illite
weathers to I/V in the coarse fraction and I/S in the fine fraciton.

Discussion
“True” vermiculite (collapse to near 10 A when exposed to K) and mixed-layer
Ch/V, B/V, M/V, and I/V are common components of soils and alluvium but are
relatively rare in shales. Vermiculite and mixed-layer mica/vermiculites are rare in
marine sediments, presumably having fixed K and reverted to mica-illite or I/S, but
do occur in continental physillites and sandstones (Weaver, 1958). It is possible
some vermiculitic layers are created by acid leaching during burial. Ch/V is
reported to be a common component of shales, carbonates, volcanogenic sand-
stones, and evaporites. In the vast majority of cases the physil is Ch/S rather than
Ch/V. In other instances vermiculite physils are found in outcrop samples where
they are commonly a product of recent weathering.
Nevertheless, detrital Ch/S and Ch/V are relatively common in sediments,
particularly the former. Soil investigations indicates that much of it was probably
formed from the weathering of chlorites; however, an appreciable amount probably
170

formed by the weathering and low-temperature hydrothermal alteration of basic


igneous rocks. The presence of detrital Ch/S and Ch/V is normally an indication of
mild climatic conditions in the source area.
Note that most of the Ch/S in carbonate, evaporite, and volcanoclastic rocks is
either authigenic or diagenetic.
Whereas I/V is rare in marine rocks, I/S is relatively abundant. Much of the I/S
is detrital, derived by the weathering of illite. The layer charge of illites is so low
that only a minor reduction in charge, perhaps only the oxidation of Fe, is necessary
to produce smectitic layers rather than vermiculitic layers. Soil studies indicate that
more charge reduction occurs in the fine-grained than coarse-grained fraction of
weathered mica-illites. The I/S in sediments is normally fine-grained (commonly
< 0.2 pm). I/V may be relatively abundant in the coarser fraction ( > 2 pm) that is
seldom analysed.

Hydroxy Interlayered Vermiculite and Smectite

As discussed briefly in the sections on smectite and vermiculite, hydroxy forms of


A1 and Fe are commonly precipitated in the interlayer space during acid weathering.
Hydroxy-A1 is by far the more abundant complex. Under evaporitic conditions
hydroxy-Mg is a common interlayer precipitate. Depending on pH, availability of
complexing cations, and interlayer charge, the hydroxy complexes can occur as
scattered islands or form almost complete sheets, soil chlorites. Among other things,
this interlayer material reduces the CEC and hinders the contraction of the
interlayer space when heated.
The physils with interlayer A1 have been given a variety of names. A few are
intergradient, intergrade, hydroxy-Al, aluminum interlayers, chloritized expansible
layer silicate, chlorite-vermiculite-intergrade, etc. (see Barnhisel, 1977). I like the
terms hydroxy-A1 or hydroxy interlayered. The hydroxy interlayered physils are best
identified by their response to heat treatment, preferably K-saturated samples. The
starting vermiculite, usually dioctahedral, has a strong (OOl), 14 A peak. As inter-
layer water is replaced by hydroxy-A1 (Fe or Mg). The 14 A peak intensity decreases
relative to the other (001) peaks. As the area covered by hdyroxy-A1 increases, the
(001) relative intensities more and more resemble those of chlorite (Fig 2-29). The
initial hydroxy-A1 units collapse to 10 A when the sample is heated at 110°C. As the
amount of hydroxy-A1 in the interlayer space increases, higher and higher tempera-
tures are required to cause contraction in the interlayer space When only a minor
amount of hydroxy-A1 is present the 10 A peak will shift to the 10 to 11 A range
when the sample is heated at 500-600°C (Fig. 3-21). When the interlayer space is
nearly filled with hydroxy-Al, the 14 A peak will shift only slightly and at high
temperatures increase in intensity, as does chlorite. Step heating and x-ray analyses
provide a qualitive measure of the amount of hydroxy interlayer material, but a
quantitative technique has not been established. A note of caution: organic matter
in the interlayer space can simulate the effects of small to moderate amounts of
171

K 25

K 110
K 300

K 550

Fig. 3-21. Soil vermiculite containing some hydroxy interlayers. The collapse of the vermiculite to around
11A when heated at 550°C indicates the hydroxy interlayers are not abundant. From Barnhisel, 1977.
Reprinted by permission of Soil Sci. Soc. Amer., Inc.

hydroxy-Al. DTA analyses can be useful in distinguishing between the two types of
interlayer material.
The laboratory synthesis of hydroxy interlayered vermiculite and smectite has
been reviewed by Rich (1968), Barnhisel (1977), Lahav et al. (1978), and Brindley
and Kao (1980). There are three general procedures: (1) a solution of the desired
hydroxy-metal cation ratio is prepared by adding a base, NaOH, to a salt, AlCl,,
MgCl,, etc., and added to a suspension of Na-montmorillonite or vermiculite; (2)
the base and salt solution are added simultaneously, drop-wise, to an agitated,
dispersed physil; (3) the salt solution is added first to the dispersed physil, and the
system then titrated with a base.
Ion hydrolysis and the formation of complex ions in the presence of physils is not
well understood. The stability upon aging of hydroxy interlayers is a function of
OH/cation molar ratio, time of aging, method of preparation, temperature, particle
size, and the character of the physil (Barnhisel, 1977). On aging, gibbsite may form
externally. This occurs within a few weeks when the OH/A1 molar ratio is 3.0, in
three months when the ratio is 2.25, and over six months when the ratio is less than
1.50 (Barnhsel and Rich, 1963). In other experiments different results were ob-
tained and no gibbsite was formed (Carstea et al., 1970). Vermiculites bond the
hydroxy-A1 strongly and prevent the formation of gibbsite (Brydon and Turner,
1972).
For hydroxy complexes to be bonded and remain in the interlayer position it is
necessary that they have a positive charge, therefore a OH/A1 ratio < 3.0, e.g.,
[Al(OH),]+, [Al(OH)]*+. It has been suggested that the Al-complex occurs as a
six-membered ring structure, AI,(OH):: (Hsu and Rich, 1960; Jackson, 1960). This
is probably the smallest A1 polymer that exists in the interlayer space that is not
172

subject to cation exchange. Larger polymers have been proposed by Jackson (1962)
and Hsu and Bates (1964). The OH-A1 polymers continue to hydrolyze and
polymerize into larger units during prolonged aging (Hsu, 1977), e.g.
[A154(OH),,]’8+. 36H,O. The various experiments suggest that “fixed” or nonex-
changeable A1 occurs in the narrow OH/A1 range of about 2.5 to 2.7. Where the
ratio is higher or lower than the ideal value a portion of the polymer may have a
ratio between 2.5 and 2.7 (Barnhisel, 1977). Analyses of natural samples indicates
interlayer A1 ranges from 0.90 to 1.84 with an OH/A1 ratio of 2.3 to 3.0 (average
2.61) (Kirland and Hajek, 1972).
Calculations indicate that a positive charge on the hydroxyl sheet is not necessary
for it to bond to a neutral 2:l layer. Long H bonds are capable of creating a stable,
though weak, structure (Bish and Giese, 1981).
Table 3-1 and illustrate the systematic development of hydroxy-A1 and hydroxy-
Mg interlayer sheets between montmorillonite layers (Brindley and Kao, 1980).
Units of NaOH (1 through 6) were systematically added to A1 and Mg saturated
Wyoming montmorillonite. The maximum addition of NaOH gave an OH/Al = 3
and OH/Mg = 2, enough to convert all A1 to a gibbsite-like arrangement and all
Mg to a brucite-like arrangement. Assuming the composition of the 2:l layer
remains constant the composition of the interlayer sheet was calculated. Note the
systematic increase in interlayer A1 and Mg, and in the OH/Al, OH/Mg ratios.
Interlayer H,O systematically decreases. Prior to the addition of NaOH the OH/A1
ratio was 2.17, corresponding to the formula [Al(OH),]’ and the OH/Mg ratio was
1.03, corresponding to the formula [Mg(OH)]+.
X-ray diffraction patterns show that the hydroxy-Mg products have less tendency
to swell in ethylene glycol and water, and greater thermal stability than hydroxy-A1
products. The latter observation reflects the greater thermal stability of Mg(OH)
compared with Al(OH),. Hydroxy-Al-montmorillonite expanded to 17 to 22 A
when treated with ethylene glycol at all stages, whereas hydroxy-Mg-montmoril-
lonite does not expand when the OH/Mg ratio is > 1.8. The (001) spacings for the
hydroxy-Mg-montmorillonite with nearly complete interlayer sheets are in the range
A.
of 14.6 to 14.9 A, and for the A1 variety the value is about 14.5 These values are
higher than for normal chlorites but are appropriate for the low layer charge of the
montmorillonite layers. The (001) value of chlorites (14.2 to 14.4 A) decreases as the
2:l layer charge increases.
A number of chemical techniques have been devised to preferentially dissolve
hydroxy interlayers. These are listed by Barnhisel (1977). Once these interlayers are
removed the swelling and exchange properties of the 2:l physil are normally
restored.
Much of the laboratory data concerning hydroxy interlayers are not directly
applicable to the natural system. Much of the soluble A1 in soils is complexed with
organic acids, and analyses of the hydroxy interlayers indicate they have a complex
composition. Experiments by Goh and Huang (1986) indicate hydroxy-Al-citrate
complexes can be adsorbed as interlayers in montmorillonite. An appreciable
amount of Fe is usually present in the hydroxy interlayers (Carstea et al., 1970) and
under alkaline soil conditions Mg may be present. The parent material in soils is
173

Table 3-2
Chemical analyses and interlayer formulae of Na-montmorilloni te and of hydroxy-Al-montmorillonites
and hydroxy-Mg-montmorillonite (0,..., 6) following the systematic addition of NaOH. After Brindley
and Kao (1980).
Na-mont-
morillonite 0 1 2 3 4 5 6
SiO, 58.4 52.4 50.6 48.7 46.2 44.3 41.4 37.5
A1 2 0 3 20.30 22.32 25.87 29.68 31.59 33.23 34.45 38.54
Fe20, 3.88 3.62 3.36 3.20 2.94 2.88 2.82 2.52
MgO 2.60 2.25 2.13 2.09 1.94 1.90 1.79 1.68
CaO 0.62 0 0.12 0.12 0.13 0.13 0.45 0.46
Na20 2.34 0 0 0 0 0 0 0.83
K2O 0.17 0.07 0.15 0.06 0.13 0.07 0.07 0.12
H 2 0 z 110°C 6.17 8.87 10.85 10.98 12.75 12.35 13.70 15.50
H,O <llO°C 4.74 8.49 5.96 4.93 3.46 3.74 3.49 3.14
Total 99.22 98.04 99.04 99.76 99.14 98.60 98.17 100.29
Interlayer composrrion
Na 0.30 0 0 0 0 0 0 0.17
K 0.01 0.01 0.01 0.01 0.01 0.00 0.01 0.02
Ca 0.04 0.00 0.01 0.01 0.01 0.01 0.04 0.05
Al 0.0 0.35 0.75 1.19 1.54 1.84 2.23 * 3.15 *
OH 0.0s 0.76 1.94 3.30 4.32 5.26 6.48 8.84
H2 0 0.35 0.84 0.84 0.31 0.46 0.03 0.10 -
0 = 0.28
OH/AI - 2.17 2.59 2.77 2.80 2.86 2.90 2.98
+
(OH H20)/AI - 4.57 3.71 3.03 3.10 2.88 2.95 -

SiO, 58.3 53.1 50.0 47.6 44.6 42.5 41.1 39.1


‘41 2 0 3 20.21 18.43 17.31 16.43 15.55 14.71 14.26 13.63
Fe203 3.68 3.42 3.28 3.02 2.78 2.66 2.42 2.41
MgO 2.79 5.25 8.82 13.16 15.98 20.19 22.86 25.26
CaO 0.50 - - - - - - -
Na20 2.77 0.26 0.17 0.15 0.11 0.11 0.06 0.36
K2 0 tr tr tr tr tr tr tr tr
H 2 0 z 110°C 6.16 7.85 8.58 10.04 11.79 13.00 14.26 14.65
H 2 0 < 110°C 4.68 11.31 10.33 8.14 7.20 6.00 4.41 4.09
Total 99.09 99.62 98.49 98.54 98.01 99.17 99.37 99.50
Interlayer composirion
Ca 0.04 - - - - - - -
Na 0.36 0.04 0.03 0.02 0.02 0.02 0.01 0.07
Mg - 0.31 0.75 1.34 1.82 2.50 * 2.97 * 3.50 *
OH 0.10 0.32 1018 2.36 3.32 4.68 5.60 6.72
H2 0 0.34 0.78 0.66 0.58 0.80 0.67 0.75 0.55
OH/Mg - 1.03 1.57 1.76 1.82 1.87 1.89 1.92
(OH + H,O)/Mg - 3.55 2.45 2.19 2.26 2.14 2.14 2.08
Extraneous gibbsite or brucite.

commonly weathered muscovite-illite rather than montmorillonite or trioctahedral


vermiculite, which are used in laboratory experiments.
Hydroxy interlayered physils are relatively common in soils (Jackson, 1963),
particularly in the temperate regions. In 1968 Rich listed 34 occurrences of pedo-
174

genic interlayered physils. Many more have been reported since 1968. In addition to
forming from micas, these chloritic physils form from smectites developed on
basaltic rocks (Hawaii and Australia) (Jackson, 1959) and volcanic ash (Japan)
(Masui and Shoju, 1969). The latter authors note that during the early stages of
weathering montmorillonite with only a minor amount of interlayer hydroxy material
is predominant. With increased weathering, vermiculite increases and hydroxy-A1 is
enhanced. Hydroxy-A1 is more strongly fixed in the surface horizon than in lower
horizons. Mermut et al. (1987) described a soil in Canada in which complete layers
of hydroxy Fe were formed in vermiculite derived from biotite.
Rich (1968) concluded that the optimum conditions for the development of
hydroxy interlayers were: moderately active weathering, pH 4.6 to 5.8, low organic
content, and frequent wetting and drying of the soil. The accumulation of interlayer
hydroxy-Al is more pronounced under well drained conditions than under poorly
drained conditions (grumusols) (Jackson, 1963; Huang and Lee, 1969). Once formed,
hydroxy-Al interlayered vermiculites are quite stable and are commonly associated
with kaolinite, with the interlayered material being in the A horizon and kaolinite in
the B and C horizons (Douglas, 1977); though, under conditions of low pH ( < 4.5)
and high organic content, A1 has been solubilized and complexed by the organic
matter, converting hydroxy interlayered physils to montmorillonite (Malcolm et al.
1969). A similar type of alteration was observed in acid (pH 3.0-3.5), oxidized tidal
flat sediments from San Pablo Bay, California (Lynn and Whittig, 1966).
Hydroxy-A1 interlayered physils have been found in fluviatile deposits (Bossi,
1972; Schultz, 1963) but appear to be rare in marine sediments. Chloritic material of
this type is present in many of the river and estuarine sediments of eastern and
southern United States but is not found in the marine muds. It is not clear whether
the hydroxy-A1 (and Fe) is dissolved in the marine environment (the organic content
is generally high) or the conversion to chlorite is completed by the precipitation of
hydroxy-Mg.
During burial and increased temperature, one might expect that the hydroxy-A1
interlayer physils would convert completely to chlorite. However, the chlorite would
be dioctahedral and little of that is found in sedimentary rock, though it has been
reported in Triassic fluviatile sediments of the Colorado Plateau (Schultz, 1963). It
is quite possible that minor amounts of dioctahedral chloritic material or chlorite
would go undetected. Weaver and Beck (1971) suggested that dioctahedral chlorite
occurs interlayered with diagenetic I/S, actually I/Ch/S.

Mica Illite

Micaceous material of one sort or another is present in nearly all soils. It may
occur as trace amounts in lateritic soils or be the dominant physil in temperate soils.
The clay fraction of the C horizon of most of the soils in the Northern Hemisphere
contains > 30% illite, and about three-fourths of these soils contain > 50% illite
(hydromica). In the Southern Hemisphere high illite is restricted primarily to
mountainous areas (Andes, South Africa, SE Australia) (Gradusov, 1974). Gradusov
175

L I In --A

Fig. 3-22. Sketch map of the world showing areas of soils with different proportions of illitic material in
clay fractions of their A, and/or A, horizons plus major areas of some other soils, as follows:
1. Soils with hydromicas 1: 70%or more; 2. 50-70%; 3. 30-508; 4: 30%or less.
11. Soils formed in ashfalls 1. humid areas; 2. and areas.
111. Soils of the Amazon basin.
IV. Soils with palygorskite.
From Graduson, 1974. Copyright 1974 Elsevier.

considers the composition of the C horizon reflects the composition of the parent
material. This is reasonably true in the temperate and cold climates but not in the
tropics. The absence of illite in the C horizon of the tropical soils is presumably due
to weathering intensity rather than its absence in the parent rock.
The distribution and concentration of illite in the A horizon (Fig. 3-22) is similar
to that in the C horizon though in the polar regions the values are much lower
( < 30%vs. > 70%). The distribution data suggests much of the micaceous material
in soils is inherited, and even though some is altered it probably persists in the A
horizon as degraded illite. Only in Italy is it indicated that there is more micaceous
material in the A than in the C horizon.
As summarized by Reichenbach and Rich (1975) micas, illites, and degraded
forms are predominant in the clay fraction of arctic raw soils, brown earths, prairie
soils, chernozems, chestnut soils, syrozem, alkali soils, intrazonal mountain soils,
and different azonal soils. Some gray-brown, gray, red, and red-yellow podzolic soils
have “remarkable contents of illite”. (Or, they are prevalent in Entisols, Inceptisols,
Mollisols, Aridisols, and Alfisols (Fanning and Kerarnides, 1977).)
As in shales and other sedimentary rock, the micaceous component of soils
176

commonly consists of at least two different minerals, muscovite and illite. Biotite is
more abundant than muscovite in igneous and metamorphic rock, but because of
the ease with which it is weathered it is found only in very young soils. Further,
during weathering under reducing conditions, Mg and Fe can be removed from the
octahedral layer and a dioctahedral mica formed. T h s transformation occurred in a
kaolinite crust formed on granite. Presumably in local areas the feldspars caused
pore solutions to have relatively high concentrations of K, Si, and A1 and low
concentrations of Fe and Mg (Konta et al., 1970).
In general muscovite (and presumably phengite) is most abundant in the silt
fraction and illite and I/S (with a high proportion of illite layers) in the clay
fraction. Both may be inherited; the illite may have formed by degradation of the
inherited muscovite; the illite and I/S may have crystallized in the soil. To a large
extent the grain size of micaceous material in soils reflects the size of the material in
the parent rock. Limited data indicate the course material is usually the 2M
polytype and the fine illite the 1M-1Md polytypes, as in shales.
When we consider soils as a source of detrital illite, it is apparent that in Recent
time most of the soil derived illite transported by rivers is inherited illite. Actually,
studies of the Mississippi River sediment (p. 197) suggest most of the illite was
obtained from the bedrock, predominantly shales and metamorphic rocks. At the
present time rivers carrying a physil suite with a high illite content are draining
areas where illite is abundant in the parent rock. However, this may not have been
true in the past. Conditions favorable for the development of pedogenic illite may
have been extensive in the Precambrian and Paleozoic, and even during younger
periods. It is a distinct possibility that much of the illite and I/S in ancient shales
was formed in soils.
What needs to be considered here is the formation of pedogenic illite in modern
soils and the modification of inherited illites and muscovites during soil weathering.
It is difficult to distinguish between the two end products.
Muscovite and illite are relatively resistant to weathering, and relatively intense
weathering is required before they are appreciably altered. As discussed in the
previous sections, during chemical weathering the octahedral sheet may be modified,
but regardless, K is stripped from the interlayer space to produce expandable,
water-bearing layers. When the proportion of expandable layers is less than about
10 to 20%, the material is usually called a degraded mica or illite (or hydromica).
-
When the expanded layers exceed 20%, the material is normally referred to as a
mixed-layer physil; when expanded layers reach on the order of 70%, the physils are
called (incorrectly) smectites or vermiculites.
Fig. 3-23 shows x-ray patterns of various size fractions of the B2 horizon of a soil
developed on glacial material in Wisconsin (Fanning and Keramidas, 1977). Note
the general broadening of the 10 A peak with decreasing size. The broadening is due
both to a decrease in flake size and the development of some expanded layers
(degraded illite). The chlorite is converted to smectite-vermiculite complex in the
finer sizes. On the basis of the x-ray patterns it cannot be concluded that the fine
10 A material is a degradation product of the coarse 10 A material. A similar set of
A
10 peaks can be obtained from various size fractions of many unweathered shales.
177

FINE CLAY

Fig. 3-23. X-ray patterns of various size fractions of Varna soilB2 horizon, Wisconsin. Mg, 25°C and K,
25°C were glycerol solvated. Note progressive widening of 10 A illite peak with decreasing particle size.
From Fanning and Keramidas, 1977. Reprinted by permission of Soil Sci. SOC.Amer., Inc.

However, in most shales the chlorite peak would be well developed in all size
fractions.
Though micas and illites break into finer sizes when they are weathered to
expandable phases, it is not clear whether there is any size reduction without major
chemical modification. It is likely that some are weakened enough by mild chemical
changes that they break during erosion and transport.
The most likely source mineral for the formation of pedogenic illite or mica is
K-feldspar, or plagioclase plus mica. Si and A1 are available in most environments;
K is the essential ingredient for the formation of illite. Divis and McKenzie (1975)
were able to produce illite during hydrothermal experiments where they heated
orthoclase in brines and distilled water at 200" and 300°C for 29 to 32 days. The
illite produced was a 1Md variety with a 10.2 A spacing. When heated in con-
centrated brines a dense surface coating of illite was obtained. In dilute brines and
distilled water scattered illite rosettes developed. During crystallization, illite ap-
parently uses portions of orthoclase as structural components. Much of the original
structure, chains of tetrahedra, is apparently retained; but as hydration occurs, some
Si and K is lost. When quartz, albite, and calcite were present, scattered rosettes
formed on them. This indicates the illite was also able to grow directly from
solution.
Various Russian studies suggest microorganisms, by concentrating K, may play a
role in the formation of illite from feldspar (Parfenova and Yaribova, 1962).
Harder (1974) was apparently able to synthesize some poorly crystallized I/S at
surface temperatures (3" to 20°C) by the precipitation of A1 hydroxides from Si-,
178

Mg-, and K-containing solutions and aging for several months. The formation was
favored by a low Si concentration, Al/Si and Al/Mg ratios similar to those in
natural illites (controlled by initial composition and pH), a very high K content, and
pH 8.5 to 9. NaCl in the solution tends to inhibit the formation of I/S. Harder
concluded that it was not possible to form illite under isochemical weathering
conditions from K-feldspar because the K is not high enough in relation to the
1A1/3Si ratio. K must be strongly enriched (Harder, 1974) or the Al/Si ratio
modified.
High resolution imaging by transmission electron microscopy of weathered
microcline (granodiorite) from a humid, temperate climate (southern New South
Wales, Australia) demonstrate that micaceous material can form from feldspar
during chemical weathering (Eggleton and Buseck, 1980). During the initial dissolu-
tion of the feldspar, circular holes are produced, which enlarge to form negative
crystals. Amorphous, ring-shaped structures develop in the larger holes. The
amorphous phase crystallizes into a well-crystallized 10 A layer silicate. In other
occurrences there is no evidence of an amorphous precursor and crystalization
occurred directly in solution holes. The 10 A physil has an irregular stacking
sequence, including lo-, 20-, and 30-A sequences. The extremely thin units are
similar to those of authigenic illites in sandstones. It is not known for certain
whether the 10 A material is illite or dehydrated montmorillonite. Based on mor-
phology the authors believe montmorillonite, illite, and 1Md muscovite are all
present. The x-ray patterns indicate I/S formed. Water diffusing through structural
defects, into feldspar, can attain K concentrations high enough for illitic material to
crystallize.
SEM studies of a weathered granite in France (Meunier and Velde, 1976) showed
secondary illites at the boundary between orthoclase and muscovite grains. The
relatively high Fe and Mg content of the illite indicates Fe and Mg migrated from
nearby biotite flakes. Kaolinite is forming higher in the outcrop.
In the lower portion of Tertiary kaolinite weathering crusts, formed on granites
and gneisses in Lower Silesia, Poland, K-feldspars first altered to 1Md illite (with
some smectite layers), whch altered to kaolinite. In the same zone (between the
saprolite and kaolinite zone) biotite was transformed to dioctahedral mica by losing
Fe and Mg and gaining Al (from feldspar) (Stoch and Sikora, 1976).
An isovolumetric study of a saprolite developed on a granite near Columbia,
South Carolina, suggested an intermediate K-rich phase preceded the development
of kaolinite (Gardner et al., 1978). They were not able to identify the K-phase, but
their data showed that K remained in the granite-saprolite to a more advanced stage
of weathering than did Na, Ca, Mg, and Fe. This should favor the growth of illite.
In a study of weathered granites and pegmatites in southeastern United States,
Sand (1956) found that much of the feldspar first altered to small books of mica
which then altered to kaolinite. Unfortunately it is not known if the mica is
hydrothermal in origin. However, Sands noted that there was a tendency for
secondary mica to be more abundant in areas where vegetation was sparse, and
presumably the p H higher. This suggests the mica is a weathered product.
Examination of a spheroidally-weathered anorthosite boulder from the Wichita
179

Mountains, Oklahoma (relatively dry, warm climate), indicated that the alteration
products were largely x-ray amorphous. A poorly crystallized illite occurred in the
outermost shell, even though there was relatively little K in the fresh rock (Fritz and
Mohr, 1984).
In addition to K-feldspar, the other most logical source of pedogenic illite is acid
volcanic rocks. Like K-feldspar, the data are inconclusive. In the Pampa region of
eastern Argentina, the soils (Pleistocene) are composed of wind-blown arkosic and
pyroclastic silt. The physils are almost entirely moderately crystallized Fe-rich illite,
with a few smectite interlayers. The illite was derived largely from weathered acid
volcanics and granites of the dry area of northern Patagonia (to the west) and blown
into the Pampa region.
Barshad (1966) reported the formation of illite from acid igneous rocks in an arid
environment; however, the mode of formation of the illite is not known. Other
studies which report an increase in illitic material from the base towards the top of
soil profiles found the increase was due to the fixation of K in smectite, I/S, or
vermiculite, and its partial conversion to illite.
Niederbudee and Fischer (1980) covered a column containing loess soil with a
thin layer of biotite and percolated a weak solution of CaCl, through the column.
Both the smectite and the vermiculite in the soil fixed K and formed illite (actually
I/S and I/V). In the desert and semiarid regions of South Africa illitic material
(10.1-10.3 A)
is more abundant in the upper A horizon than in the lower horizons.
This “reversed weathering sequence” is due to the fixation of K in I/S and I/V
(Van der Merwe and Weber, 1963). Analyses of the dryland soils (pronounced
summer dry season) of the southwestern United States (Nettleton et a/., 1973)
A
showed that 10 material was commonly abundant in the A horizon and vermicu-
lite and/or biotite/vermiculite in the B and C horizons. In this instance both K’
and NH; are believed to be released by plant decay. A similar mechanism
apparently accounts for the formation of illitic material from Fe-rich beidellite in
grassland soils (chernozemic to podzolic) of Saskatchewan (Arnaud and Mortland,
1963). In the arid and semi-arid lowlands of Morocco (Paquet and Millot, 1972)
degraded illite fixes K and becomes better crystallized. Mica in the surface horizons
of the Near East aridic soils is believed to have formed from smectite and soluble K
(Singer, 1987). Some of the arid soils of Iran have a relative concentration of illite in
the A horizon which is believed to be be wind transported dust (Mahjoory, 1975). A
similar situation exists in Hawaii where it has been shown, by K-Ar analyses, that
the illite is eolian in origin (Dymond er af., 1974).
“Illite formation” by K-fixation is equivalent to the “rejuvenation” that occurs
when stripped micas are deposited in sea water (Weaver, 1958):
degradation aggrrgat ion
illite ---f stripped illite + illite
K K

The final illite is basically an inherited illite that has been recycled and not an
authigenic or pedogenic illite.
It should be pointed out that the soil scientists have known for years that
alternate wetting and drying increases the amount of fixed K and the proportion of
180

60

50

40
I
0
W
0
30
W

<
-1

Y
t
j 20
- ./. ILLITE - 157c - 4s

ia

0
-0.3 -0.4 -0.5 -0.6 -0.7
LAVER CHARGE
-
(EOUIVALENTS PER HALF UNIT CELL)

Fig. 3-24. Percentage illite layers versus layer charge for K-smectites subjected to 100 W D cycles in water
at 60°C and 1 Sr-exchange. Numbers in parentheses refer to percentage of octahedral charge. Best fit line
is for rnontmorillonites having 69% or more octahedral charge. From Eberl et al., 1986. Reprinted by
permission from ACS Symp. Series 323,296-326. Copyright 1986 Amer. Chem. SOC.

illitic layers in expandable physils (Weaver, 1958; Eberl et al., 1986). Gaultier and
Maury (1977) alternately wetted and dried samples at temperatures ranging from
40” to 240°C and found that the temperature did not affect the amount of K that
was fixed. After more than 100 W-D cycles, samples do not expand in water;
however, ethylene glycol produces a partial expansion. Three successive extractions
with various cations showed the higher charged cations were more effective in
removing “fixed” cations; the amount of unexchanged K ranged from 62% for
NH: leaching to 37% for A13+ leaching.
Eberl et al., (1986) subjected a variety of K-smectites with various layer charges
to as many as 100 W and D cycles. The percentage of illite layers formed by the W
and D process was proportional to the total layer charge (Fig. 3-24), though the
tetrahedral charge was slightly more effective than the octahedral charge. The
percentage of illite layers ranged from 8 to 51 for samples W and D 100 times at 60”
and washed with Sr Cl,. Most of the change occurred in less than 20 W and D
cycles. As the total milliequivalents of interlayer cation are not changed signifi-
cantly, it is likely that the formation of 10 A layers is due to dehydration of the K,
which has a low hydration energy, rather than an increase in layer charge.
181

After three leachings with Sr, an appreciable portion of the “fixed” K was
removed, indicating it was not held as firmly as in natural I/S and illite; however,
most samples did retain some strongly bonded K after three washings. The per-
centage change in illite layers between one and three Sr washings correlates with the
angle of tetrahedral rotation (a).Increasing the rotation decreases the size of the
holes in the basal oxygen planes of the 2:l layers and the K ions cannot be buried as
deeply in the 2:l layers.
The equivalent of fixed K per illite layer decreases from approximately 1.0
(equivalents per half-unit cell) for 5 to 10% illite layers to approximately 0.35 for
40% illite layers.
When Na-smectite was mixed with K-feldspar and subjected to 100 W and D
runs the percentage of illite layers formed was similar to that found for the
K-smectite. Muscovite, glauconite, and phlogopite were slightly less effective.
These experiments, and many others, strongly suggest that in nature I/S can
form from smectite, in the presence of K and/or K-minerals, by periodic (daily or
annual) wetting and drying. Tidal flat environments should be ideal, except the K
has to compete with Mg hydroxyls for interlayer sites. However, it is not likely that
more than 50% probably less than 30%, of layers can be contracted by this method
unless the starting material is a stripped mica or illite rather than smectite.
Norrish and Pickering (1983) found that it is common for the illite content to
increase towards the surface in Australian arid and semi-arid soils. These illites have
12 to 14%Fe,O,, which is more than found in most illites. The authors suggest it is
a pedogenic illite formed under conditions similar to those in which the Fe-illites
form in alkaline lakes (p. 274).
All in all, there is little evidence for the large-scale neoformation of illitic material
in modern soils, though it is quite possible that a significant amount forms in aridic
soils. Weathering conditions favorable for the formation of illitic material (arid-al-
kaline) were probably more extensive in the past. As discussed in the chapter on
sandstones, authigenic illite or I/S with 5 to 20% smectite layers is a common
component and some appears to have formed at relatively low temperatures
( < 50°C?). Illite is ubiquitous in ancient continental red desert sandstones; how-
ever, it has not been established when it formed.
Certainly during recent geologic time we can conclude that most detrital illite or
I/S (with high illite) was derived from the parent rock, possibly via an intermediate
soil phase. The climate in the source area could range from cold to temperate to hot
and dry. Abundant detrital illite can also be generated in the tropics if elevation and
relief is high, as in the Amazon Basin. At best, detrital illite reflects low to moderate
weathering intensity. The character of the peak, sharpness, can be used to further
refine the source area conditions. Peak sharpness, in general, decreases from the
poles towards the equator.
One cannot assume detrital illites in Paleozoic and Precambrian rocks are not
pedogenic in origin without obtianing K-Ar age data. However, most K-Ar data
provide an age spectra for illitic material in shales, usually related to grain size.
There is generally a size fraction that has a pedogenic age. More work is needed in
this area.
182

Palygorskite and Sepiolite

Palygorskite, and to a much lesser extent, sepiolite are common constituents of


calcareous and calcrete soils in dry Mediterranean and arid climatic zones. Paly-
gorskite has been found in soils of north and south Africa, the Middle East, High
Plains of the United States, southwestern South America, and south-central Australia
(Fig. 3-25) (for references see Callen, 1984). In areas where the rainfall exceeds
about 30 cm palygorskite alters to montmorillonite (Paquet and Millot, 1972;
Bigham et al., 1980) or chlorite (Weaver and Beck, 1977). For palygorskite to be
preserved or to precipitate, soil waters have to be alkaline and have a relatively high
Si and Mg content. These conditions are usually restricted to arid and semiarid
climates. Callen (1984) notes that many of the calcretes containing palygorskite are
quite old, up to several hundred thousand years in Australia, and do not reflect
present climatic conditions.
Many of the soils which contain palygorskite are formed on Cenozoic and
Mesozoic sediments that contain palygorskite suggesting much of the soil palygors-
kite is inherited. In arid climates eolian transport is an important mechanism for
adding detrital palygorskite to a soil. For example, much of the attapulgite in the
Arabian Gulf and Red Sea is wind transported. It is difficult to prove that
palygorskite is pedogenic, but several examples strongly indicate it can form in soils
either by alteration of montmorillonite or by precipitating from solution.

PALYGORSKITES 0 m.y. LATE PLIOCENE - HOLOCENE

1
Fig. 3-25. Late Pliocene- Holocene palygorskite sepiolite occurrences. Dots are generalized DSDP and
oceanic occurrences, diagonal shading is continental data. Prefix D indicates soil or calcrete. Cross-hatched
areas are soils superposed on sedimentary basins with palygorskite. From Callen, 1984. Copyright 1984
Elsevier Pub. Co.
183

Singer (1984) reviewed the published data on palygorskite in soils and concluded
most of the modern pedogenic palygorskite formed in soils affected by rising ground
water, where rainfall was < 40 cm/yr, and the most common landscape was that of
flood plains or low terraces. Ground waters are frequently saline or alkaline with
pH values around 8; the soils are highly saline. Fine-textured calcareous, alluvial, or
aeolian sediments are common parent material for these soils. Calcite is the most
common carbonate associated with the palygorskite. Gypsum crystals are frequently
present, and along with calcite, may be coated with secondary palygorskite (Eswaran
and Bananji, 1974). Authigenic palygorskite commonly occurs as cutans or mats
coating ped surfaces.
Conditions that favor the development of calcareous crusts (calcretes, caliches,
and duricrust) also favor the development of palygorskite (Singer, 1984). These
crusts usually form in semi-arid regions where water is drawn to the surface by
capillary action and evaporated. Among other places, examples of pedogenic
palygorskite associated with calcareous crust have been described from Morocco
(Millot et al., 1969), Australia (Singer and Norrish, 1974), and western United
States (Allen and Pashai, 1981). Quaternary calcareous beaches along the coastal
plain of northwest Egypt have a caliche capping that contains a mesh of palygors-
kite fibers. The palygorskite crystallized in vadose fresh water from Mg that was
liberated when high-Mg calcite was converted to low-Mg calcite; corroded quartz
and feldspar grains were the source of the Si and A1 (Hassouba and Shaw, 1980). It
is of interest to note that authigenic palygorskite also formed in lagoonal-sabkha
gypsiferous marls that were deposited in the depressions between beach ridges.
Nettleton et al. (1985) studied 33 Aridisols and Entisols profiles from southwest-
ern United States and Jordan which contained palygorskite. Horizons with a high
palygorskite content had a saturated paste pH of 7.3 to 8.3. They concluded that
palygorskite is forming in soils that have desert pavements, about half those
examined. The gypsiferous soils of Iraq contain abundant authigenic palygorskite.
As the rainfall increases, palygorskite decreases and montmorillonite increases
(Bananji et al., 1975).
Though numerous studies of arid soils from Saudi Arabia indicate inherited and
eolian palygorskite are common, several investigations indicate much of the paly-
gorskite is pedogenic. Elprince et al. (1979) found that in eastern Saudi Arabia
palygorskite was present in saline sabkha soils formed on a variety of geologic
materials. Mackenzie et al. (1985) found palygorsbte in soils formed on a variety of
source rocks in the Wadi-ar-Rimah, central Saudi Arabia. The Wadi area is subject
to sporadic precipitation and flooding during the late winter. Where permeability is
low, shallow ponds form and evaporate to leave salt crust giving rise to sabkha soils.
Where permeability is higher, the soil waters gradually become richer in Si and Mg
and palygorskite and/or smectite form. In both areas the palygorskite is associated
with smectite and the authors believe that when palygorskite is being formed,
smectite is dissolving and vice versa. The thermodynamic calculations of Weaver
and Beck (1977) (Fig. ) show that the stability fields of palygorskite and montmoril-
lonite share a common boundary and only minor changes in the concentration of
Mg and/or Si will favor the development of one mineral over the other.
184

Another type of occurrence of pedogenic palygorskite is found in the coastal area


of southeast South Australia (Hodge et al., 1984). A series of parallel calcareous and
quartz dune sands are aligned parallel to the shoreline. The interdune flats contain
shelly sands, marls, and clays on which palygorskite-rich (with associated
montmorillonite) ground-water rendzinas and associated swamp soils have formed.
The ground water has a composition similar to seawater and may be the source of
Mg, but seasonal flooding supplies fresh water to the area.
The conditions under which soil palygorskite forms is fairly well established. Mg
can be supplied by either marine or fresh water. During evaporation the Mg/Ca
ratio can be increased by the precipitation of calcite or gypsum. In other occur-
rences Mg is supplied by the conversion of high-Mg calcite to low-Mg calcite. The
high pH allows for some dissolution of quartz and other silicates to produce the

Fig. 3-26. White clay vein in sandy soil at the boundary between the Upper and Lower Miocene, Georgia.
Fibers of palygorskite and opal spheres extend from the edge of the vein. Bar = 1 pm.
185

necessary Si. Though it is usually ignored, appreciable A1 is also required for the
formation of palygorskite. Under high pH and high salt concentrations, some
physils can be partially broken down and A1 and Si made available for reorganiza-
tion into a chain structure. The common association of palygorskite and montmoril-
lonite strongly suggest that montmorillonite is often the source of A1 and also of
some of the Si and Mg. When the pH is decreased by increased rainfall, palygorskite
is desilicated and reverts back to montmorillonite.
Though palygorskite weathers relatively easily, it has persisted in a number of
paleosoils. Playgorskite coating fossil ped surfaces has been reported from an
unconformity in the basal Permo-Triassic of northwest Scotland (Watts, 1976) and
from a soil zone between the Lowtr and Middle Miocene of southwest Georgia
(Weaver and Beck, 1977). Fig. 3-26 shows palygorskite fibers and silica spheres
(opal-cristobalite) coating a quartz grain and tapering fibers growing into veins in
the Miocene soils. Smooth mats of palygorskite form on the vertical surface of peds
(cutains).
Though palygorskite is a relatively rare physil it is a component of soils, playas,
alkaline lakes, and lagoons in the arid to semi-arid climatic belt. Palygorskite is a
common component of marine sediments that lie seaward of continental palygors-
kite deposits. This is true for both recent (Fig. 3-25) and ancient deposits (p. 699).
There is a reason for this. Palygorskite is much lighter than other minerals and
should be preferentially transported by water and wind. It forms in near-surface or
surface environments that are periodically evaporated to dryness and can easily be
eroded by the wind. Many of these areas are periodically inundated by flood-like
waters which would transport appreciable material to the major drainage system.
Once the palygorskites reach the continental shelf they are transported to the deep
sea by turbidity currents and slumps (Weaver and Beck, 1977). It is likely that the
soils themselves are only a minor source of detrital palygorskite.
If palygorskite forms in normal marine as well as in continental and coastal
environments, then it is useless as a paleoclimaticc indicator. If it does not form in
the oceanic environment, other than hydrothermal, which I believe, then it is an
excellent climatic indicator: arid to semi-arid with seasonal rainfall.

Miscellaneous Physils

Talc and pyrophyllite usually occur as minor componentss in a few soils. The
latter mineral is inherited but talc can form during weathering. Talc, along with
Fe-smectite, formed during the weathering of an olivine pyroxenite in the western
Ivory Coast (Paquet et al., 1982). The talc contains appreciable tetrahedral A1 and
octahedral Fe. A typical formula is:

(Fei,15
Mg 2.51 )(si3.88A10.12 )OlO (OH),
186

Interpretation

It is obvious from the discussion on weathering that climate is only one of the
factors that influences the formation and transportation of physils. Time, relief,
permeability, rock type, etc. are all factors that determine the composition of a clay
suite that will be delivered to a depositional basin. Local conditions commonly
override the global climate pattern, e.g., smectite in the humid tropics. Dual source
areas, either lateral or vertical, further complicate the interpretation. Further,
differential flocculation, differential settling, authlgenic modifications, pellitization
and current patterns in the waters of the depositional basin will serve to mask the
composition of the source area. Interpreting paleoclimatic conditions from physil
suites, without supporting data, is risky. But, nothing ventured, nothing gained.
The distribution of kaolinite, illite, and smectite in soils and outcropping sedi-
ments extends over nearly the full range of latitudes. A relatively large proportion of
transported detrital physils are derived from sedimentary and metamorphic bedrock
and do not reflect climatic conditions. This has been even more true in the past than
at the present time. The modern type soil profile did not develop until the
Cretaceous. Of course, another problem is the time lag, which can be tens of
millions of years between the development of a weathered crust and its erosion.
Two physils, palygorskite and chlorite, have some restricted environmental
significance. Abundant palygorskite reflects a warm semi-arid to arid climate in the
source area. It is both formed and preserved under these conditions. Chlorite is
extremely sensitive to chemical weathering, and abundant detrital chlorite that
produces sharp x-ray peaks indicates a cold, or possibly arid, climate. However, it
must be kept in mind that chlorite in sedimentary rocks forms diagenetically at
relatively low temperatures. In modern marine sediments chlorite, and to a lesser
extent illite, become more degraded (broader x-ray peaks) as the equatorial region is
approached; however, the same trend, with depth, exists in most temperate soil
profiles.
Kaolinite, if it can be proven that it comes from soils, indicates moderate to high
rainfall but not necessarily tropical conditions. Smectite, in addition to the fact that
they form in soils under a wide variety of climates, have the added complication
that they can form from volcanic detritus in the depositional environment.
The distribution and climatic significance of clays in recent marine muds is
reviewed in Chapter V. It is evident that on a global basis there is only a limited
match between climatic zones and physils, and this occurs primarily in the Atlantic
Ocean.
The French, particularly G . Millot, H. Chamley, and C. Robert, have been active
in using physils for determining paleoclimates. Examples are given in Chapter IX.
Some of the changes in physil suites in the ocean that are attributed to climate
actually reflect shifts in ocean current patterns caused by continental drift. Fig. 3-27
+
shows the ratio of chlorite mica/total physils in a series of wells on the Continen-
tal Rise of the northwest Atlantic Ocean (Lancelot et al., 1972). The abrupt increase
near the beginning of the Pleistocene presumably reflects a true climatic change,
cooler temperature (but also more glacial erosion). On the other hand, a similar
187

ZONE <2um

r
(BLOW, 1 9 6 9 ) :. . I4 .5 .6
I .7
1 .a .9

W
z
w

-
i-23
v1
w

2
-+ -&-?! ,
P A

a-22 - 0, 9; wl _-
[u
- A A

-0
z--22 0
g
A
a
W
Y 1 x o
--
-19
-
-18
w
z
V
W -
X
a
%-17 X?_O f Oh

‘iI +
2
-16 -
0
-
-15 0

-14 0

-13
w

i-
z
V
W

2-12
x
r: -11 +
-10

,9 SITE A = amphiboles p r e s e n t

mineralogical change occws between the Eocene and Oligocene in the northeast
Atlantic (Latouche, 1978). In this instance the change coincides with the opening of
the northeastern Atlantic and the movement of cold Arctic bottom waters, and
physils, southward.
A study of the oxygen isotopes ratios of planktonic foraminifera and the physils
in a 15 m long core (representing 33,000 years) seaward of the Niger Delta shows a
good correlation between changes in the physil suite and rainfall. Dry and wet
periods (based on isotopic data) are marked by diffrential soil erosion in various
geographic areas of the large drainage basin. During humid periods there was an
188

increase in detrital kaolinite and goethite which was derived from the upstream
equatorial part of the river basin. During dry periods the content of smectite,
chlorite, and palygorskite increased. These physils were derived primarily from the
downstream, poorly drained soils (Pastouret et al., 1978).
There is a problem in correlating global average climate with global physil
composition. For example, the Cretaceous is presumed to be the dryest period (Fig.
9-4), yet kaolinite deposits are abundant and kaolinite content of shales is relatively
high (Fig. 9-1). Another example is the Eocene. It was a very wet epoch and
abundant kaolinite was formed. Also, it was the epoch in which probably the most
palygorskite, indicative of an arid climate, was formed on land.
There are a number of temporal related factors to consider when interpreting the
climatic significance of physils. One major concern of paleoclimatologists is the
probable changes in the equator-to-pole temperature gradient through time. I have
discussed chemical weathering and physil distribution in the framework of the
present-day temperature gradient; however, several lines of evidence suggest that
the equator-to-pole temperature gradient was considerably less at times in the past
than it is today ( e g , Frakes, 1979; Parrish et al., 1982). For example, the equator-
to-pole temperature gradient during the Cretaceous was much less than today and
there is no evidence of polar or glacial ice.
The make-up of “soils” has changed drastically with time (Yaalon, 1962).
Protosoils developed directly from original mafic crust and prevailed until the
appearance of life. High proportions of Fe2+ and poor separation of Si, Al, and Fe
are characteristic of Precambrian weathering. Primitive soils, incipient regolith
covered with incompletely leached soil, prevailed during the late Precambrian and
early Paleozoic, until the appearance of land plants. Beginning with Devonian time
( - 400 m.y. ago) land plants spread rapidly, presumably increasing the intensity of
leaching and weathering developing a “rudimentary pedosphere”. The fully evolved
pedosphere with clearly differentiated profiles and soil types developed with the
appearance of flowering plants in the Cretaceous ( - 130 m.y. ago).
Each stage in soil development should have increased the intensity of acid
weathering and the rate of removal of the more mobile elements and increased the
residual concentration of Al-rich phases. The amount of kaolinite in soils should
have become increasingly dominant in younger soils and sedimentary rocks. This
has not happened. Many soils are eroded before they reach maturity but, equally
important, only about half the modern soils are acid and a fair portion of them are
tundra. Around 20% of all soils are Aridisols.
Theoretically the primitive soil period, late Precambrian and early Paleozoic,
should be better suited than modem soils for the crystallization of the more
chemically complex physils. Due to the lack of plant life, leaching was limited and
the soils were more apt to be alkaline. Smectite, illite, and chloritic minerals should
be the physils most likely to form under these conditions. During this time period
granitic metamorphic and acid volcanic rocks outcropped over a much larger area of
the earth surface than in the following periods. Among other things, this indicates
an abundance of parent rocks with a high K content. It seems likely that during this
time period the major pedogenic physils were illite and I/S, which were the source
of much of the illite and I/S in the low Paleozoic shales.
189

CHAPTER IV

CONTINENTAL TRANSPORT AND


DEPOSITION

RIVERS

Introduction

The large bulk of physils in recent environments are detrital in origin. Most
physils are transported by rivers and deposited in near shore environments. In this
section the erosion transport and deposition of physils are discussed. Additional
data on river physils are presented in the sections on deltas and estuaries.

Transportation

Various estimates have been made of the mass of solids delivered by rivers to the
ocean basins each year. Values range from 13.5 to 64 billion short tons per year.
Holeman (1968) gave a widely quoted value of 20.2 billion tons. More recently,
Milliman and Mead (1983) suggested a value of 13.5 billion tons. The smaller value
is primarily due to a decrease in the estimates of the load carried by Asian rivers.
Table 4-1 lists the average sediment discharge of the 21 major rivers.
Nearly 50% of the suspended material supplied to the oceans comes from Asia
which contains only 16.9% of the world’s total drainage area. Approximately 43% of
the Asian suspended load is transported by the Ganges and Yellow Rivers. At times,
the Yellow River flows contain as much as 40% sediment by weight. Holeman also
points out that only a small portion of the eroded material in watersheds is
transported to the oceans. In the Potomac River Basin only 5% of the eroded
material reaches the Potomac estuary. Only 24% of the sediment moved by the
Yellow River reaches the ocean (Milliman and Meade, 1983).
What is the percentage of physils in the suspended load? The data are sparse. The
suspended load of the mountain tributaries of the Amazon River (Gibbs, 1967)
contain from 50 to 74% physils of which approximately one-third are smaller than 2
pm. The tropical rivers carry from 83 to 98% physils with 70 to 100% being finer
than 2 pm. The total weight of physils in the tropical rivers is much less than in
mountain rivers. Assuming that the average suspended load contains 60% physils, a
value similar to that for the percent physils in shales, then each year rivers transport
190

Table 4-1
Quality of Data Base For 21 Largest River-Sediment Discharges To The Ocean. After Milliman and
Meade (1983).
River Average Sediment Adequacy o f
Discharge (1Oht/yr) Data Base
1. Ganges/Brahmaputra 1670 Inadequate
2. Yellow (Huangho) 1080 Good
3. Amazon 900 Inadequate
4. Yangtze 478 Good
5. Irrawaddy 285 Inadequate(?)
6. Magdalena 220 Inadequate
7. Mississippi 210 Good
8. Orinco 210 Sufficient
9. Hungho (Red) 160 Inadequate
10. Mekong 160 Sufficient
11. Indus 100 Sufficient
12. MacKenzie 100 Poor to fair
13. Godavari 96 Inadequate
14. La Plata 92 Inadequate to sufficient
15. Haiho 81 Good
16. Purari 80 Inadequate
17. Zhu Jiang (Pearl) 69 Sufficient to good
18. Copper 70 Sufficient
19. Danube 67 Good
20. Choshui 66 Sufficient
21. Yukon 60 Sufficient

approximately 8-12 billion tons of physils to the oceans. At least half of this
material is probably finer than 2 pm.
Most clay size material is carried in suspension. The bed load seldom contains
much clay except in the form of shale chips, and mud balls; however, the bed load
near the mouth of the Mississippi River is reported to contain nearly 20% of clay
size material (Leopold, et al., 1964). This material may be in the form of flocculated
clay granules. The amount of bed load is difficult to measure but is generally
assumed to be less than a tenth the suspended load (Milliman and Meade, 1983).
though in arid and semi-arid climates it can constitute half the total load.
The two most important mechanisms in the suspension of sediment are turbu-
lence and diffusion. Turbulent motion results from eddies which are constantly
being formed by the shearing action of the fluid, while already existing eddies are
being destroyed by viscous friction. The diffusion mechanism involved two essential
features. One is the simple transport that results from the turbulent velocity
fluctuations, and the other is a mixing of the transported fluid with its new
surroundings. In the diffusion process, the flow of sediment is in the direction of
decreasing concentration or opposed to the concentration gradient. The tendency is
to equalize the concentration.
It is obvious that before material can be suspended it must be removed from the
bed, therefore the bed load must be in motion before suspension can occur. As this
191

bed motion develops, an area of high sediment concentration forms just over the
bed. It is in this area that the distinction between bed load and suspended load
becomes obscure. The larger turbulent eddies are capable of scouring into this area
thus lifting sediment particles into true suspension. The finer sizes can be main-
tained in suspension by the smaller eddies present, and the coarser particles fall
back to the layer of concentration above the bed. The net effect of this process is to
suspend the fines and concentrate the course material near the bed.
Quantitative data on the factors that effect the suspended sediment load and
particularly the suspended clay load is limited. The amount of suspended sediment
transported is influenced by the amount and distribution of rainfall, temperature,
topography, rock type, and vegetation. These factors, to a large extent, determine
the rivers capacity as well as the rate of weathering and thus determine the
availability of physils as well as the streams available to transport them. The
suspended sediment load increases with increasing velocity, but the nature of the
relation is controlled by the factors just mentioned (Leopold et al., 1964).
The total dissolved load and the relative amount of dissolved load varies greatly.
Leopold et al. (1964) reports dissolved loads ranging from 1%to 64% of the total
dissolved and suspended load. The average value is 20%. They report that in the
United States the proportion of dissolved load increases with increasing annual
runoff. In the wetter regions, the dissolved load is a function of the availability of
salts in the rocks (weathering) and not the available runoff.
The concentration of dissolved material in a river water decreases with increasing
flow, whereas the suspended load increases with increasing flow. In arid and
semi-arid regions where sediment production is high, the dissolved load is a
relatively small part of the total, but the concentration is high. Where precipitation
is greater, runoff higher, vegetation abundant, and chemical weathering effective,
the relative percentage of dissolved load increases (but the concentration is low)
(Leopold et al. 1964).
The relative amount of clay-size material in the suspended load varies not only as
a function of such things as degree of weathering and amount runoff but as to
position in the river channel and day-to-day variations in the amount of flow. Most
streams have sufficient velocity to transport clay-size material. The amount of clay
present is determined by its availability (lithology, weathering, and runoff).
Particle-size analyses of suspended sediment from five moderate sized Georgia
rivers (Kennedy, 1964) showed that the clay content ranged from 25 to 50%.
Individual samples contained as much as 75 to 80% clay. During a storm the
concentration of suspended clay was considerably higher than that of sand and silt
during the early part of the rise in discharge; however, as the discharge continued to
rise, the proportion of sand and silt increased. Some of the rivers had their highest
concentration of clay at h g h rates of discharge and others at lowest discharge.
Kennedy noted a general increase in the proportion of suspended clay in stream
water after intense rainfall which he believed reflected the fact that fine grained
material is derived mainly from the land surface during storms and the coarse-grained
material already in the stream bed is the source of much of the sand. Thus, the
concentration of clay in suspension would be related more directly to intensity of
192

y1

U
0,

d 400 6

C
0
8
0

'! 1
lil
5
V
200

Tributary
Fig. 4-1. Variation of suspended solids of Amazon River and its tributaries. A, Concentrations of main
stream of Amazon. Open circles, concentrations in dry seasons; closed circles, concentrations in wet
seasons of indicated tributaries. From Gibbs, 1967, Geol. SOC.Amer. Bull., 78, 1203-1232.

rainfall and only indirectly to water velocity.


Fig. 4-1 illustrates how the concentration of suspended solids vary as a function
of relief and season in rivers of the Amazon Basin (Gibbs, 1967). Concentrations are
much higher in high relief areas (Andes) and during the wet season. Kennedy (1964)
found that 75% of the total suspended load was transported during floods (10% of
the total time). A study of Two 0-Clock Creek in the Canadian Rockies showed
that 87% of the total sediment load was exported during the one flood generated by
peak snow melt (McPherson, 1971). In a study of small midcontinent streams, Love
(1936) established that it is common for 96% of the total load to be transported in 2
to 3% of the total time.
Gibbs also established that the mean particle size of the suspended load in-
creased as the relief and amount of runoff increased. He calculated that relief was
193

much more important (8 times) than precipitation or temperature in controlling the


particle size of the material transported. Kennedy determined that on a river rise of
only a few feet the range in grain size could be as great in the river cross section at a
given time as it was for a fixed location through the rise and fall of the river. Careful
planning is required to collect meaningful river samples.
In the Amazon River and its tributaries, Gibbs (1967) found that in rivers of low
concentration ( < 30 ppm), low-relief tropical-environment, the amount of material
in suspension did not vary with depth. The mean particle size was < 2 pm. For
rivers with 30 to 250 ppm, mountain tributaries, samples taken at 0.9 total depth
had 20% greater concentration than surface samples. The size of the suspended load
particles was larger than for the low concentration rivers and the increase was due
to an increase in the amount of larger particles with depth. For the mountain
tributaries, 90 to 95% of the rivers’ solid load is in suspension. At the mouth of the
Amazon, 95% of the total load is the suspended load. The value at the mouth of the
Congo is 96% (Spronck, 1941), and for the Mississippi, reported values range from
99 to 80% (Henry and Bader, 1961)

Mineralogy

Little effort has been made to determine how the mineral composition of the clay
suite of a river varies as a function of time. During the years 1959 to 1961 I
collected a series of suspended samples from the Arkansas f iv e r at Ponca City,
Oklahoma (Weaver, 1967). The Arkansas River heads in the Rocky Mountains of
Colorado, flows through eastern Colorado, western and central Kansas before
turning south into Oklahoma. The river has no major tributaries. Samples were
taken from the same location near the middle of the river, approximately one foot
below the surface. X-ray analyses were made of the less than 2pm fraction of each
sample. Illite and I/S are the dominant physils.
Fig. 4-2 shows how the clay suite varied during portions of an extensive storm.
Illite and I/S, which vary inversely, have a wide range of values. The clay suite
closely reflects the character of the surrounding sediments and the distribution of
the rainfall.
During times of normal discharge, the illite to mixed-layer ratio is approximately
one. Local runoff increases the proportion of mixed-layer clay. As water from
storms to the north in Kansas and Colorado reach Ponca city, the proportion of
illite increases and the color of the clay changes from brown to gray. Generally,
montmorillonite follows illite. In the waning stages of floods which extend into
Colorado the illite and montmorillonite dominate the clay suite, presumably indicat-
ing that most of the suspended clay load at this stage is a material from far to the
northwest in Colorado.
Fig. 4-3 shows the fluxuation in the illite and the I/S content of the suspended
clay suite over a two year period. The variation, during a single flood, can be nearly
as large as the total range of variation. Repeated settling and resuspension over long
periods of transport time tends to homogenize the physil suites that complex rivers
194

Water level
10

Normal 0 -

20 -

9AM 7PM 5AM 3PM IAM 11AM 9PM


July 15, 1959 July 16,1959 July 17,1959
Fig. 4-2. Variations in suspended clay suites of the Arkansas River during a major storm which covered
the area from Ponca City, Oklahoma, northwest to Colorado. Color change coincides with change in clay
suite. From Weaver, 1967. Copyright 1967 SOC. Econ. Paleo. Miner.

0 w
50
w
w

30
w
i
w

70 - B Illite
I/S
195

deliver to the ocean but it is unlikely (and largely unknown) that the homogeniza-
tion is complete. Further mixing is accomplished by burrowing organisms and
marine currents.
In a study of the semi-arid Cheyenne River drainage basin of eastern Wyoming,
Rolf and Hadley (1964) determined that the suspended clay load was not similar to
the clay suite in the surrounding soils. The mineral composition and texture of the
clay and silt fractions of the transported material changes as it passes from upland
slopes (by sheet erosion, finger gullies, and rills) to the stream channels. The
weathered material is sorted en route to the stream channels, the texture becoming
finer and the physil suite simpler. The degraded micas (vermiculitic) d o not reach
the stream. Further, the suspended clay suites are different for low and high flows, a
situation similar to that reported for the Arkansas River.
A dissimilarity between the soil and suspended clay suite was noted in a small
mountain valley of the northern Appalachian Mountains. The suspended physils in
a trout stream were illite and chlorite derived from the relatively unweathered shale
outcrops from the steep creek banks, rather than weathered chlorite and illite from
the thin soil horizon. In the larger, lower relief Susquahanna River, a higher
percentage of the suspended load was derived from the heavily farmed flood plain
soils and degraded illite and chlorite are relatively abundant.
In Puerto Rico the highlands of the Cordillera Central are composed of Creta-
ceous basalts and andesites. The highly weathered top soil ( - 4 m) on the humid
north flank (windward) is composed of halloysite, gibbsite, and quartz. Montmoril-
lonite, chlorite, and vermiculite (14A phases) gradually increases in abundance
below this zone becoming the dominant clays at 10 m. The rivers draining north
from this region carry a mixture of halloysite and the 14A phases indicating the
whole soil profile is being eroded and the river clay suite does not reflect the surface
weathering conditions in the source area. Gibbsite is not present in the river muds
suggesting it rapidly altered to some other phase. The volume of volcanic sand
transported by the river is much greater than the volume of clay. The physil suite of
the marine sediments after a moderate amount of burial will likely have a high
content of montmorillonite and/or chlorite and be strikingly different than that
formed by the present day climatic conditions. It should be noted that on the dryer
lee slope, montmorillonite-vermiculite is the dominant physil in the soils and in the
rivers (Ehlmann, 1968).
In his classic study of the Amazon Basin Gibbs (1967) demonstrated the role of
relief in determining the rate of erosion and amount of suspended solids supplied to
the ocean. A young mobile belt (Andes Mountains) is present in the western part of
the basin but for much of its length it flows over the Brazilian Shield. Based on
suspended sediment concentration and discharge he calculated that 82% of the
eroded material was derived from the 12% of the basin comprising the mountainous
Andean environment-type.
Table 4-2 summarizes the physil data for Gibbs’ three relief zones. Though the
low-relief tropical rivers have a relatively high percentage of physils their total
contribution to the Amazon River is small.
Fig. 4-4 shows the mineral composition and size distribution of the suspended
196

Table 4-2
Summarized physil data for Amazon Basin (After Gibbs, 1967).
_____

Tributary Percent physils Percent of Removal Rates/yr


basin of total sample in total
environment each size fraction sample
<2p 2-2op >2op 10‘ g/km2 10l2 g/basin
Mountainous 18-58 7-28 2-8 50-74 156-163 63-66
Mixed 60-79 3-12 0-5 73-79 46-145 6.8-97.9
Tropical 70-98 0-19 0 83-98 0.8-58 0.03-12

load of a stream draining a tropical and a mountainous environment and at the


mouth of the Amazon River. Physical weathering is dominant in the mountains as
demonstrated by the abundance of 2M mica, unweathered chlorite, and feldspar.
Montmorillonite is apparently derived from “calcic” rocks in the Andes. The
abundance of kaolinite and gibbsite in the low-relief tropical streams indicates that
chemical weathering is the dominant weathering process. In this example it is
apparent that relief, rather than climate, largely determines the physil suite that is
delivered to the ocean. The composition of the physil suite is determined primarily
by the source rocks, however, the controlling source rocks comprise only 12% of the
overall source area.
To the south of the Amazon River, the San Francisco River, whose drainage
basin is predominantly in the humid tropics, carries a high content of kaolinite. Off
shore muds contain 40 to 50% kaolinite (Griffin et al., 1968).
In contrast to the Amazon River, the Mississippi River drains a relatively low
relief craton area. The climate range is more restricted, ranging from temperate-
humid to semi-arid or arid. A study of The Mississippi River Basin (Potter et al.,
1975) floodplain muds showed that the source rocks were the principal control on
the physil suite delivered to the Gulf of Mexico. The basin can be divided into two
major source areas. The outcropping material in Western two-thirds of the basin is
largely Cretaceous and Tertiary rocks and Pleistocene loess (with a lesser amount of
upper Paleozoic). All three types have a relatively high content of montmorillonite
and varying amounts of illite, kaolinite, I/S, and chlorite. The eastern rocks are
largely of Paleozoic age. Illite is the major physil in most of theses sedimentary
rocks. Varying amounts of I/S, kaolinite, and chlorite are present. Potter et al.
(1975) analyzed the < 10pm fraction, so their percentage values are somewhat
different than those obtained using the < 2 pm fraction, but are more typical of the
river suspended load.
The physil suites of the river muds are similar to those of the source rocks. Fig.
4-5 shows that montmorillonite is dominant in the muds from the western and
central portions of the Mississippi and absent from the eastern tributaries and the
southwestern region where Paleozoic rocks outcrop. The high values in the north
central area are due to the presence of Pleistocene loess overlying Paleozoic rocks.
Kaolinite (app. 20 to 25%) is relatively uniformly distributed but is more abundant
in the areas containing Paleozoic rocks. Illite and chlorite are most abundant in the
river muds from the eastern portion of the basin (Fig. 4-6) where the Paleozoic rocks
197

Tropical
txingu 1

1'1'I ' 1 I 1 ' 1 ' 1 ' I ' \ I


Gibbsite
Mountainous
(Maranon)

"'.,

/
I Kaolinite \

1c 8060 40 20 10 8 6 4 2 10.8 0.6 0.4 0.2


Particle diameter in microns
Fig. 4-4. Size and mineralogic frequency distribution of the suspended solids of a tropical and of a
mountainous environment tributary and of the Amazon. From Gibbs, 1967, Geol. Soc. Amer. Bull.. 78,
1203- 1232.

outcrop. There is a close association of the physil suite of the river muds to that of
the outcropping rocks. Potter, et a1.(1975) compared the crystallinity (K.I.) of the
10A peak of the river muds with those of the Paleozoic shales and concluded that
the unweathered rocks rather than soils were the primary source of the physils. This
may merely reflect the youth of the soils. The kaolinite values are higher than one
would expect from the composition of the source rocks.
Fig. 4-7 shows the average physil suites of the muds in the main streams. It is
evident that by far the largest volume of suspended physils is derived from the
198

Fig. 4-5. Basin-wide distribution of smectite, which is present in most of the alluvial muds west of the
Mississippi River and absent from most alluvial muds of the Ohio Basin. From Potter et al., 1975.
Copyright 1975 Elf-Aquitiane.

western portion of the basin. The Missouri has an average annual suspend load of
240 X 10-tons and the Ohio 15 X 10-tons (Holeman, 1968). The muds immediately
below the confluence of the Ohio River and the upper Mississippi River have the
same average composition as those of the Missouri River. Fig. 4-8 shows the
mineral-size distribution of a typical sample from this area. Note most of the “clay
minerals” are not clays ( < 2 pm).
Table 4-3 shows the composition of the Mississippi River muds south of Cairo
(approximately 600km north of the river mouth), at the mouth of the river, and
prodelta muds from the Mississippi Delta. The physil suites are very similar
indicating there has been little physical segregation or authigenic modification of
the physils.
It should be noted that the average montmorillonite is actually an I/S containing
approximately 80% montmorillonite layers and it comprises 70 to 80% of the
< 2 pm fraction.
199

Fig. 4-6. Basin-wide distribution of illite. which is most abundant in portions of the Ohio Basin. although.
like kaolinite, it is present virtually everywhere throughout the Mississippi basin. From Potter et al.. 1975.
Copyright 1975 Elf-Aquitiane.

Both the Mississippi River and Amazon River suspended physil suites, at the
mouth, are determined by the composition of the source rock; however, in the
former the major contributing source rocks occupy approximately 70% of the
drainage basin and in the latter only 12%. Mostly, this is due to the difference in
relief. Thus, rather than rivers suspended physil suite being representative of the
whole basin, from a practical standpoint the suite is primarily indicative of the
effective source area, which may or may not be representative of most of the basin.
East of the Mississippi drainage basin the numerous small rivers demonstrate the
interrelation of source, climate, and relief. The climate of the S.E. United States is
humid subtropical and chemical weathering is more intense than in the interior.
Though montmorillonite is the dominant physil in the Cenozoic Coastal Plain rocks
the soils, except those near the coast, have a high kaolinite content. This is reflected
in the composition of the suspended load of the Mobile River (western Alabama)
and the Apalachicola River (Alabama-Georgia boundary). The suspended physil
200

Chlorite

Kaolinite
Of

Fig. 4-7. Physil suites in the mains streams of the Mississippi River system. The western source is
dominant. From Potter et al., 1975. Copyright 1975 Elf-Aquitiane.

suite ( < 2 pm) of the former contains 40 to 50% kaolinite and the latter 60 to 80%
(Griffin, 1962).
From south to north along the eastern margin of North America the climate
changes from humid subtropical to humid continental to continental subarctic.
Chemical weathering and suspended river clay suites are closely related to climate,
though there are small scale exceptions.
Fig. 4-9 is a very generalized clay mineral distribution map of the eastern United
States and the Continental Shelf. This distribution is not based solely on the
composition of the soils, but is intended to represent the composition and distribu-
tion of the physils that are available for erosion and transport, and thus includes the
soils and rocks from the surface to the base level of the streams. In some situations,
the soil contributes most of the suspended physil load (southern Appalachians); in
others, the soil contributes little, and most of the physils are obtained by the erosion
201

10 prn

-
2 6.3 20 63 200
log particle size
Fig. 4-8. Fraction analysis of Mississippi alluvial mud from levee bank on the east side of New Madrid
Bend, 25-A-I, Fulton Co.. Kentucky. From Potter et al., 1975. Copyright 1975 Elf-Aquitiane.

of relatively unweathered argillaceous rock from the banks of the streams (northern
Appalachians).
Kaolinite is the dominant physil in the deeply weathered saprolitic soils of the
southern Piedmont. The soils formed largely on illite-chlorite metamorphic rocks.
Northward, as weathering decreases, soil profiles are thinner and considerably less
kaolinite is available in the erosion zone. Illite and chlorite, along with vermiculite
(weathered), in the metamorphic rocks and shales are the main physils available for
stream transport.

Table 4-3
Average composition of marine gulf of Mexico deltaic muds and Mississippi river muds ( (10 p m
fraction), After Potter et al, 1975.
Minerals Mississippi River Mississippi Marine
muds below Cairo Mouth Gulf muds
Mixed-layer clay minerals 0 0 0
Smectite 40 40 45
Mite 25 20 20
Chlorite 5 7 5
Kaolinite 20 16 15
Quartz 10 14 12
Feldspar 0 3 3
Carbonate 0 0 0

Samples 2 12 11
202

r I

I Explanation
Predominant clay mingrals
63;21- Montmorillonite
m-Kaolinite
a- Illlte> Chlorite

Fig. 4-9. General surficial clay mineral distribution in eastern United States and Continental Shelf area.
From Neiheisel and Weaver, 1967. copyright 1967 SOC.Econ. Paleo. and Miner.

The southern Piedmont is fringed by the Upper Cretaceous Tuscaloosa Forma-


tion which also has a very high kaolinite content. The southeastern Coastal Plain is
an area of low relief and contains Tertiary sediments in which montmorillonite is by
far the most abundant physil; the montmorillonite commonly weathers to kaolinite
but the zone of alteration is relatively thin.
The physils in the surficial Continental Shelf sediments are predominantly illite
(Biscaye, 1965), derived from the central and northern Appalachians and carried
southward by longshore currents.
The major portion of the Piedmont Province of South Carolina and Georgia is
drained by the watersheds of the Santee, Savannah, and Altamaha Rivers. Some of
the Coastal Plain streams are also tributary to these rivers in their course to the sea.
As the diagnostic kaolinite (and hornblended) suite from the Piedmont and Tusca-
loosa Formation is carried to the estuaries or tidal deltas along the coast, it is
diluted by the contrasting mineral suite (montmorillonite) of the Coastal Plain
sediments. The composition of sediment deposited along the coast reflects the
proportional amount of the two source areas in the watershed.
203

Fig. 4-10. Composition of muds in a Georgia river system and estuary (Altamaha Sound) over a distance
of 350 km. Physil suite is influenced by rock type and tidal currents.

The data in Fig. 4-10 illustrate how the physil suite of muds from a central
Georgia river system is influenced by the various rock types the rivers flow through.
The Piedmont and Cretaceous kaolinite-vermiculite suite is progressively diluted by
the influx of montmorillonite from the younger Coastal Plain sediments. Sample
No. 9 is from the mouth of a major tributary, the Oconee River, which drains a
similar sequence of rocks. The physil suite of the seaward most sample is influenced
by tidal influx which is presumably the source of the illite (20%).The high kaolinite
at the head of the estuary suggest there has been some differential settling.
The relative amounts of the Piedmont to Coastal Plain drainage areas in the
watersheds of the major rivers decreases in the southwesterly direction because of
the greater width of the Coastal Plain and increased density of the Coastal Plain
tributary rivers in this direction. The approximate ratio of Piedmont to Coastal
Plain drainage area is 4:l for Santee watershed, 7:3 for Savannah watershed, and
2:3 for Altamaha watershed. It is thus not surprising to find Piedmont source
kaolinite in greater abundance at the mouth of the Santee River (72%) than at the
mouth of the Savannah (65%) and Altamaha Rivers (60%).The muds of the Pee Dee
River slightly north of the Santee River contain 86% kaolinite. Illite is present in
both the Piedmont and Coastal Plain Rivers in amounts ranging from 2 to 30% with
most samples containing between 10-15%. It does not appear to be diagnostic of
either source.
In contrast to the larger rivers the small Coastal Plain river muds generally
contain 70 to 90% montmorillonite and relatively little kaolinite. Rivers of inter-
mediate size (mixed source) have intermediate amounts of the two physils. I should
add that this relation applies specifically to fresh water muds. The composition of
the physil suites at the drowned river mouths are usually influenced by a tidal
transport of physils from the Continental Shelf (see Estuaries).
Thus, in a subtropical environment (S.E.U.S.)kaolinite is the major physil being
transported to the ocean, whereas at the mouth of the tropical Amazon River,
204

kaolinite is a relatively small amount of the physil suite. In both areas most of the
transported physils are derived from the higher elevation areas of the basins. In the
subtropical region chemical weathering occurs in the mountain areas where eleva-
tions are moderate and local relief is from 100 to 150 m, whereas in the Amazon, the
Andes Mountains have a high elevation and in excess of 4000 m and the weathering
is largely physical. In these two examples a combination of elevation and relief
determines the composition of the transported physil suites. Because of the dif-
ference in discharge the Amazon River delivers more tons of kaolinite to the
Atlantic ocean then do all the rivers of the S.E. United States.
The physil suites carried by the numerous rivers running into Chesapeake Bay
vary in composition but the physil suites contain on the order of 50% illite and
varying amounts of I/S, chlorite, and kaolinite. The physil suite is primarily
controlled by source rock, though a small portion of the I/S, kaolinite and
vermiculitic material is probably derived from soils.
The suspended physils ( < 2 pm) in the upper portion of the St. Lawrence estuary
(St. Lawrence River) of Canada have an average composition of 64% illite, 34%
chlorite, 1.5% kaolinite, and a trace of montmorillonite. Most of the physils were
apparently derived from the Cambro-Ordovician shales and slates along the south-
ern flank of the river. A tributary river (Saguenay), which drains the Canadian
Shield to the north of the river has a high content of illite and little chlorite
(D’Anglejan and Smite, 1973). In this humid cold temperate climate the river
suspend sediments appear to closely reflect the composition of the surrounding
rock.
As one moves north along the east coast of the United States the character of the
source rocks remain unchanged (largely illite-mica and chlorite); the elevation and
relief is similar to that of the southern Appalachians, as is the annual rainfall. The
major difference is a northward decrease in temperature. As a result of the lower
temperature, chemical weathering is less intense and the river physils more closely
reflect the composition of the bed rock than those in the more southern rivers. The
length of weathering time is also a factor. The soils and saprolites of the southern
Appalachians are much older than those to the north.
Along the Arctic coasts of Labrador and Baffin Island the physil suites in near
shore marine sediments largely reflect their local origin and the distribution of
source lithologies. Chemical weathering is negligible (Piper and Slatt, 1977). The
predominant physils are illite ( > 60%) and chlorite ( - 20%) with kaolinite and
montmorillonite generally comprising less than lo%, but locally are more abundant.
The illite-chlorite suite is related to outcrops of metamorphic rocks and kaolinite
and montmorillonite to unmetamorphosed sedimentary rocks. This example and
others indicate kaolinite can be abundant in Arctic muds but it is presumably
always detrital.
The Columbia River in the N.W. United States shows the effects of two
contrasting lithologies on the streams physil suite (Knebel et al., 1968). The
northern portion of the river, above the Grand Coulee Dam, drains a rugged
topography formed on plutonic and metamorphic bedrock. The climate is cool
subhumid to humid; organic litter and organic acids are abundant. The average
205

composition of the river muds is 70% illite, 8% montmorillonite, and 22% chlorite
plus kaolinite. The cool temperature and high relief (physical weathering) are
apparently more of a factor than abundant water and organic acids (chemical
weathering) in determining the composition of the river transported physils. Most of
the remaining portion of the drainage basin is in the basaltic Columbia Plateau.
Relief is low and the climate is arid to semi-arid. Small tributary streams in the
basalt have an average composition of 7% illite, 89% montmorillonite, and 4%
chlorite plus kaolinite. The suspended illitic mountain suite is progressively diluted
by the volcanic derived montmorillonite as the suspended load moves seaward
through the Columbia Plateau. The river muds in the lower reaches of the main river
(Bonneville Dam) have an average composition of 24% illite, 55% montmorillonite,
and 21% chlorite plus kaolinite. This is equivalent to roughly one part mountain
suite to two parts volcanic suite. This ratio is similar to that of the outcrop areas of
the two types of source rock. Even though the plateau climate is relatively dry an
appreciable amount of montmorillonite is being formed from the volcanics. This is
presumably a tribute to the high solubility of volcanic glass.
Konta (1985) was able to study the suspended sediment from 13 of the world’s
major rivers; Mackenzie (NW Canada), St. Lawrence (eastern Canada), Orinoco
and Caroni (northern South America), Parana (SE South America), Nile (NE
Africa), Niger (NW Africa), Orange (southern Africa), Indus, Ganges, Brakmaputra,
and Padma (India region), Waikato (New Zealand). The rivers occur in climatic
regions ranging from arctic to tropical. Samples from each river were collected
periodically over a one to two year period.
Fig. 4-11 shows the average composition of the suspended minerals from the 1 3
rivers. Though Konta concluded that each river has its individual fingerprint, the
physil suites in most of the rivers are quite similar. Illite-mica, along with minor
chlorite and kaolinite, comprise the physil suites of most of the rivers. Kaolinite is
predominant in only one tropical river (Niger) and montmorillonite is present in
significant amounts in only two rivers. Konta devised a measure of mineral maturity
based on the sum of stable minerals (quartz + illite + feldspar + amphiboles +
carbonates) and chemical maturity (5% Al: % Na + Mg + Ca). These maturity values
correlate positively with rainfall and negatively with relief and continentality
(temperature and geographic latitude, with the equatorial region having the lowest
value).
It is evident from the various river studies that there is no simple relation
between the physil suite transported by rivers and climate in the drainage basin.
Relief, elevation, source rock, river length, and flood periodicity can also be major
factors in determining the physil suite transported by rivers.

ESTUARIES
Settling
As physils are carried downstream, they can rather abruptly encounter high
salinity ocean waters (deltas, steep coasts) or the transition can be very gradual. The
206

100-
%

80 -

60-

40-

20-

0-

Fig. 4-1 1. Svnoptic view of average percent composition of crystalline suspended matter in some rivers.
From Konta, 1985. Copyright 19x5 University Hamburg.

latter situation is typical of submerged shorelines where estuaries have developed in


drowned river mouths.
In most rivers the concentration of ions is low and the individual physil plates or
packets have a large diffuse electric double layer. Negatively charged physil particles
are surrounded by a swarm of neutralizing (counter-) ions. The ions and the shell of
water they are in are more or less fixed in the proximity of the particle surface and
move with the particle (Fig. 4-12). The charged surface and the shell of counter-ions
comprise the diffuse or electric double layer. How far the electric double layer
extends depends primarily on the concentration and charge of cations in the water
( o r other liquid). In fresh waters where the cation concentration is low the double
layer conimonly has a thickness in excess of 30 to 70 A. The positively charged
double layers are so thick that the physil particles can not approach each other
closely enough for electrostatic attraction (edge-face charge) or for van der Waals
attraction to occur. The attractive force decreases rapidly with distance. The physils
are dispersed. As the salinity or ion concentration increases, the double layer is
compressed, the physil particles can come close enough to each other so that the
attractive forces are stronger than the repulsive force and coagulation or floccula-
tion occurs. The double layer can also be compressed by adding higher valence
cations. The flocculating power of bivalent ions is approximately 20 to 80 times that
of univalent ions and the flocculating power of trivalent ions is 10 to 100 times that
of bivalent ions. As an increase in salinity causes a decrease in the width of the
double layer, the flocs should contract and become denser. The flocculation process
207

Electric potential
surrcundtrq the
Darticle

E
Rigid l a m
attached to Plane of shear
particle
(stern l q e r )
Bulk of solution

Extent of
diffuse layer
of counterions
Concentration of
positive ions

Concentration
ofnegatiw ions
Fig. 4-12. Concept of the Zeta Potential. From Riddick, 1968. Copyright 1968 Livingston

is, to a large extent, reversible. If the salinity is decreased flocs will disperse (Van
Olphen, 1963; Riddick, 1968).
The Zeta potential is considered to be a measure of particle repulsion. It is the
electric potential in the double layer at the interface (shear plane) between the
double layer and the surrounding liquid. When the Z.P., in millivolts, is relatively
large, either positive or negative, the particle repulsion (dispersion) occurs. As the
ion concentration increases and the Z.P. approaches zero, flocculation occurs.
The negative charge on fresh water physils decreases with increasing ionic
strength. Charge inversion generally occurs at salinities of about 2% and most
sediments have a net positive charge by a salinity of about 6%. Particle size does
not appear to be a factor (Pravdic, 1970). Martin et a / . (1971) found charge reversal
did not occur till salinities of about 20%. The difference is believed to be due to
variations in the nature of organic compounds bound to the physils. Though
flocculation occurs by neutralization of the negative charges and compression of the
electric double layer, if the positive charge becomes large enough the particles will
208

repulse each other and flocculation will not occur. Presumably, in estuaries most of
the flocculation occurs before this state is reached.
The formation of aggregates or flocs is a two-step chemical-physical process.
First, the double layer must be compressed by a change in the water chemistry (also
referred to as particle destabilization). Second, there must be transport to allow
interparticle contact. This may be induced by Brownian motion, differential veloci-
ties of suspended particles (larger or denser particles) and velocity gradients
(turbulence). In most river-estuarine systems turbulence is the major mechanism
causing particle collision, though differential velocity settling can be a significant
mechanism. In addition, colloid aggregation can be caused by cross-linkage by long
chain polymers, e.g. humic and fulvic acids.
For the case where transport is a slower reaction than the rate of destabilization,
the rate of coagulation is given by the product of the rate of transport (frequency of
collision) and the relative stability of the colloids (efficiency of collision). In the case
of partially destabilized suspensions the efficiency of collision decreases for ex-
tended times of reaction and eventually comes practically to a halt. A maximal size
of aggregates is reached (Hann and Stumn, 1970).
In waters of low ionic strength (“fresh water”) the stability of colloids is high and
coagulation occurs slowly (months to years). If the colloids are destabilized to a
small extent, the maximal size of the aggregates attained after an extended time is
not too different from the initial size of the colloid.
In waters of high ionic strength (estuarine and sea water) the stability of
suspended colloids is reduced extensively and rapid coagulation occurs. Maximal
size is reached within a short time compared to the time of settling. The maximal
size of the aggregate increases as the relative stability of the colloid decreases (as the
salinity increases). Also, because of differences in size and surface charge, the
various physils have different stabilities at a given salinity. The calculations of Hahn
and Stumm (1970) indicate the following order of increasing stability: montmoril-
lonite > kaolinite > illite. Illite coagulates at a lower salinity than montmorillonite.
Thus, colloidal theory predicts that as river transported physils come in contact with
waters of increasing salinity, illite should flocculate and settle out first, followed
closely by kaolinite and lastly montmorillonite (differential flocculation).
Experimental studies of differential flocculation and sedimentation has produced
mixed results. Table 4-4 summarizes some of the results obtained by Whitehouse et
a/. (1960), who studied flocculation and settling rates under static conditions. The

Table 4-4
Representative Settling Velocities in Quiescent Sea Water (cm/min). After Whitehouse. et al. (1960).

Physil Salinity (ppt)


0.9 1.8 3.6 10.9 18.1 32.5
Illite 0.89 0.90 1.05 1.10 1.10 1.10
Kaolinite 0.80 0.81 0.81 0.81 0.81 0.81
Mon trnorilloni te 0.0023 0.0036 0.0078 0.041 0.076 0.088
209

lo6
I
t \ 17.5ppt

I
0 10 20 30
Time (min)
Fig. 4-13. Flocculation rate of kaolinite in synthetic estuarine solutions at three salinities. From Edzwald
and O’Melia, 1975. Copyright 1975 The Clay Miner Soc.

relative settling characteristic of the three physils are similar to those obtained from
the theoretical calculations of Hahn and Stumm. They also found that the chlorite
settling rate was similar to that of kaolinite and the vermiculite settling rate, at high
salinities, was similar to that of illite. Humic acids decreased the settling rates of
montmorillonite and kaolinite, particularly at low salinities. The settling rate of illite
was not significantly altered by humic material.
More recent coagulation studies have taken account of the turbulence factor. A
rate equation that describes coagulation under turbulent conditions was developed
by Smoluchowski (1917) and modified by others:
dn 4
-= - -a!@nG
dt I7

where a is the collision efficiency factor, @ is the volume of solid material per unit
volume of solution, n is the concentration of particles at time t, and G is the mean
velocity gradient. For a completely stabilized suspension a approaches zero and for
a completely destabilized suspension a! = 1 (all collisions lead to agglomeration).
Parameters measured are dn/dt (change in number of particles with time), n, and @;
therefore, a! can be calculated. Experiments are commonly conducted by stirring
with a blade reactor or by rotating double cylinders.
The results of experiments by Edzwald and O’Melia (1975), shown in Fig. 4-13
and 4-14, indicate the number of particles remaining in suspension as a function of
time. Illite (Fithian) aggregates at a slower rate than either montmorillonite or
210

Mont

Kaolinite

0 10 20 30 0
Time (min)
Fig. 4-14. Flocculation rate of kaolinite, montmorillonite, and illite at a salinity of 17.5 ppt. From
Edzwald and O’Melia, 1975. Copyright 1975 The Clay Miner SOC.

I I I I I
0 4 8 12 16 3
Salinity (ppt)
Fig. 4-15. Stability values as a function of salinity for kaolinite, montmorillonite, and illite. Physils are
destabilized (flocculated) at higher salinities. From Edzwald and OMelia, 1975. Copyright 1975 The Clay
Miner. SOC.
21 1

1 1 I I
0 4 8 12 16
Salinity (ppt)
Fig. 4-16. Stability value as a function of salinity for the clay size fraction of three sediments from the
Pamlico Estuary. Numbers increase in a seaward direction. From Edzwald and OMelia. 1975. Copyright
1975 The Clay Miner. Soc.

kaolinite. Fig. 4-15 shows the relative stability ( a ) of the three physils as a function
of salinity. Illite is more stable than kaolinite which is more stable than montmoril-
lonite. Thus, the physils should be deposited in inverse order (montmorillonite first,
illite last) in an estuary.
Edzwald and O’Melia’s analyses of three muds from the Pamlico estuary (N.C.)
indicate the stability of the natural muds (Fig. 4-16) is lowest in the upstream, fresh
water part of the estuary (No. 3) and progressively decreases in a seaward direction
(Nos. 12 and 18). Sample No. 3 had a high kaolinite content and No. 18 had a
relatively high illite content. Though the natural samples tend to confirm the
experimental data that kaolinite is less stable than illite, it does not prove that
differential flocculation is the main cause of the physil trend in the estuary as it is
likely that much of the illite (marine) is supplied to the mouth of the estuary by tidal
currents.
The physil settling sequence determined by Edzwald and O’Melia is the opposite
of that determined by Whitehouse et al. (1960). They believe this is because the
latter authors reagitated their sample after each withdrawal (from settling tube),
thus breaking the floccules and effectively eliminating the factor of floccule growth
as a function of time. As Fig. 4-15 shows, all three physils form floccules at low
salinities but the growth rate of the flocs, with increasing salinity, is less for illite
than for the other two clays and the flocs could be transported farther.
212

0.15-

/ /‘
/ I

0.51 / /

01 I I 1 I I
0.3 0.5 1 5 10 20
Solinity %.
Fig. 4-17. Relationship of the collision-efficiency factor ( a ) and salinity for illite. kaolinite and
montmorillonite. From Gibbs, 1983, and Edzwald and OMelia, 1975. Copyright 1983 Soc. Econ. Paleo.
Miner.

Gibbs (1983a) conducted a similar series of turbulent-flocculation experiments


and obtained the results shown in Fig. 4-17. The comparison of the two sets of data
illustrates the problems involved in obtaining consistent results in these experi-
ments. Gibbs’s sequence of physil stabilities is, for a wide range of salinities, the
opposite of that obtained by Edzwald and O’Melia (1975), though Gibbs’s relative
stability values vary as a function of salinity. Gibbs attributes his higher collision
efficiency values ( a ) to the fact that his sample collection technique resulted in less
breakage of flocs. Collision efficiency values increased with decreasing grain size,
though the opposite trend was observed in experiments with natural muds. Experi-
ments with several natural muds showed an increase in coagulation with increasing
salinity and increasing time. Gibbs concluded it was not possible to calculate
reliable collision efficiency factors for natural muds using the data he obtained with
the standard physil samples, because the presence of organic and metallic coatings
on natural muds had not been accounted for. Gibbs (1983b) removed the organic
coating, with Na hypochlorite, from some river muds and found that this increased
the speed of coagulation. The difference in the coagulation rate between coated and
uncoated samples is greatest a low salinities and becomes minor at salinities higher
than about 15%0.
213

In his first study, Gibbs (1983a) observed that coagulation began at salinities as
low as 0.5-1.0%0. Flocs form readily at 3% salinity. The time required for half the
particles to coagulate, in a given experiment, ranged from 15 to 55 minutes with the
time decreasing with increasing salinity. Thus, the physils could be transported
through a considerable salinity range before enough collided to form a sizable floc.
Dyer (1972) found that if physil concentrations are sufficiently high, illite and
kaolinite will flocculate completely by a salinity of about 4%; montmorillonite is
more slowly flocculated over a wide salinity range.
Differential flocculation can be caused in the laboratory but the results are
largely dependent on the experimental techniques. Whether differential flocculation
occurs in nature is less certain. Hunter and Liss (1979) measured the electrophoretic
mobility (velocity of particles per unit of applied electric field) of suspended
material from 4 estuaries. They found that the particles (physils, quartz, calcite,
organisms, etc.) were negatively charged at all salinities, with the charge decreasing
with increasing salinity. This uniformity of charge on the various suspended
particles is believed to be due to surface coatings of organic matter and metal
oxides. Laboratory experiments have not yet duplicated the natural system.
The organic coating material has carboxyl groups (COOH) which readily release
the H + at pH values > 4. The resulting negative charge is only partially neutralized
by the absorption of larger cations thus a permanent negative charge is produced.
Most humic and fulvic acids are negatively charged. They can be bound to
physils in a number of ways. Two common bonding types are referred to as cation
bridges and water bridges. The former process involves the replacement of a water
molecule from the hydration shell of the exchangeable cation (generally polyvalent)
by an oxygen of the anionic organic polymer. The bridge is relatively weak and the
absorbed humic and fulvic material can be displaced by leaching with NaCl or other
salts of monovalent metals (Greenland, 1971). This may account for the decreases in
organic content with increasing salinity in same estuaries. Water bridges form when
an uncharged portion of a humic molecule form hydrogen bonds with the water of
hydration surrounding the exchangeable cations. When a clay-organic complex is
dried, the two components will be held together by van der Waals bonds and the
organic material is difficult to extract.
At a pH < 8, Fe and A1 hydroxides have a positive charge and form a columbic
bond with the negatively charged organic material. The hydroxides may occur on
the physil surfaces. A similar type of bonding occurs at the broken edges of physils
where positive charges may occur. Another type of bonding is referred to ligand
exchange. In this type of absorption, the carboxyl groups in humic and fulvic acid
form a bond with the surface oxygens of the hydroxide layer. The anion is
incorporated with the surface hydroxyl layer. The complex formed between the
organic anions and hydroxides of Fe and A1 is much more stable than the complex
formed by the cation or water bridge mechanism involving physil surfaces.
The presence of humic and fulvic material increases the stability of physil
suspensions. When electrolytes are added to a solution containing excess humic and
fulvic acids and physil-organic complexes, the cations react first with the excess
214

Diameter ( p )
Fig. 4-18. Progressive flocculation of sediment stirred in 3% NaCl shows the final distrihution (19 h) and
the distribution measured every 0.5 h during the first 6 h. 0 h (initial); --- 6 h: ~ . 18 h (final).
From Kronck. 1973. Reprinted by permission from Nature, 24. 348-350. Copyright Macmillan Mag..
Lid.

organic material. This reduces the effect the cations can have on the physil double
layer (Narkis et al., 1968).
Not only do the physils flocculate or aggregate in some fashion when river water
empties into brackish or marine water, but much of the organic material and ions
such as Fe, Mn, Al, P are flocculated. The inorganic ions are largely complexed by
the organic material which, in turn, coats the physils and other minerals preserving
their negative charge (e.g., Sholkovitz, 1976; Hunter, 1983). Johnson (1974) studies
the material at the sediment-water interface in coastal environments and observed
that 69% of the quartz particles were encrusted with organic matter and organic-
mineral aggregates comprised form 13 to 71% of the sediment. Nearly all of the clay
and silt size particles were incorporated in an organic matrix. The organic-mineral
aggregates are largely fecal in origin.

Size

There is not much information on the size of artificially created flocs but
Edzwald and OMelia (1975) used Whitehouse et a/.’s settling rates i n near normal
sea water to calculate equivalent diameters: illite = 14.2 pm, kaolinite = 12.2 pni,
montmorillonite = 4 pm. They considered these values to be on the low side.
Fig. 4-18 shows the results obtained by mixing a fine glacial clay in a 3% NaC‘l
solution. The sample was stirred continually and periodic measurements of particle
size were made using a Coulter Counter. Floc size and sorting systematically
increased with time. It required 18 hours for a stable, well-sorted floc assemblage to
21 5

develop. The small floc size is presumably due to the absence of organic material
and divalent cations (more effectively decrease the size of the electric double layer).
Microscopic observations of flocs from the marine water adjacent to Nova Scotia
(Kranck, 1973) show they appear as irregular aggregates of individual particles.
Most were smooth, subround aggregates. Larger flocs consisted of a low density,
flexible, loose network of grains. Organic material is present in some flocs. The
particles in the individual flocs are poorly sorted. Flocs formed the largest particles
in all samples but also occurred as very small aggregates. Fig. 4-19 to 4-21 (Arnone,
1972) are SEM pictures of flocs from the Satilla River and Estuary of central
Georgia. The river has a relatively high content of dissolved and particulate organic
material. Fig. 4-19 shows an open floc composed of organic material and clay
aggregates; Fig. 4-20 shows a denser packed montmorillonite-organic particle, - 15
pm diameter, that may be a resuspended “mud-floc”; salinity was 0.11%0.Fig. 4-21
is a floc from the mouth of the estuary composed of physils, organic matter, and
diatom fragments. The presence of sulfur (EDX) indicates organic material is
present in all flocs. The flocs as seen in the SEM probably appear much denser than
in the natural state.
Arnone’s data show that the percent of dissolved Fe and Al decreases abruptly in
the interval between a Na concentration of 3 and 50 ppm. Over the same interval
the Fe and A1 content of the suspended particulate material increases from 3.5 to
5% and 6 to 98, respectively, suggesting some of the Fe and A1 is incorporated into
the flocs. The AI/Fe ratio of the flocs remains near 2 throughout much of the
estuary. This value is low for physils, indicating appreciable “free Fe” is present in
the flocs. Neiheisel (1973) found hydrous Fe oxides (lepidocrocite and goethite)
present as coatings on quartz grains and diatoms in the flocs from Delaware Bay. Fe
probably also coats the physils. The flocs should make good starting material for the
formation of glauconite.
The flocs observed in suspension in the Upper Satilla Estuary (Arnone, 1972)
have a honeycomb, hexagonal arrangement of fibers forming chains (Fig. 4-19).
These early formed flocs ( - 3 ppm Na) contain an average of 18% carbon
(approximately 36% fulvic and lower molecular humic materials) along with varying
amounts of kaolinite, montmorillonite, and Fe and A1 complexed with the organic
material. The flocs are small (maximum size <0.5 mm) and oval shaped. The
organic carbon content of the flocs decreases rapidly until the mixing zone of the
salt water wedge (50 ppm Na) is encountered; the floc concentration increases
rapidly in the mixing zone. The flocs are larger (maximum size 120 mm) and have a
more irregular shape. At higher salinities the flocs are smaller. At 50 ppm Na the
-
flocs contain 9%C; the C content then decreases gradually to 3% near the mouth
of the estuary (7000 ppm Na). The decrease in organic C probably indicates some of
the fulvic-humic material has dissolved in the higher pH waters and/or that
additional physils and shell fragments from the sea and marshes were incorporated
into the floc.
The suspended material in the St. Lawrence Estuary contains 5 to 40%organic
material, based on ashing. Ashing probably gives a high value because of the loss of
physil water (DAnglejan and Smith, 1973). There was no systematic change with
216

Fig. 4-19. Mixed organic (upper)-physil (lower) floccule from low salinity portion of Satilla River. Bar
equals 1 pm. From Arnone, 1972.

salinity; however, they had no low salinity samples. The inorganic material is largely
illite and chlorite. Eighty-five percent of the particles are smaller than 7 p m with
mean diameters ranging from 5 to 7 p m at all stations. Near bottom flocs are larger
than surface ones. Sorting is considered to be good and skewness is negative.
Fig. 4-22 shows the size distribution of suspended material from several marine
environments adjacent to Nova Scotia (Kranck, 1973, 1975). The data were ob-
217

Fig. 4-20. Dense flocks composed primarily of montmorilloniteand organic matter in low salinity portion
of Satilla Estuary. Bar equals 1 pm. From Arnone, 1972.

tained with a Coulter Counter. The solid curve shows the size distribution of the
flocs (dominant component) and grains in suspension. The modal size increases with
increasing current velocity. He was unable to produce fully flocculating suspensions
with modal values larger than 4 pm and suggested this was due to the absence of
organic material. The dashed lines show the size distribution of the suspended
material after it is ashed to remove organics and dispersed. The flocculated material
has a fairly symmetrical, near normal distribution. The flocs are apparently quite
stable, as vigorous agitation and storage for several months did not cause any
change in the size distribution curves. The dispersed grains have a less well defined
modal size that is usually shifted towards the finer sizes.
Kranck suggested that, as in thoroughly flocculated suspensions, the flocs settle
uniformly as a front; and, as the particle size is not uniform, the particles must have
218

Fig. 4-21. Floc from mouth of Satilla Estuary containing physils, organic matter, and diatom fragments.
Bar equals 10 pm. From Amone, 1972.

different densities. Fig. 4-22 shows the relation between the grain size mode
(dispersed material) and floc size mode. Where the grain size is fine the flocs are
small but average approximately three times the size of the grains. The differential
decreases with increasing grain size (and velocity) and the grain and floc size modes
are identical at a value of 64 pm (very fine sand). It may be that the grain size mode
represents the size of minerals such as quartz and feldspar. The grain size of the
dispersed physils would have to be appreciably finer than the grain size mode in
order to produce flocs only 2 to 3 times larger, or less, than the grain size mode.
On the basis of a series of settling experiments with natural samples Kranck
constructed the curves in Fig. 4-23. The curves represent the particle size distribu-
tion at a fixed depth (20 cm) as a function of time. The size distribution after a short
settling time (2-4 = 13 to 52 min) has a well developed mode of individual grains
21 9

Diameter (pm)
Fig. 4-22. Typical grain size spectra of particulate matter from coastal waters with high inorganic content.
The solid curves show the natural distributions and the broken curves the defleocculated inorganic
grains. Origin of samples: A, Northumberland Strait; B, Bay of Fundy; C an D, Petpeswick Inlet. From
Kranck. 1973. Reprinted by permilssion from Nature, 24, 348-350. Copyright Macmillan Mag.. Ltd.

(silt) and a flat tail of flocs. With increased settling time (6-8 = 3 to 12 hrs) most of
the grains have settled and the suspended material consists largely of flocs, ap-
proaching a normal distribution, with a broad mode at approximately 4 pm. The

6o

40
t
-
2-4

Y.

20 -

0 ' I 1 I I I I
220

floc modal value is similar to that found in the St. Lawrence Estuary and Chesapeake
Bay.
Volume-size distribution (based on settling rate) of suspended sediments from the
turbid zone of the Chesapeake Bay (Schubel, 1969) are well sorted and have mean
diameters ranging from 2.3 to 8.6 pm, with most values falling between 3 and 7 pm.
Sample diameter increases with depth. The particle size for most samples ranged
from 0.25 to 70 pm. Dispersed and nondispersed samples have the same size
distribution, suggesting there was little if any flocculation. It is possible Schubel’s
dispersion treatment (sonic) did not disperse organic bonded, tight flocs.
In an SEM study of suspended sediments in the turbidity maximum of Chesapeake
Bay, Zabawa (1978) found three types of agglomerates: floccules of mineral grains
arranged in face-to-face, edge-to-face and edge-to-edge modes of grain contact;
fecal pellets (up to 50 pm); grains bound together by mucal slime secreted by
bacteria and other microbial structures. The pellets are commonly the only ag-
glomerates dense enough to settle through the upward vertical currents, however,
flocs become attached to the pellets and are carried to the bottom of the estuary.
Studies by optical microscope (Gibbs et al., 1983) indicated that suspended
samples from the Delaware River were individual inorganic particles with an
average grain size less than 15 pm. In Delaware Bay, where the salinity was 0.3%.
the surface water contained flocs whose modal diameter was 70 pm; the near
bottom flocs had a modal diameter over 120 pm. The size of the flocs in the near
bottom waters decreased progressively to 50-70 pm, down the Bay. The flocs in the
surface water reached a maximum size of approximately 90 pm at a salinity of 0.5%
and then progressively decreased to approximately 12 pm. It is not known whether
the decrease in size is due to settling, breakage, or dilution. The larger flocs are
composed of about 90% water. Perhaps the flocs become smaller seaward because
they contract and become denser. Laboratory experiments by Gibbs (1984) suggest
that floc density decreases as flocs increase in size.

Deposition

Most of the information for the following summary of deposition processes was
obtained from Meade (1969, 1972), Schubel (1969) and Dyer (1972).
Estuaries are classified on the basis of their salinity distribution and water
circulation as highly stratified, moderately stratified and vertically homogeneous.
Estuaries actually have a continuous range of stratification and the degree of
stratification commonly varies throughout the year as river discharge varies. Most
estuaries are moderately stratified.
In stratified estuaries the lighter fresh river water flows seaward on top of a
denser salt water tidal wedge. As the incoming flood-tide is stronger than the
outgoing ebb-tide the net flow is landward. The position of the interface between
these two bodies of water depends on the magnitude of the river flow and the tidal
flow, and upon the geometry of the estuary. In a given estuary the interface moves
22 1

Water flow

transport
-
Suspended matter
-4 -
4 - - -

---, . . . .. ... . . .
od.~g$+Aa'.vPsvd': .. v.:,g;y&5?-"-:t=.b*.
i
L C C

River Ocean
Fig. 4-24. Schematic representation of the relations between water density, water circulation, turbidity
maximum, and sediment accumulation near the landward limit of sea salt in a river-mouth estuary. From
Meade, 1972, Geol. SOC.Amer. Mem. 133, 91-120.

seaward during the river flood stage and landward when the river water volume is
moderate or low (most of the time).
The two opposing currents transport suspended sediment in opposite directions.
The river suspended sediments are obtained by erosion of the drainage basin and
physil suite, in some way, reflects the nature of the source rock. The tidal currents
obtain their suspended load from the off-shore marine environments and scour of
the estuary bottom and banks. The main part of the tidal load very commonly has a
physil suite that was transported by the river. The reworked or resuspended portion
of the suspended load is usually a mixture of material from the two source areas.
The distribution and origin of physils in an estuary is generally complex and the
transported depositional processes can not normally be well understood unless a
large number of samples are collected over a long time interval.
The net non-tidal estuarine circulation tends to the formation of a "sediment
trap" or turbidity maximum in the estuary in the area where the upstream flow of
the tidal currents gradually slows and the net flow is seaward at all depths (Fig.
4-24). As the river water flows seaward over the lower salty layer some of the
particles settle out of the upper river water into the lower layer and are carried
upstream. Presumably some flocculation occurs at this time as the physils settle into
222

the saline layer. Because of the low salinity value necessary to induce flocculations i t
appears that much of the initial flocculation may occur upstream of the “fresh”
water - salt water interface. Most river velocities are high enough that these loose
floccules would be transported downstream past the fresh water -salt water
interface. As some of these flocs settle into the high ionic strength waters the double
layer should contract and the flocs become denser and also smaller.
Many of the particles that settle into the lower layer are carried back upstream
and are deposited near the fresh water - salt water transition (“sediment trap”)
where the velocity of the lower layer is low (Fig. 4-24). Sediment is resuspended
during both incoming and outgoing tides and settles during slack water. I t is also of
interest to note that the “push effect” of the incoming tide can move estuarine
suspended material many kilometers inland of the turbidity maximum.The turbidity
maximum is in part created by the resuspension of the smaller particles in the
bottom sediment during both the incoming and outgoing tides and by vertical
mixing. The vertical mixing causes the smaller particles to be transported from the
lower layer to the upper layer and the process is repeated many times. Conceivably
flocs could grow progressively larger during each recycling until they eventually
become too large to be resuspended. However, this growth process is presumably
counteracted to some extent by the breakage of flocs by turbulence and partial
redispersion of flocs as they move from the lower salty layer to the upper fresher
layer.
Particles with settling velocities much less than the mean vertical mixing veloci-
ties would be concentrated in the upper layer and transported to the lower reaches
of the estuary or out to sea. In many estuaries there appears to be a natural
background of 10-20 mg/l of suspended sediment with an equivalent diameter of
3-4 pm. As the fall velocities are of the same magnitude as the mean velocity of
vertical secondary currents they are in perpetual suspension. If suspended sediment
is present in high enough concentrations and turbulence is moderate, flocs will settle
as layers. In many estuaries, especially at slack water, a “fluid mud” layer floats o n
the bottom. In mid channel these layers are resuspended by flood tides. Local
currents move some of the “fluid mud” into quiet areas outside the main channel.
There the fluid mud consolidates slowly, presumably by the expulsion of interfloc-
cule water and collapse of the flocculate structure. As consolidation proceeds the
moisture content decreases exponentially with depth and there i s a linear increase in
shear strength. The mud becomes a part of the lithologic column. The process is
normally interrupted by “abnormal” episodic floods or tides.
Fig. 4-25 shows the distribution of clay and the water content of the clay mud in
the near-surface sediments of the James River estuary. The turbidity maximum
occurs at about 50 km from the mouth. The water content closely parallels the clay
content. The amount of organic material, in the total sample, decreases systemati-
cally from 3% in the upstream end of the estuary to 0.5%at the mouth. Unlike many
estuaries the site of maximum deposition is not at the inner limits salt water
penetration but in the middle and outer portions of the estuary. In these latter areas
deposition is in the shallow flanks adjacent to the main channel. Nichols (1972)
223

A
75
Water content
- -.
c
50 .- - -.
., . .-.-
?I
25
a
'1. Clay
0

B
-> 400-
E 300.
;200-
I

+ 100-

80

River
70 60
- 50 40 30 20
Distance from mounth
10 0 km

Mouth

Fig. 4-25. Variations along the length of the James River estuary, Virginia. A, water and clay content in
percent; B. oxidation-reduction potential (Eh) in mv; C. hydrogen-ion potential (pH); D, calcium
carbonate content in percent; E, total carbon and organic material in percent. From Nichols, 1972, Geol.
Soc. Amer. Mem. 133.

suggests deposition occurs in these areas because they are near the level between the
lower, more saline layer and the upper layer where there is no net motion.
It is apparent that in larger estuaries a portion of the river suspended load is
trapped and deposited in the estuary. For example, Morton (1972) found that in
Narragansett Bay, Rhode Island, only 10% of the river sediment load was trans-
ported through the estuary to the ocean. Only 9% of the total input (river and shore
erosion) of suspended material to the Chesapeake Bay escapes seaward. Shore
erosion accounts for 13% of the suspended material in the upper bay and 52% in the
middle bay. The nominal rate of sedimentation in the Bay amounts to 3.7 mm/year
224

(Biggs, 1967). In addition, in many estuaries offshore sediment is also moved in and
deposited. However, sediment does escape from estuaries. Most of the previous
discussion describes the situation in an estuary under normal or common flow
conditions. During flood stage the fresh water - salt water boundary can be pushed
completely out of the estuary and the river sediment load can be transported
directly to the sea. As most of the river sediment load is transported by floods, it is
possible that a major portion of the total river load escapes deposition in the
estuaries. For example, the Delaware River carried more sediment in two days of
flooding than it carried in any other full year over a 16-year period. Much of this
sediment may eventually be trapped in the nearby marshes or moved back into the
estuary.
Most of the estuaries that have been studied in any detail are those that are
subjected to dredging. Dredging allows the salt wedge to move farther into an
estuary, thus increasing the depostion of sediment in the estuary. It is believed that
in matured, undredged estuaries the estuary is progressively filled, becomes shal-
lower, and the landward penetration of the salt water wedge systematically de-
creases. A relatively high proportion of the river sediment would be delivered to the
open ocean. Present estuaries are maintained by the continued rise in sea level,
dredging, and periodic floods (Meade, 1972).

Biodeposits

An appreciable portion of the bottom muds in estuaries and marshes are


apparently biodeposits, fecal pellets and pseudofeces. As has been discussed, tidal
velocities are often great enough to prevent the deposition of floccules and ag-
glomeration by filter feeders may play a major role in forming particles large
enough to settle to the bottom.
Filter feeders, such as tunicates, oysters, barnacles, copepods, clams, mussels and
scallops obtain food by filtration of suspended particles, largely physils and organic
material in the 1 to 5 pm range. Some species reject material before ingestion and
eject it as loosely compacted aggregates termed pseudofeces. Ingested particles pass
through the digestive tract, where acid juices may partially decompose the physils,
and are voided as compacted fecal pellets which range in length from 50 to 3,000
Ilm.
The rate of fecal pellet production is dependent on the concentration of the
organisms that produce the pellets and the particles used in their production. Pellet
survival is variable and is possibly related to the effectiveness of the mucous cement.
Horizontal movement of pellets by currents and winnowing of unpelleted material
can produce patchy distribution of pellets in bottom sediments. Further, the pellets
can be broken up by feeding detritivores (Lewis and Syvitski, 1983).
Fecal pellets from various filter feeders vary in composition. Oyster pellets
contain a mixture sand, silt, clay, diatom fragments, sponge spicules and algae.
Physils make up from 70 to 90% of the pellets and organic carbon from 4 to 8%.
225

Barnacle pellets are composed largely of silt and clay (Haven and Morales-Alamo,
1972).
Laboratory studies show that an individual oyster can deposit on average 1.62
grams dry weight of pellets per week (Haven and Morales-Alamo, 1972). As much
as 25% of the bottom sediment in the York River Estuary, Virginia is fecal pellets,
largely from oysters (Haven and Morales-Alamo, 1968). Moore (1936) found that in
the Clyde Sea Culumus and euphuusice deposited fecal pellets at the rate of 33.4
milligrams per square centimeter per week. Verway (1952) calculated that mussels in
the Wadden Sea deposited between 25,000 and 175,000 metric tons of detritus
annually. Rhoades (1963) estimated that the clam Voldia, which constituted less
than 10%of the bottom fauna, was capable of reworking sediment in Buzzards Bay
and Long Island Sound faster than it accumulated. Briggs and Howell (1984)
calculated that the zooplankton and filter-feeding bivalves in Delaware Bay filter a
volume of water equal to that of Delaware Bay 3 to 5 times per year. They deposit
4 X 10" kg of feces, which settles to the bottom. This mass of feces, consisting of
phytoplankton and suspended sediment, is 200 times greater than the annual fluvial
input of sediments. Biotic filtration in some other estuaries is even larger. Not only
are large amounts of physil and organic material reworked and probably chemically
modified by the filter feeders, but they are compacted into pellets with a settling
velocity equivalent to very fine to fine grained quartz grains. Thus, much of the
clay-organic pellets are going to be deposited in areas of relatively high velocity
currents where clay physils and loose floccules would not normally be deposited.
The biota affects the cohesiveness and erodibility of estuarine and lacustrine
muds. In addition to larger animals and plants, most muds contain an abundance of
microorganisms and microfauna (pass 0.5 mm sieve and retained on 63 p m sieve).
The small organisms grow fast and are the first to utilize organic detritus that falls
to the sediment-water interface. Where sufficient light is available, microalgae such
as diatoms, dinoflagellates, and bluegreens will grow. The shear strength of sedi-
ments has been positively correlated to the organic content, which in turn correlates
with the number of bacteria. Reduced erodibility is thought to be, in a large part,
due to adhesive substances secreted by bacteria and microfauna and the filaments of
the bacteria (Montague, 1984).
SEM studies of the suspended sediments in the turbidity maximum of Chesapeake
Bay showed the presence of fecal pellets and angular flocs in addition to fine
background material of small flocs (Zabawa, 1978). Microbial structures, algae,
fungi and bacteria commonly colonize flocs and pellets. These organisms secrete
mucal slime webbing, which stabilizes the network of mineral grains and acts as a
glue to bond minerals which come in contact with the host complex (Paerl, 1975).
Zabawa noted that the resuspended bottom sediment (by tidal currents) contained a
relatively large number of pellets to which flocs were attached, presumably by the
mucal slime. Thus, flocculation occurs but most of the particles large and dense
enough to settle in this turbid area are fecal pellets.
A simpler model of physil sedimentation in an estuarine environment is provided
by a study of Howe Sound (Syvitski and Murray, 1981). A similar study was made
of Bute inlet, a fiord in the same general area (Syvitski et af.,1985). Howe Sound is
226

a fiord in southwestern British Columbia fed by glacial melt water carrying a high
content of glacial flour. A sill that occurs 17 km from the head of the fiord
effectively prevents sediment being transported from the sea. There are no lateral
currents and no shallow water, coastal environments. Thus, the fiord is an example
of a one-sediment source with deposition restricted to deposition in a deep, narrow
( 3 km) trough.
Glacial flour enters the surface layer of the fiord as a fresh-water sediment plume
which moves quickly down the inlet while gradually mixing with the underlying sea
water. During times of high river discharge, over 50% of the initial suspended load
in the surface layer is transported out of the fiord. The concentration of the surface
suspended load decreases linearly with distance from the mouth as salinity in-
creases.
The particles that enter the fiord are individual physil plates and other silicate
grains, silt-size grains of quartz, feldspar and hornblende which contain physil
particles on the surface, and small clay clasts, commonly spherical (small river mud
balls). As the surface layer mixes with the marine water, flocculation begins. Physil
plates, < 20 pm, flocculate edge-to-edge. The initial floc (10 to 70 pm) density is so
low that there is no appreciable gravity settling, but some are forced down by
diffusion and mixing of the fresh and marine waters. The flocs attract other
particles, including non-platy minerals and organic detritus. “Organic cohesive
forces” attract the organic detritus, and organic material (i.e., bacteria, mucilage)
bind the flocs into more stable agglomerates. Fig. 4-26, 4-27, and 4-28 show the
nature of the suspended material as a function of depth and distance seaward of the
river mouth.
Zooplankton graze on the flocs and agglomerates and produce fecal pellets (90%
inorganic material, largely physils) which settle to the bottom relatively rapidly. The
larger flocs, whch escape the zooplankton, increase in size with depth and develop
face-to-face and face-to-edge particle orientation, eventually forming stable three-di-
mensional flocs. The smaller flocs retain their two-dimensionality with depth and
grow into snake-like flocs. These flocs can form large vertically aligned streamers
(observed from submersible). Some of the agglomerates, colonized by bacteria,
develop large, formless three-dimensional blobs (also observed from submersible).
This careful study, in an uncomplicated environment, showed that physils
delivered by river to a marine environment form flocs. The floc density can be
increased by organic activity to the stage that they are dense enough to settle to the
bottom (agglomerates); others form large flocs that may remain in suspension for
long periods of time (streamers and blobs). Zooplankton are effective in sweeping
physils from suspension and concentrating them into relatively dense fecal pellets.
Getting physils to the bottom is a complicated process, and the relative importance
of the various processes can vary with environmental conditions.
Syvitski and Murray (1981) also noted that water turbidity was no indication of
the downward flux of particles. Discrete particles or low density flocs may be
abundant, but if the other factors which produce dense aggregates are not operative
the downward flux can be minor.
Wells (1987) observed that in a coastal barrier island setting (Cape Lookout
227

Fig. 4-26. Scanning electron photomicrographs of samples taken throughout the water column 17 km
seaward of the Homathko River mouth, British Columbia. (A) 0 m-fine suspended particles; (B)
0 m-surface water floccule; (C) 5 m-floc size increasing; (D) 50 m-large compact agglomerate: (E)
50 m-copepod mineral-rich fecal pellet; (F) 250 m-flocs and silt grains. Scale in micrometres. between
white bars, is given below each picture. From Syvitski et al., 1985. Copyright 1985 Elsevier Pub. Co.
Courtesy J.P.M. Syvitski.
228

Fig. 4-27. Scanning electron photomicrographs of samples taken through the water column 50 km
seaward of the Homathko River mouth: (A) 0 m-surface water aggregates; (B) 0 m-surface water
aggregates and filaments; (C) 50 m-decreased floc size, chain diatom; (D) 25 m-increased biogenic
components in flocs; (E) 250 m-larger floccule, floccules linked by bacterial filaments; (F) 500 m-sus-
pensates dominated by bacteria. Scale in micrometres between white bars is given below each picture.
From Syvitski et al., 1985. Copyright 1985 Elsevier Pub. Co.
229

Fig. 428. Scanning electron photomcrographs ot samples taken throughout the water column 75 km
seaward of the Homathko River mouth: (A) 0 m-dinoflagellate matting; (B) enlargement of (A) showing
mineral grains attached to matting; (C) 20 m-large mucous coating binding grains; (D) 50 m-fecal
pellet, and agglomerate (E) showing biogenic debris; (F) 250 m-submicron size particles in agglomerate.
Scale in micrometres between white bars is given below each picture. From Syvitski et al., 1985.
Copyright 1985 Elsevier Pub. Co.
230

Bight, N.C.) mud was accumulating at the rate of 10 cm/yr. despite the fact that
particle density in the overlying water was low and current velocity was high. Diver
observation indicate the sediment is settling as large (up to 1 cm) agglomerates of
marine snow. These agglomerates are bound by bacterial mucilage which also glues
the agglomerates to the bottom limiting the amount of resuspension. Turbulence not
only does not break the agglomerates but allows them to grow by coming in contact
with other particles that can be entrapped by the mucilage.
The studies of Wells (1987) and Syvitski and Murray (1981) indicate bacterial
binding, as well as pellet formation, can be significant factor in the deposition of
physils in relatively high energy environments.

Physils in estuaries

The physil suites of the muds in the estuaries along the east coast of the United
States all show some change when they are traced from the fresh water river muds
to the mouth of the estuary. The change in the physil suites with increasing salinity
has been ascribed to diagenetic or authigenic modification of the physils, differential
flocculation and, more recently, to the interaction of the river source and the marine
source (tidal). The latter process appears to be the more significant. Furthermore,
the physil suites apparently do not vary with the grain size of the sediments.
Petrie, Jr. (1972) analyzed samples from along a series of lateral transverses in the
Neuse River estuary and found that the physil suites of sediments ranging from 99
to 5% sand were similar. Years ago I analyzed sand-mud sample pairs from the
Brazos River, Galveston Bay and Galveston Island, Texas, and found no significant
difference in the physil suites. At least locally current velocity does not appear to be
an effective mechanism for sorting clay size physils.
We will examine the distribution of physils in estuaries located along the east and
west coasts of North America and the Gulf of Mexico, starting with the northeast-
ern North America and moving clockwise.
The physil suites in the Arctic fiords (9) along the eastern coast of Baffin Island
are relatively uniform, containing 80 to 90 + % illite (with minor I/S), a maximum
of 13% chlorite, 11%kaolinite and 12% smectite. A minor amount of corrensite is
present in two fiords. The Sandy and silty muds range from well-graded turbidites
to bioturbated muds (Hein and Longstaffe, 1983). The physil suite reflects the low
weatering intensity of the arctic climate.
In a survey study of the coastal sediments along the eastern margin of the United
States, Hathaway (1972) showed that the estuarine physil suites ( < 2 pm), in
general, reflected the north-south weathering trend discussed in the section on
rivers. The suites are modified by the influx of physils from the continental shelf.
Illite and chlorite are the dominant physils in the estuaries extending from Maine to
Chesapeake Bay. The southern estuaries are filled with various mixtures of kaolinite
and montmorillonite depending on the relative abundance of Piedmont (kaolinite)
and Coastal Plain (montmorillonite) rocks in the drainage basin. Dioctahedral
vermiculite is present from Chesapeake Bay south. Fig. 4-29 shows the distrubution
231

2s

Fig. 4-29. Map showing distribution of illite in western Atlantic ( < 2pm, carbonate-free fraction). From
Hathaway, 1972, Geo. Soc. Amer. Mem. 133, 296-316.

of illite along the eastern coast of the United States. Chlorite shows a similar
distribution pattern but is less abundant (absent to 4 parts in 10). Fig. 4-30 shows
the distribution of kaolinite. The regional distribution of montmorillonite is similar
to that of kaolinite.
Illite and chlorite are the dominant physils in estuaries as far south as Chesapeake
Bay. Though the northernmost rivers carry enough of these physils to account for
the amount in the estuaries, the relative amount carried by rivers progressively
decreases southward and most of the estuaries south of Long Island Sound contain
a higher proportion of illite, and commonly chlorite, than is present in the river
physil suite.
232

Fig. 4-30. Map showing distribution of kaolinite in western Atlantic ( < 2pm. carbonate-free fraction).
From Hathaway, 1972, Geol. SOC.Amer. Mem. 133, 293-316.

A reasonable explanation proposed by Hathaway (1972) is that during the


Pleistocene glacial melt waters carried rock flour containing unweathered illite and
chlorite to the outer edge of the Continental Shelf. As the sea level rose and the
shoreline transgressed the shelf, physils were winnowed from the glacial deposits
and marine currents transported them landward and southwest along the coastline.
Tidal currents then pushed some of the physils and sands into the estuaries,
increasing the illite and chlorite content in the seaward portion of the estuaries.
The < 44 pm fraction of the shoal deposits of the Delaware River contain a
physil suite composed of 61% illite, 18% chlorite, 13% kaolinite and 8% montmoril-
lonite (Neiheisal, 1973). The physil suite is compatible with a river source. Seaward,
in the Delaware Bay, the average composition is 67% illite, 26% chlorite, 7%
233

kaolinite and 6%montmorillonite. It was not established if the difference was due to
the influx of marine suspended material or shore erosion of Coastal Plain sediments.
In the Rappahannoch River and estuary, which empties into the Chesapeake Bay,
the change in the physil suite with an increase in salinity includes a decrease in
kaolinite, an increase in the “crystallinity” of illite and the abrupt occurrence of
chlorite and feldspar in the saline portion of the estuary (Nelson, 1960). Farther
south, in the James River estuary, the trend is similar. Bottom samples from the
James River estuary and the Chesapeake Bay (Feuillet and Fleischer, 1980) nicely
show the type of profile that develops when both a river and a marine source
contribute physils. Kaolinite and dioctahedral vermiculite decrease seaward as illite,
chlorite and montmorillonite increase. As seen in Fig. 4-31 the rate of increase or
decrease for the various physils changes about midway down the estuary, a few
kilometers below Hog Island. The James River is a partially mixed estuary. The
tidal wedge extends upstream to near Hog Island and is overridden by the lighter
river water. The null point and turbidity maximum occur several kilometers down-
stream of Hog Island, the same location where the rate of increase and decrease in
the relative abundance of the various physils occurs.
The distribution of the physils strongly suggests the downstream change is due to
the dilution of the river physil suite, relatively high in kaolinite and vermiculite, by a
marine suite with a relatively high content of illite and chlorite. This distribution
pattern is not evidence that flocculation has not occurred. Flocs normally have a
low density and a range of sizes. Once formed, they need not settle immediately but
can be transported by relatively slow currents for a considerable distance.
Fig. 4-34 shows the distribution of physils in the Pamlico River and estuary of
North Carolina (Edzwald and OMelia, 1975). Kaolinite decreases and illite in-
creases. No change was reported for chlorite and “chlorite-like intergrades.” The
trend was believed to be due to differential flocculation. However, Nelson (1973)
found both well crystallized illite and chlorite were more abundant in the tidal
portion of the estuary and concluded they had a marine origin. A similar trend
occurs in the nearby Neuse River and estuary (Griffin and Ingram, 1955).
Hathaway (1972) found that dioctahedral vermiculite was a significant compo-
nent of the river physil suites from the Chesapeake Bay area to southern Georgia. It
is most abundant in the Neuse River and Santee River areas (Neiheisal and Weaver,
1967). This is presumably the material that earlier workers referred to as diagenetic
chlorite. The dioctahedral vermiculite is a weathering product of muscovite and
illite. Its absence north of Chesapeake Bay is presumably due to less intense
weathering in the colder climate. The vermiculite content of the shelf and estuaries
is commonly less than that in the adjacent rivers. I t is probable that some of the
dioctahedral vermiculite has a high enough layer charge to extract K from the sea
water and revert to a mica or illite (Weaver, 1958).
From North Carolina southward, the major rivers, draining the highly weathered
Piedmont, carry a clay suite composed primarily of kaolinite. The physil suite of the
Pee Dee River of South Carolina contains 80 to 90% kaolinite. The kaolinite values
systematically decrease southward to central Georgia where the Altamaha River
physil suite contains 60% kaolinite. The decrease in kaolinite is related to an
234

Distance downstream (km)

40

20

--
30- Dioctahedml vermiculite
I
20 - I
.. .- . ....-
I
10- - -.:.
*--I.. -r. -. a-

10
c
5

.-
c
40

20

10

30 t Montmorillonite I I

Fig. 4-31. Relative abundances of physils in James River estuary. Salinities shown on ordinate arc
average autumn surface values. Least-square fits are calculated on basis of two line segments. Point o f
greatest change in slope, indicated by dashed vertical line. divides the line segments. Note scale change
for relative percent. From Feuillet and Fleischer, 1980. Copyright 1980 SCK.Econ. Paleo. Miner.
235

Chlorite and 146: 40%


Mont ~ 5 %

6 20-

8 16 24
Nautical miles downstream
Fig. 4-32. Clay mineral composition of the Pamlico sediments downstream from the railroad bridge at
Washington, N.C. From Edzwald and OMelia, 1985. Copyright 1975 Clay. Miner. SOC.

increase in the amount of Coastal Plain sediments in the drainage basins (Neiheisal
and Weaver, 1967). The suspended sediment of the small Coastal Plain rivers
commonly contains more than 70% montmorillonite.
An excellent example of an estuary with its drainage basin entirely in the Coastal
Plain is the Coosawatchie River, Broad River, Port Royal Sound in southern South
Carolina (Neiheisal and Weaver, 1967). The Coosawatchie River muds ( < 2 p m
fraction) contain 75 to 95% montmorillonite and minor amounts of illite, sepiolite
and palygorskite. Kaolinite is present in amounts of 1 to 2%. The muds of the tidal
Broad River contain 20 to 37% kaolinite, 9 to 37% illite and 37 to 69% montmoril-
lonite. Fig. 4-33 shows the kaolinite/montmorillonite (K/M) ratio values for the
two rivers and the offshore area north of the river mouth and along the east side of
the estuary. Values are relatively low at the head and along the west side of the
estuary and to the south of the river mouth. Illite has a similar distribution,
averaging 28%on the east side of the estuary and 12% on the west side. The offshore
samples north of the estuary mouth contain an average of 33% illite, the samples
south of the mouth 19% and the samples some distance offshore of the mouth 34%.
The percent illite is closely related to the K/M ratio. It is apparent that littoral
currents and flood tides carry a relatively kaolinite and illite rich physil suite to the
east side of the estuary. Much of this material is deposited there. The remaining
material mixes with the montmorillonite-rich suite from the Coosawatchie River and
during ebb flow deposits the mixed physil suite along the west flank of the estuary
and south of the river mouth. The segregation within the estuary is believed to be
due to Coriolis force. Similar patterns have been observed in other estuaries.
The composition of the physil suite (33% illite, 35% kaolinite and 32% montmoril-
lonite) north of the estuary mouth is similar to that in the Santee Delta and
Charleston Delta (diverted Santee), 65 km to the north, suggesting the Santee River
is the ultimate source of much of the physil material being transported into the
Broad River estuary.
Please note, in regard to interpreting the paleogeography of ancient sediments,
that, in the situation just described, if one were using the K/M ratio to determine
the shoreward direction, one would most likely make the wrong interpretation.
236

%. 5 miles

- > 0.80k / m
Current

t
N

__

kaolinite-rich physil suite from the Ocean into the estuary. From Neiheisel and Weaver, 1967. Copyright
1967 S o c . Econ. Paleo. Miner.

A similar complex distribution pattern also exists in Brunswick Harbor, central


Georgia (Neiheisal and Weaver, 1967). The estuary is formed by the Coastal Plain
Turtle and Brunswick Rivers. The river muds have a K/M ratio of approximately
0.5 and contain 10% illite (Fig. 4-34). The K/M ratios in the shoal area and at the
harbor entrance are in the range of 0.8 to 1.0. The illite content is 10 to 12%. Most
of the offshore sediments have a K/M ratio less than 0.8 and an average illite
content of 17%.The offshore area does not appear to be the principal source of the
shoal material. The suspended load of the large Altamaha River, approximately 20
kni to the north, has a K/M ratio larger than 1.0 and has a low illite content. The
two rivers are connected by a salt water coastal stream (Mackay River) that is
aligned parallel to the coast in back of a barrier island complex (St. Simons Island).
Engineering reports indicate that when the Altamaha River is at flood stage,
appreciable water moves into the Mackay River and flows south towards Brunswick
Harbor. Flood tides presumably push much of this water into the Brunswick Harbor
area where some of the Altamaha River sediment, along with offshore suspended
material, settles out during slack tide. The K/M ratio, 1.5. of the Mackay River
muds confirms that suspended physils from the Altamaha have been transported
237

\ / I

Fig. 4-34. Kaolinite/montmorillonite ratio of muds near Brunswick, Central Georgia. The high K/M
ratio in the harbor shoal muds indicates the Altamaha River was a major source of physils. Neiheisel and
Weaver, 1967. Copyright 1967 Soc. Econ. Paleo. Miner.

southward through the coastal river. In this example, the muds deposited in the
small estuary were derived from three different sources and transported by three
different current systems.
Fig. 4-35 illustrates the problem of relating the physil suite of the suspended
sediment load to that of the bottom sediments. The Satilla River estuary is a
moderately sized Coastal Plain river located in south Georgia. The suspended values
are an average of three samples collected during different months of the year.
Kaolinite and montmorillonite comprise 90% of the physil suite ( < 2 pm). In the
fresh water portion of the river, the K/M ratio is approximately two (60% kaolinite,
30% montmorillonite). With increasing salinity, the kaolinite and montmorillonite
values converge, reaching a K/M ratio of one at a CI concentration of 2500 ppm.
The ratio remains near one throughout the remainder of the river. This pattern
could be interpreted as the result of differential flocculation of kaolinite at the head
of the estuary. However, the composition of the bottom sediments indicates this is
not the case. Most of the bottom sediments have a K/M ratio near one, similar to
that of the suspended physil suite in the high salinity area. Much of the downstream
suspended material is presumably resuspended bottom and shore material. The
K/M ratio drops to 0.5 at stations 7 and 6 coinciding with a significant increase in
238

0.0035 0.013 0.06 0.16 0.23 028 0.44 0.72 0.86


% ChlWlde
0 suspended
0 Bottom

0 0

/

A

0 @ 8
Y
Offshore
5 krn
L
10 9 8 7 6 5 4 3 2 1 Stotions
Fig. 4-35.Kaolinite/montmorillonite ratio of suspended and bottom physils ( < 2 pm) in Satilla River
and Estuary. Data replotted from Windom et al., 1971.

salinity. This is probably the area where the turbidity maximum occurs. The
offshore marine physil suite has a K/M ratio of approximately 0.8. The situation
suggests that the smaller amount of suspended material transported by the Satilla
River is dwarfed by the amount transported into the estuary from offshore. The
offshore physils with a K/M ratio of 0.8 and Satilla fresh water physils with a K/M
ratio of 2.0 presumably combine to produce the K/M ratio of 1.0 in the estuary.
The K/M low near the head of the tidal wedge may indicate the montmorillonite
is preferentially flocculated at these low salinity values. The preferential flocculation
may be related to the rapid flocculation of organic material which occurs in the
same general area.
Another observation that needs to be explained is the reason for the relatively
high kaolinite content of the Satilla, a Coastal Plain River. Most of the drainage
area of the Satilla River is underlain by Pleistocene-Pliocene sands and gravels.
Relatively intense weathering in these porous sedimemts has altered much of the
montmorillonite to kaolinte.
Charleston Harbor, South Carolina, affords an excellent example of what en-
gineers can do to a small estuary and affect the distribution of physils (Neiheisal
and Weaver, 1967).
Prior to 1942, Charleston Harbor was a salt water estuary receiving fresh water
discharge from the Ashley, Wando, and Cooper Rivers which drained but 1,188
square miles of low relief Coastal Plain formations. In February 1942. the Santee
River was diverted to the Cooper River (for hydroelectric generation), thereby
adding 14.512 square miles of Piedmont and Coastal Plain drainage to the Charles-
ton Harbor watershed. The increased discharge in the upper Cooper River caused
scour of bed and banks while additional sediment was contributed from the
Piedmont source Santee River watershed. The added discharge of fresh water into
the salt water estuary created a “mixed estuary”, i.e., the creation of a salt water
wedge mechanism which served to trap much of the increased sediment load of silt
and clay in the harbor as shoals. Historic records indicate that sediment deposition
was but 10-15% of that presently occurring in the harbor. The amount o f material
239

Fresh water flow


I
-Diverted Santee watershed-14.512~m t y Pre-diverson watershed-1 188 Sq mi 4
Watershed runoff ++-Tidal
1
cycle control
Dam control I
Mostly fresh water +Pred
-
-3
I

I and ebb tlow ,bottom flood


I I I
I Lakc I,-- TailraceCanal I

I , ,
0 10 20 30 40 50 60 70 80
Distance in miles for sediment sample locations along profile
1

Kaolinite and montmorillonite as per cent clay- mineral Suite I


I
I I
I I

i Kaolinite
I
; 0 2
I
!,o- A Montmorillonite I I
? 0 :Suspended kaolinite I I +
in Tailrace Canal I I I
I !!

?"'
>
:20-

ia ~

I J r
0" ' ' ,++,= ,= , I , I
I ,

0 10 20 30 40 50 60 70 00
Distance in miles downstream for harpoon samples in Santee -Cooper system

Fig. 4-36. Distribution of diagnostic clay and sand size minerals in physiographic unit comprising
Charleston Harbor watershed. From Neiheisel and Weaver. 1967. Copyright 1967 Soc. Econ. Paleo.
Miner.

dredged from the ship channel increased from 120,000 cubic yards prior to 1942 to
over 7,000,000 cubic yards in 1961.
Fig. 4-36 shows the kaolinite and montmorillonite content ( < 55 pm) of bottom
muds in an area extending from the diverted Santee, through Charleston Harbor to
the Continental Shelf. The Santee River has a physil suite composed of approxi-
mately 80% kaolinite and roughly equal amounts of illite and chlorite-vermiculite.
240

After the Santee River suspended load, with a K/M ratio > 10, leaves Lake
Moultrie the ratio decreases to about 4. The decrease is due to hydraulic scour of
the montmorillonite-rich banks. Scour is enhanced by the large increase in water
volume caused by the diversion. The bottom muds of the Cooper River have an
average K/M ratio of about 0.8. The difference between the K/M ratio of the
suspended load and the botton muds suggests much of the scoured material is
deposited in the river.
In Charleston Harbor and at the harbor mouth, the high kaolinite areas roughly
coincide with the shoal areas. In most samples, more than 80% of the shoal material
is finer than 44 pm. The average K/M value of southwestern Charleston Harbor
muds, where most of the shoaling occurs, is 1.10 (range 0.5 to 3.5); whereas the
sediments in the northeastern part of the harbor, where depostion is minor, have a
value of 0.68 (range 0.4 to 1.0). The latter value is similar to that for the
prediversion muds. Cores from the dredged channel west of Drum Island have a
much lower kaolinite content at the base than the top. The base of the core is
presumably prediversion sediment. Offshore the K/M value averages 1.55. This is
similar to the value of 1.60 for the original Santee Delta sediments 75 km to the
northeast.
Suspended samples, taken over a full tidal cycle, in the harbor have average
K/M ratios ranging from 1.6 to 4.7. For stations at the tidal inlet and offshore
values range from 1.2 to 2.1. During the spring high run-off the contrast is even
larger. The surface waters in the harbor have K/M ratios of 4.7 (flood tide) and 2.7
(ebb tide). The harbor mouth and offshore values are 1.3 to 2.0 (flood) and 1.9 to
2.0 (ebb). This indicates much of the river transported kaolinite remains in the
harbor. Salinity values are lowest in the western part of the harbor, where the major
shoaling occurs. This indicates the Cooper River discharge is directed to this area.
At an offshore station north of the harbor entrance the bottom suspended load is
greater at flood than ebb tide. At a station south of the harbor entrance the
situation is reversed and the concentration is greater than at the northern station.
This suggests that more bottom suspended sediment is leaving the harbor, being
moved southward by the longshore currents, than is entering it. The K/M values at
the north and south of the harbor entrance are similar, but the southern suspend
physil suite contains only half as much illite (11% vs. 23%) as the northern suite.
Most illite values inside the harbor range between 20 and 30%. This suggests some
illite is preferentially trapped inside the harbor.
Fig. 4-37 illustrafes the variability of the suspended physil suite throughout a
tidal cycle. The station was at the mouth of Charleston Harbor. The K/M ratio
values vary considerably. Vertically, at a fixed time, the ratios can differ by a factor
of two. The ratio at'a specific sample depth can vary by a factor of three over a
period of several hours. However, the ratio values have a well developed mode with
50% of the values ranging from 1.5 to 2.0. The average ratio value for the ebb
samples is 1.96 and for the flood samples 2.13. The average illite values are 23% and
25%, respectively. Spot sampling can provide atypical samples. In this example
approximately 50% of the samples have values significantly lower or higher than the
mean or modal values.
241

Time ( h )
Fig. 4-37. Suspended sample measurements in Charleston Harbor on 20 January 1966 during spring tide.
After Neiheisel and Weaver, 1967.

There appears to be relatively little information on the physil suite of Recent


shallow water carbonates. There is some data on the carbonate muds of Florida Bay
(Weaver, 1960; Manker and Griffin, 1971). The bay is separated from the Atlantic
Ocean by the Florida Keys reef tract. The physils in the carbonate residues (0.1%)of
rocks from the Atlantic are largely illite, chlorite and talc. The reef tract is a
relatively effective barrier and the physils in Florida Bay residues (1 to 8%) are
largely a variety of chloritic montmorillonite. In the western part of the bay, normal
montmorillonitic material remains predominant; however, progressively eastward it
expands less when treated with glycol and does not contract when heated. Scattered
islands of Mg hydroxide are presumably present in the interlayer space. Close to the
reefs the physil suite is a mixture of material from the two sources.
The bay is a semi-evaporate basin at certain times of the year. The high salinity
may have caused Mg hydroxide to precipitate in the interlayer positions of the
montmorillonite. However, the Pleistocene carbonates that form the north flank
(Everglades) of the bay contain a chloritic montmorillonite which could be the
source of the material in the bay. The Pleistocene physils probably formed under
environmental conditions similar to those that now exist in the Florida Bay region.
“Smectites” which will not collapse to -
lOA at -
200°C have been reported in
the “estuaries” of the Gulf Coast, south Florida and the southern portion of the east
coast of the United States. Most authors believe this is due to the precipitation of
Mg hydroxide islands between layers of montmorillonites and that a chloritic
mineral is being formed. These physils collapse at temperatures near 400”C, in
comparison to a collapse temperature -
600°C for chlorite. There is no question
that something other than the exchangeable cations and water is present in the
interlayer position of the smectites. The interlayer material is most likely Mg
242

5-
*
I
0

1-
- - - - - - Kaalinite-tr.
- - - -- -- - 8

Guad
rivter
Guod
delta San Antonio, Mesquite, Aransas bays %t @3
/flats
Fig. 4-38. Relation of clay mineral composition to geographical divisions in the bays near Rockport,
Texas. From Grim and Johns, 1954. Copyright 1954 Natl. Acad. Sci.

hydroxide but it could be organic material and/or Fe hydroxides. Surprisingly, this


has not been convincingly established.
An early study of physil muds from San Antonio Bay, Mesquite Bay and Aransas
Bay, Texas, produced the data shown in Fig. 4-38. Grim and Johns (1954) interpre-
ted the seaward decrease in montmorillonite and increase in illite and chlorite as
being due to diagenesis. The similarity of the offshore physil suite to that in the bays
at a considerable distance from the Guadalupe River source suggests that the trend
is due to systematic mixing of physils from the river with those from the inner shelf
(tidal). Grim and Johns also noted that samples from the barrier islands contained
more illite than those from the bays or shelf. I found a similar situation on
Galveston Island. The barrier island muds contained twice as much illite (26%)as
the bay (13%) and offshore (12%) samples. This suggests some winnowing has
occurred. Similar concentrations of illite occur in some ancient beach deposits.
Russell (1970) and Drever (1971) report that the physil suite in Banderas Bay,
west coast of Mexico, is similar to the suite transported by the Rio Ameca, the only
river emptying into the bay. However, they report the physil suite at the river mouth
is composed of 60% montmorillonite, 35% kaolinite/halloysite and 5% illite; the
average values for the bay are 52% montmorillonite, 28% kaolinite and 20% illite.
The difference, particularly for illite, suggests there has been an appreciable contri-
bution from the sea. In bay sediments with a high content of H,S, Drever found
that with depth the non-exchangable Mg increased by 0.2 to 0.4%and Fe decreased
by a similar amount. He suggested that under reducing conditions Mg replaced Fe
in the montmorillonite octahedral sheet. The released Fe forms pyrite.
A study of San Francisco Bay, California, (Fig. 4-39) indicates some of the
complications involved in relating the suspended physil suite to that of the bottom
muds (Knebel et al., 1977). The San Francisco Bay system is a near right angle
estuary. The northern leg (east-west) is fed by Sacremento-San Joaquin Rivers that
supply most of the sediment to the bay. More than 80% of the sediments are
delivered in the winter. This leg is a well- to partially well-mixed estuary. The
southern leg (north-south), in contrast, receives only a small volume of riverborne
243

p Winter
2 3t Spring
g

b p
summer
I Fall

1.2- 0 Bottom mud

1.0- 0
T
s
Y
- 36
26
.- 0.8-
G.-
6
2 ;:
0.6dr
0
Range of
turbidity m x
0.4-
L
RV 3 6 9 12 14 17 21 24 27 30 32 36 Stations
Northern reach Southern reach
Fig. 4-39. Illite/montmorillonite + chlorite + kaolinite ratio of suspended and bottom sediments in San
Francisco Bay, California. Illite appears to be preferentially suspended in the southern reaches of the
bay. Plot of data from Knebel et al., 1977.

sediments from small local streams and no tidal wedge is present. Most of the
suspended sediment is due to tidal currents and wind generated waves.
The physil suites (2-20 pm) of the bottom muds have a relatively uniform
composition throughout the bay system. The suite is composed of roughly equal
amounts of illite and chlorite plus kaolinites and slightly less montmorillonite
(average: illite 35%, chlorite plus kaolinite 45%, montmorillonite 20%). The average
suspended suite from 13 stations is composed of 42% illite, 36% chlorite plus
kaolinite and 22% montmorillonite.
Fig. 4-39, which is a replot of some data from Knebel et al. (1977), shows the
ratio range of illite in the suspended physil suite at the individual sampling stations.
Samples were collected in the winter, spring, summer, and fall. Also shown are illite
ratio values for the bottom muds. The ratio values at a given site can vary by as
much as 0.4 (about 40% of average value). These numbers give some idea of the
problems invloved in working with suspended samples. Nevertheless, there appear
to be significant physil differences within the estuary system.
There is a significant change in the suspended physil suite in the vicinity of
station 12. This is near the location of the turbidity maximum during the period of
maximum discharge (winter). The winter suspended suite upstream of station 12
should be primarily river transported material. During the other three seasons, the
saltwater wedge moves farther upstream and presumably resuspends and transports
some of the winter physils deposited at station 12 farther upstream. The lowest
suspended illite values occur when there are turbidity maxima at stations 3, 6, 9 and
12. Apparently at these times strong vertical currents resuspended the entire physil
suite.
244

Downstream from station 12, the suspended illite ratio values increase and have a
wider seasonal variation. Much of this material, particularly in the southern reach, is
material resuspended by waves and tidal current. The suspended illite ratio values
are roughly twice those of the bottom muds. The distribution strongly suggests that
differential resuspension of illite occurs in the southern reach. This winnowing into
suspension of illite may be because the illite does not develop flocs as readily as
other physils, flocs with illite are smaller or less dense, turbulence breaks flocs and
illite is preferentially released. The distribution in Fig. 4-39 suggests the suspended
illite may be transported northward and deposited in the vicinity of stations 21, 17
and 14. Some is probably transported out of the estuary into the ocean. The above
interpretation differs somewhat from that of Knebel et al. (1977).
Studies of the physils in fiords and estuaries of Alaska (Kunze et a]., 1968;
O'Brien and Burrell, 1970), where the physils are derived from melting glaciers,
indicate there is no significant difference between the fresh water and salt water
physil suites. The published x-ray patterns indicate the physils are composed of
approximately 40% chlorite and 60% biotite. These physils are derived from igneous
and metamorphic rocks commutated by glacial abrasion. The physils in Glacier Bay
are essentially unweathered, though the width of the biotite peak suggests some
vermiculite layers are present. Farther north, in the Takn Estuary, minor amounts of
vermiculite, montmorillonite and mixed-layer biotite-vermiculite are present. These
physils were presumably derived from biotite and indicate some mild chemical
weathering occurs even in a glacial environment.

MARSHES AND TIDAL FLATS

Deposition

This summary of the physical characteristics of tidal deposits is primarily from


reviews by Frey and Basan (1978) and Davis (1983).
Along coasts that do not contain major deltas much of the physil material occurs
in marsh and tidal flat muds. However, because of the abundant supply of physils,
marsh sediments are abundant adjacent to deltas. Fig. 4-40 show the distribution of
environments in a typical intertidal zone. Marsh sediments are largely silty clays and
clayey silts. Salt marshes occur along the gently dipping coasts of most continents
and occupy the zone between fresh-water environments and marine to brackish
bays, lagoons, and estuaries. Where low energy conditions exist, as in the eastern
panhandle of Florida, marshes can border the open ocean. In areas of active coastal
uplift, such as the Pacific coast of North America, marshes are narrow and cover
relatively small areas adjacent to protected bays.
Along the Atlantic-Gulf Coastal Plain marshes are nearly continuous from Texas
to New York (except for southern Florida). Aside from the Mississippi Delta region
most of the large marshes are located behind barrier islands and are actively filling
coastal lagoons. The marsh usually occurs in the upper portion of the intertidal zone
245

SEDIMENT TRANSPORTING TIDAL CURRENTS


Decreasing capacity - Decreasing amount of material in transport
4
Decreasing competency - Decreasing size of material in transport

A Salt marsh No reworking by waves or organisms


Fairly slow Sedimentation
Sediment bound by plants. Filter effect of plants
B Higher mud flats Very limited reworking by waves and organisms
Fairly rapid sedimentation
Sediment bound by algae
IClInner sand flats

D Arenicola sand flats


Limited reworking by waves,extensive reworking by Corophium
S low sedimentation
Extensive reworking by waves and Arenicola
I
Slow Sedimentation
E Lower mud flats Little reworking by waves o r organisms
Rapid sedimentation
F Lower sand flats Little o r no organic reworking, minor wave-action, strong
tidal currents
Slow sedimentation
Fig. 4-40. Surface sediment facies distribution over intertidal zone of The Wash, eastern England.
Reproduced by permission of the Geological Society from “Intertidal flat sediments and their environ-
ments of deposition in the Wash”, by G. Evens in Geol. Soc. London, V21, 1965.

and the unvegetated tidal flat in the low to middle intertidal zone but can extend to
the supratidal zone.
The salt marshes typically contain a higher percentage of fine sediment (clay-silt)
than other coastal environments, implying a low energy environment. Most marshes
contain a luxuriant growth of rushes and grasses which act as baffles to slow the
current and trap the finer grained suspended material. Macroinvertebrates ingest
large quantities of suspended organic and inorganic material. This material is
secreted as relatively large pellets. In Georgia salt marshes various invertebrates
secrete from 100 to 800 g/feces/year/m? (Kraeuter, 1976). Benthic microbes
probably play a role in sediment trapping and substrate stabilization. Flocculation,
particularly joint clay-organic flocculation, is a factor but under most conditions
flocculation probably occurs before the suspended sediment reaches the marsh.
In the Mississippi Sound the “goast shrimp” Culliunussu produce up to 1 million
fecal pellets per square meter per day in tidal pools and shoreface areas. Much of
the mud in the adjacent marshes is composed of pellets washed into the marshes at
flood tide and filtered out by the marsh grasses. In one area, mud deposits
containing more than 40% oyster fecal pellets are at least 12.5 m thick over an area
246

of about 15,000 m2 (Pryor, 1975). Some tidal flats in Virginia contain more than
40% fecal pellets and include pellets from 18 species (Harrison, 1971).
The organic content of marsh muds is high, commonly in the 20 to 30% range in
Georgia and as low as 4 to 6% in New England marshes where productivity is lower.
Most marsh muds have a relatively high content of H,S and “free Fe”. As a result
ancient marsh shales have a relatively high content of pyrite.
Salt marsh deposits generally have a tabular shape and are seldom more than a
few meters thick. Due to bioturbation deposits may show no stratification. Others
are well stratified. Bedding is usually due to alterations of organic-rich layers, in
some instances peat, and clay and/or silt-rich layers. Tidal channels contain coarser
material than the marsh. Shell fragments can be abundant.
Tidal flats are low-relief deposits that fringe the shore lines of most low-relief
coastal areas. The extent of the tidal flats is related to the slope and tidal range. In
the Yellow Sea area they may be up to 25 km wide. In the Paleozoic many tidal
deposits were apparently much wider. Tidal flats may be composed predominantly
of mud or sand or both. The classical tidal flat deposits consist of interbedded thin
layers of mud and sand. The sand is deposited during the high velocity period of
flood and ebb currents and the mud is deposited during and near slack water.
Studies of the Wadden Sea tidal deposits (Van Straaten and Kuenen, 1958) show the
grain size decreases from the inlets between barrier islands, towards the interior.
The high velocity tidal currents keep the physils and physil flocs in suspension and
resuspends any that might have settled out during slack tide. As the water becomes
shallower and current velocities decrease the physil-rich material settles out to form
tidal flat mud deposits on the inland side of the tidal basin. In this example the fine
suspended material is derived entirely from the sea. Thus, the grain size decreases
landward, in contrast to other coastal environments where the grain size decreases
seaward. The high turbulence in the tidal flat favors the development of large flocs
which settle rapidly (Einstein and Krone, 1961).

Physils

There is a considerable amount of information about the physil composition of


estuarine and shelf sediments but relatively few studies have been made of tidal
muds.
Weaver and Windom (1988) studied 25 marsh cores collected (by helicopter)
from the area between the mouth of the Pee Dee River in northern South Carolina
and northern Florida (Fig. 4-41). The major rivers transport a physil suite rich in
kaolinite and the smaller Coastal Plain carry a montmorillonite-rich suite (see
Rivers), In general the kaolinite content of the marsh physils ( < 2 pm fraction)
decreases from 70 to 80% in the northern part of the area to 30 to 40% in the
southern part of the area.
The marsh muds throughout the area have a higher kaolinite and lower illite (0 to
7%) content than the physil suite of the offshore sands. The southward decrease in
kaolinite parallels the decrease in the relative amount of kaolinite in the major river
241

Atlantic Ocean

Percent kaolinite
(251% illite

Florida A
Fig. 4-41. Composition of kaolinite in the rivers and marsh muds of the southeastern United States.
Average kaolinite and illite (in parentheses) values for offshore sands are also shown.

muds. A detailed study of the marsh and river muds in the area between the
Savannah and Ogeechee Rivers, Georgia, shows that the marsh muds have a
kaolinite content (60 to 70%) similar to that of the freshwater river muds and
considerably higher than the muds near the mouth of the estuaries ( - 30%).
The physil distribution pattern strongly suggests that most of the marsh physils
are, in one way or another, derived from the rivers. During the flood stage river
suspended material can be transported to the marshes via the salt marsh streams
running parallel to the coast. I t also appears likely that some of the river physils that
escape the estuaries at flood stage are transported by tidal currents into the marsh
before they have a chance to settle on the continental shelf. Thus, most organic and
inorganic pollutants adsorbed on river physils are likely to be trapped in the
marshes rather than transported out to sea. It should be noted that some marsh
sediments are derived from the estuaries, courtesy of the Army Corps of Engineers.
Based on a study of the physils and the direction of tidal currents Kelley (1983)
concluded that the marsh muds, primarily illite with minor chlorite and montmoril-
lonite, of southern New Jersey were derived from resuspended Delaware Bay
bottom sediments and Delaware River suspended sediments. The suspended material
moves out of Delaware Bay during ebb tide, moves northeast along the Cape May
248

Peninsula, and then at flood tide is moved through the tidal inlets of Cape May
Peninsula and is deposited in the marshes and lagoons.
The physils in the tidal flat muds of the Wadden Sea, Netherlands, are all carried
in from the open sea by tidal currents (Van Straaten and Kuenen, 1958). In this
instance there are no major rivers adjacent to the area where the tidal flats are
developing.
Thus, the source of physils in salt marsh and tidal flat muds can be either
adjacent rivers, estuaries, or from the continental shelf. Mixed sources are likely to
occur in many situations.

DELTAS

Formation

The following short summary on the nature of deltas and deltaic deposition was
taken largely from Coleman and Wright (1971), Reineck and Singh (1973), Wright
(1978), and Davis (1983). Most major deltas have a relatively high percentage of
physils, particularly in the more seaward portions. Large deltas can develop enor-
mous thicknesses ( - 10,000 m) and contain large volumes of physils. Deltas are
depocenters of sediments that develop where a river discharges into a large body of
water. The primary factor determining whether a delta will develop or whether the
sediment will be swept into the open ocean is the balance between the sediment load
of the river, the slope of the continental shelf, and the effective wave energy on the
shelf. The role of the river and the amount and type of physils it will deliver to the
delta (or non-delta) is determined by the drainage area, the climate, the relief, and
the water discharge regime. Large deltas develop primarily on tectonically quiescent
coasts.
A delta contains a wide and complex variety of depositional environments.
Further, deltas have a wide range of shapes and environmental patterns. Only the
very general features will be discussed. Fig. 4-42 shows the three basic physiographic
zones present in most deltas. The subaqueous portion of the delta lies below
low-tide water level. Sediments generally become finer with increasing distance from
the river mouth. The seaward-most portion (prodelta) is commonly composed
primarily of physil-rich muds deposited from the river suspended load. The prodelta
muds have a high clay content and generally are well laminated, though the
laminations are difficult to detect without x-ray radiographs. Marine shells are
common. The prodelta environment contains the bulk of the volume of deltaic
sediments.
The subaerial portion (topset) of a delta normally consists of a thin veneer of
sediments that can be divided into two zones. The lower delta plain is the area of
riverine-marine interaction and lies between the low and the high tide water levels.
In humid climates physil-rich marsh deposits are a major feature of the lower delta
plain. The sediments are typically structureless and consist of organic clays inter-
layered with silts. Bioturbation may be extreme. Evaporites and barren flats may
249

I I

Fig. 4-42. Components of a delta. From Coleman and Wright, 1971; and Davis, 1978. Copyright 1978
Springer-Verlag.

develop in and climates. The upper delta plain is the older portion of the subaerial
delta and is not significantly affected by marine water. It is dominated by riverine
depositional processes. Muds are mostly present in swamp and lacustrine deposits.
Some of the swamps are well drained and periodically exposed to oxidizing
conditions. Fe oxide and calcium carbonate nodules may be formed during these
periods (Coleman, 1966).
In regard to the deposition of physils, deltas can be divided into two general
types. In the fluvial-tidal low-energy delta, much of the river suspended material is
deposited as mud in the subaerial and prodelta parts of the delta or in adjacent
environments (Mississippi Delta). In contrast, the river suspended material in wave
dominated lugh-energy deltas, much of the suspended material can be transported
long distances along the coast (Amazon Delta) or transported over the shelf margin
and deposited in the deep ocean (Columbia River).
During flood stage, rivers breach or overflow their natural levees and deposit
sediments on the subaerial deltaic plain; however, most sediments are transported to
the mouth of the river where they respond to a rapid diminuation of water velocity.
River-mouth process can be complex and will only be discussed briefly.
"... river-mouth effluent diffusion and sediment dispersion patterns depend on

the relative role of three primary forces: (1) the inertia of issuing river water and
associated turbulent diffusion; (2) friction between the effluent and the bed im-
mediately seaward of the mouth; and (3) buoyancy resulting from density contrasts
between issuing and ambient fluids" (Wright, 1978). The first two processes com-
monly cause the rapid deposition of the coarser material and the development of a
distributary mouth bar. Most of the suspended physils in the lighter river water
apparently remain in suspension as the lighter river water passes over the distribu-
250

Fig. 4-43. Deposition and sorting of sediments at a river mouth. From Scruton, 1960; Davis, 1978.
Copyright 1978 Springer-Verlag.

tary mouth bar and flows seaward over the denser ocean water (buoyant effluents).
A general decrease in velocity seaward coupled with turbulent mixing between the
two waters should promote flocculation similar to that which occurs in estuaries
(Fig. 4-43).
As in estuaries, the geographic location of the fresh water-salt water interface
depends largely on the river discharge rate. During the low water stage, when the
physil content is low, the flood tides can move into the river channels and
presumably cause physil flocculation in the channel. Some of the flocculated
material is likely transported seaward by the ebb tides or by the river waters during
flood stage. The intrusion of the salt water wedge into the channel causes a partial
impounding of river water which exits through crevasses into interdistributary bays
(splays) (Wright, 1978). If there is sufficient vertical mixing at the fresh water-salt
water interface, some of the flocculated material could be transported into the
interdistributary bays.
In areas where the tidal range is large (Amazon River) mixing by tidal activity
destroys vertical density stratification. The buoyancy effect at the river mouth is
negligible and flow in and out of the river is subject to tidal reversals. There is
extensive upstream transport of bed load by flood-tide currents which leads to
extensive sand accumulation in the channel. As currents are slack at high flood-tide
when the water levels are highest, silts and clays from suspension are deposited
along the high water levels of the river banks (tidal flats).
During river flood stage, the salt wedge is pushed from the channel, intense
turbulent diffusion decreases the buoyancy effect and increase the rate of decelera-
251

tion. Suspended sediment is deposited rapidly. However, not too rapidly, as turbid
water plums have been observed extending as much as 100 km offshore of the
mouth of the Mississippi River.
Ocean waves tend to remold the configuration of the shoreline. Their effective-
ness depends on the rate of supply of river sediments and the wave energy which
reaches the shoreline. One of the major roles of waves is probably to resuspend
some of the clay material and allow it to be transported seaward or along the coast
where it can be trapped in marshes and bays.
It should be noted that as well as prograding seaward rivers may abandon a
deltaic lobe, for various reasons, and start a new one. This produces lateral shifts in
the loci of deposition and can spread prodelta muds over wide areas of a coast.
The mechanisms of deposition at the freshwater-saltwater interface are similar to
those in estuarine environments. Syvistki et al. (1985) were able to observe and
measure the suspended material in the pro-delta region of a small delta in a British
Columbia fjord. In the upper pro-delta, as the river plume flows over the marine
water, much of the riverine suspended load, sand, coarse silt and < 10 pm flocs, is
deposited. The settling velocity of large particles decreased with increasing grain
size. In the lower pro-delta ( - 8 km from the river mouth) where the freshwater
thins seaward, the vertical flux of particles is controlled by biogeochemical interac-
tions such as pelletization of fine particles and flocculation (which occurs within
rather than below the surface layer in contrast to the upper pro-delta). The pellets,
Fe-rich, are produced by indiscriminate filter feeding plankton. At a depth of 40 m,
floccules are > 10 pm in diameter and reasonably compact. The interaction of
bacteria with the pellets and floccules increases with depth and seaward distance. At
depth, mucoid filaments form stable interconnecting webs.

Mississippi Delta

The Mississippi River drainage basin covers approximately 40% of the continen-
tal United States. It empties a suspended load of 213 million tons/year into the
relatively quiet Gulf of Mexico where it forms an elongate, fluvial-dominated delta
composed predominantly of silt and clay. X-ray analysis of the < lOpm fraction
indicates much of the river suspended load is transported directly through the river
mouth and deposited in the prodelta environment with little change in mineral
composition (Table 4-3). Analyses of bulk samples (Fig. 4-44) indicates the nature
of the fractionation that occurs as the suspended load is deposited. Quartz and
feldspar are relatively abundant ( - 70%) in the river muds. These minerals decrease
in abundance seaward and the physil content increase from 30% to 60% to 70%.
Shell fragments (calcite) are present in the distill prodelta muds. Relatively little
quartz and feldspar is transported into the prodelta environment.
The < 2 pm fraction contains 80 to 90% I/S ( - 1:4), 10 to 15% illite, and - 5%
kaolinite plus chlorite. This physil suite has remained relatively constant for at least
the past 15 m.y. X-ray analyses indicate there is no significant change in the physil
252

Carbonates

River Lagoon (bay) Delta (shelf) Near shore Deep


1 O % L (slope)
Number indicates samples analyzed
Fig. 4-44. Plots showing the average mineral composition of a series of recent mud samples from the
vicinity of the Mississippi River delta. From Shaw and Waver. 1965. Copyright 1965 SOC.Econ. Paleo.
Miner.

suite, except for diagenetic conversion of the montmorillonite to I/S, to a depth of


5,500 m (maximum depth sampled) (Weaver and Beck, 1971).
The river physil suite is not materially affected by seawater, except for the
exchangeable cations, and there is no apparent physical sorting. The physil suites in
sands, silty muds, and muds, and in various environments, channel. marsh, delta
fring, and prodelta, are essentially identical (Griffin and Parrott, 1964). However,
Griffin and Parrott did find that the abandoned St. Bernard sub-delta sediments
had a larger montmorillonite/kaolinite peak height ratio than the modern Missis-
sippi River sediments. The difference in peak ratio could be accounted for by a
- 5% change in the kaolinite content.
John and Grim (1958) studied a series of bottom samples extending from a
distributary river mouth 80 km east to the open shelf. They reported a decrease in
montmorillonite from an average of 65% to an average of 50% (their technique
underestimated the montmorillonite content). Mite and kaolinite plus chlorite
increased slightly, 9% and 1%, respectively. Actually, the physil suites of the samples
from the six deltaic subenvironments are essentially identical. The minor contrast is
between the river physil suite and that of the deltaic suites. Milne and Early (1958)
also found the muds on the shelf edge east of the delta contained more illite than
those at the mouth of the Mississippi River. The increase in illite may reflect the
influence of the illite-rich suite suspended suite originating in the Amazon River.
Samples immediately to the west of the modern delta d o not appear to have a high
illite content. X-ray patterns indicate the physil suite consists of approximately 85%
montmorillonite, 10% illite, and 5% kaolinite plus chlorite (McAllister. 1964).
Though most of the physils transported by the Mississippi River are deposited in
the deltaic complex, an appreciable amount escapes and is carried well o u t into the
Gulf of Mexico. During flood stage the suspended material is carried far beyond the
,.
0.
JS /'
Y o '
\ O.
1 0 .
0. r
CSZ
254

PERCENT SMECTITE PERCENT KAOLINITE

BENTHIC SAMPLES BENTH4C SAMPLES


ISUMMER 1976 AND FALL 19771 ISUMMER 1976 AND FALL 197r1

IN KAOLINITE
BENTHIC SAMPLES
(SUMMER 1976 AND FALL 19771

L I
Fig. 4-46. A,B) Kaolinite and smectite contents of bottom sediments for summer 1976 and fall 1977
sampling periods (N.W. Florida). Contour interval 108. Percentages are of the total clay mineral
fraction. C) Differences in percentages of kaolinite in the clay mineral fraction between summer 1976 and
fall 1977 sampling periods. From Doyle and Sparks, 1980. Copyright 1980 Soc. Econ. Paleo. Miner.

montmorillonite and kaolinite and the difference in the kaolinite percentages


between two sampling periods. Montmorillonite is dominant at the outer shelf edge
and decreases landward as the amount of kaolinite increases. As the rivers of
peninsular Florida supply relatively few physils to the shelf area, the distribution
reflects the interplay between Mississippi River physils (montmorillonite) and
Apalachicola River physils (kaolinite). The peak width at half-height of the 7 A
peak increases from north to south. Doyle and Sparks (1980) suggest this is due to
degradation in the marine environment. The increase in the 7 peak width could A
also be due to an increase in the chlorite content to the south. Griffin (1962) noted
there was a general increase in the amount of chlorite to the southeast. The chlorite
was apparently carried into the Gulf from the Atlantic Ocean. Surprisingly, as
shown in Fig. 4-46c, the physil suite in bottom sediments can change significantly
over a period of about one year. Further, the suspended suites at a given location
can vary significantly from hour to hour (Fig. 4-47). The variations in the suspended
suite correlate with minor changes in salinity and presumably reflect periodic
movements of the eastern Gulf waters (kaolinite) westward into areas normally
dominated by the Mississippi physil suite or vice versa. Temporal variation in the
composition of physil suites is to be expected in estuaries. Apparently short-term
temporal variations can also occur on the continental shelf if there are two or more
c

a
E
y 40
b
100
00
60
Suspended clay mineralogy (bottom)
station 2639 DM -3

b
y 40
d
80
100

60
~l Suspended clay mineralogy (top)
station 2639 DM -3
............
:.:. :

Kaolinite
.. .. .. .. .
... ... ,...
.
. II!. .
. .. ... ...
..
255

20 20
0 0
1030 1030 1030 1030 1030 1030 1030 1030 1030 1030
2/20 I 2/21 1 2/22 I 2/23 I 2/24 12/25 2/20 I 2/21 I 2/22 I 2/23 I 2/24 12/25
Time Time
Fig. 4-47.A,B) Variations over a five-day interval in the suspended clay mineralogy in near bottom and
near surface waters of Mobile Bay (Alabama). From Doyle and Sparks, 1980. Copyright 1980 SOC.Econ.
Paleo. Miner.

physil suites supplied to the shelf and if the shelf currents do not consistently flow
in the same direction at the same velocity. In addition to the physils shown in Fig.
4-47, talc is present in many of the samples ranging from trace to dominant. Is it
crop dust?
Fig. 4-48 shows the concentration of suspended material and general physil
composition of the suspensates in the surface waters of the northern Gulf of
Mexico. The nearshore suspensates are predominantly mineral detritus and < 20%
combustible organic matter. The suspensates in waters seaward of the shelf break
typically have a small physil content and are composed of > 70% combustible
organic matter (Manheim et al., 1972). The suspended physil suites are vaguely
similar to those of the underlying bottom sediments - kaolinite in the east with
montmorillonite increasing to the west. Illite is most abundant in the vicinity of the
Mississippi Delta. The x-ray patterns are of very poor quality. Manheim et al. (1972)
suggest zooplankton, that sweep in physils and organic matter and aggregate them
into fecal pellets, may account for the apparent rapid settling of physils (indicated
by lack of complete mixing) in the marine environment.

Niger Delta

The Niger Delta on the tropical west-central coast of Africa is considerably


smaller than the Mississippi Delta, and the river has about half the discharge. The
Niger is about equally balanced between wave and tidal domination, but the river
exerts a significant influence. As a result it has a smooth outline and no interdistrib-
utary bays.
The Niger River transports most of the detritus directly to the sea (Porrenga,
1966). The composition of the river muds ( < 2pm) are quite variable. The kaolinite
content ranges from 25 to 50% and montmorillonite from 20 to 45%. Illite is
relatively constant at about 25 to 30% (percentages recalculated from peak area
values). Despite the variability in the river physil suite, the kaolinite and montmoril-
lonite in the sub-marine portion of the delta forms a series of arcuate zones parallel
256

Montrnorillonite dominant
Montmorillonite-Kaolinite -0.25- Concentration contour (rng/l)
Kaolinite dominant Inferred concentration contour
Carbonate
Fig. 4-48. Distribution of total suspended matter in mg/liter and dominant mineral suites. Depth
contours are in meters. From Manheim et al., 1972. Copyright 1972 Amer. Soc. Lim. Ocean.

to the delta shoreline (Fig. 4-49). The inner zone contains less than 20% montmoril-
lonite; kaolinite is the dominant physil. The montmorillonite/kaolinite ratio sys-
tematically increases seaward with the outer zone containing more than 40%
montmorillonite. Porrenga (1966) believes the general seaward increase in
montmorillonite is due to differential flocculation. He also found the suspended
physil suite 40 km off shore contained more montmorillonite than the nearshore
suspended suite.
In contrast to the Mississippi Delta where the physil suite of the subaerial
portion of the delta is relatively uniform, the subaerial surface sediments of the
Niger Delta contain a variety of physil suites (Olorunfemi et al., 1983). The surface
physils in the central portion of the delta are kaolinite and halloysite. This is the
area of most intense leaching by meteoric waters. The physils are presumably
formed during recent weathering. Beidellite, illite, I/S, and kaolinite are present in
the eastern delta. The beidellite formed from young volcanic rocks. Montmorillonite
and nontronite comprise the physil suite in the tidal influenced coastal portion of
the delta.
257

Fig. 4-49. Montmorillonite in < 2 p fraction of Recent sediments of the Niger Delta. From Porrenga,
1966. Reprinted with permission from Clays Clay Miner. 14th Conf., 221-234. Copyright 1966 Pergamon
Journals, Ltd.

It is apparent that the available source areas and the physil suites are more
complex than envisioned by Porrenga. The information is too incomplete to
conclude that the offshore physil zonation is due to differential flocculation. Fecal
pellets, comprising from 1%to roughly 100%of the sediments, are present in nearly
all the marine sediments (Porrenga, 1966). They range in color and degree of
induration from gray and soft to dark green and dark brown and hard. The
distribution of the three types have a well defined pattern.
The green pellets (“early-stage glauconite”) occur in areas where very few physils
are being deposited. The pellets are composed of I/S (or G/S) containing 70 to 80%
smectite layers. The x-ray patterns indicate the I/S is similar to that in the clay
matrix; however, kaolinite and illite are not present in the pellets. The Fe,O,
content of the pellets (19%) is approximately twice that of the matrix and was
presumably scavenged primarily from organic material and Fe-oxides supplied by
the tropical river.
The brown pellets occur in water shallower than 64 m and consists of a poorly
crystallized (broad peaks) chamosite and goethite. The Fe203 content (20%) is
similar to that in the green pellets (I/S). Some of it occurs as goethite. The
chamosite has 5% more MgO (8.3%) and 5% less A1203 (8%) than the green pellets.
It is apparent that the structure and chemistry of the detrital physil (I/S) has been
appreciably altered to produce the chamosite. The gray pellets have a less regular
distribution than the green and brown pellets. In water deeper than 180 m they
258

constitute almost the whole of the bottom sediment. Their mineralogy and Fe
content is similar to that of the matrix muds. Presumably various animals have
different effects on the physils they process.
The high kaolinite content of the Niger sediments is presumably due to relatively
intense weathering in a tropical environment. The kaolinite is a climate indicator. In
contrast, the physil suite of the Mackenzie River Delta (Arctic Ocean) is composed
of equal parts kaolinite and illite with minor chlorite and montmorillonite; the
kaolinite is a source rock indicator. The physils were derived from Cretaceous rocks
which contain an average of 50% kaolinite.
In contrast to deltas from the more temperate regions, the physils associated with
the tropical Niger Delta are undergoing appreciable change. Many of the subsea
physils are being transformed by biological activity, and the subaerial ones are being
transformed by weathering.

Amazon Delta

The Amazon River Delta is an excellent example of a delta influenced by strong


coastal currents which transport the detrital physils thousands of miles. The
Amazon has filled an embayment and formed a large deltaic plain, but the
northwest flowing Guiana Current has prevented the development of a seaward
protrusion. Though muddy brown water has been observed 200 to 500 km offshore,
much of the Amazon physils are transported northwest and are deposited close to
shore as mudflats (Fig. 4-50). Appreciable material is transported at least as far as
the Orinoco Delta, 1800 km (Van Andel, 1967).

200m.
-c d,

Mud

Shelly sand with variaMe


admixture of mud
Calcareous sand,algae and
fossil reefs

Fig. 4-50. General distribution of bottom sediment types on the continental shelf along the northeast
coast of South America. The mouth of the Amazon River is at Belem. From Eisma and Van Der Marel.
1971. Copyright 1971 Springer-Verlag.
259

%
40

Montrnorillonite

.-+----a
Koolinite

20 -

0 200 400 600 000 1000 1200 14(


Distance f r o m river ( k m )
Fig. 4-51. Clay mineral composition of bottom muds versus distance (to the northwest) from the mouth
of the Amazon River. From Gibbs, 1977. Copyright 1977 SOC. Eion. Paleo. Miner.

Fig. 4-51 shows a plot of the average physil composition ( < 2 pm) of the shelf
bottom sediments in an area extending from the mouth of the Amazon River, 1400
km to the north. The montmorillonite content increases and illite and kaolinite
decrease with increasing distance from the source. A similar, but more complex,
trend occurs perpendicular to the shore. The implication is that most of the mica is
deposited near the river mouth and little is available for deposition farther from the
river mouth. Another possibility is that, as suggested by studies of suspended
physils, after the initial deposition, micas preferentially remain in suspension. Eisma
and Van der Marel (1971) studied the physils in the mud banks between the
Amazon and Guyana coast and found a similar trend. They also found the swelling
physils along the Guyana coast contracted less when treated with KCI than those of
the Amazon, indicating they had probably taken up some K from the seawater.
They concluded the physils along the Guyana coast, 1400 km north of the Amazon,
had been derived from the Amazon River. Transport time, in suspension, would be
about one month. The mud banks move northwest at the rate of 1.5 km/year and
would reach the Guyana coast in about 1000 years.
To test whether the distribution on the shelf was due to differential flocculation,
Gibbs (1977) placed some < 2 pm suspended material in seawater (quiet and
stirred) and periodically x-rayed the flocculated material that settled to the bottom.
He found no difference in the physil suite with time and concluded differential
flocculation was not a factor in the natural system. The various species of suspended
physils in the Amazon River have a relatively distinct size range. Mica-illite is the
coarsest (100 to 0.4 pm), kaolinite is finer (10 to 0.4pm), and montmorillonite is the
finest (0.9 to 0.1 pm). Gibbs concluded that the distribution he observed on the
260

shelf was primarily due to physical sorting by size; flocculation occurs, but due to
turbulence and low sediment concentration may not be an important process on the
open shelf. However, illite-mica apparently has a high-float quotient.
The composition of the suspended physil suite in the shelf waters is considerably
different from that of the bottom sediments. The latter have an illite/kaolinite ratio
near 0.8 at the mouth of the Amazon; the value decreases to 0.6 1400 km to the
northwest. The suspended suite near the river mouth has a ratio value of 2.3 which
increases to 7.3 in the outer shelf (recalculated from 10 A/7 A values). The water
near the river mouth contains 64 mg/l of terrigenous matter; the concentrations
decrease to 0.06 mg/l to the east and northwest. One explanation for the difference
in the ratios is that the suspend samples included the total physil suite, whereas only
the < 2 pm fraction of the bottom sediments was analyzed. The river suspended

Underlying bottom sediment


Suspended mineral particles deep water
Q.M.1
M

M .Q 1 M0.C

RC12-3 1305fm RC12-4


2

Q.M.1
1 Mo.C

'424-7 15991m V24-19


M.Q M

V24-9 1928tm V24-21

3 5 4 358 472 9 4 100 3 5 4 358 472 9 4 100


3.33 426 5.0 72 )/ 14C Angstrom
units
3 33 426 150 72 14(
25 20 15 10 ) e g r e 2~ e : 25 20 15 10

Fig. 4-52. X-ray diffraction scans of mineral particles suspended in deep water and of underlying bottom
sediment sampled at the same location in the Gulf of Mexico. Specimens mounted in preferred
orientation. Abbreviations are: 1. illite; M, mica; C, chlorite; K, kaolinite; T. talc: Mo. montmorillonitr:
Q, quartz. From Jacobs and Ewing, 1969. Science, 163,805-809. Copyright 1969 AAAS.
261

suite mica is relatively more abundant in the coarser sizes. For the coarse mica to
preferentially remain in suspension would require that it has a relatively large
diameter/thickness (D/T) ratio. That this may be the explanation is suggested by
the nature of suspended physils in the Caribbean Sea and the Gulf of Mexico.
The Amazon River’s discharge is estimated to be 18% of the total drainage of all
the world’s rivers. This water flows to the northwest and causes a salinity reduction
far into the Caribbean Sea. Various Atlantic currents pass through the Caribbean
and onto the Gulf of Mexico. Jacobs and Ewing (1969) have shown that physils
from the Amazon River are transported into the Caribbean and from there on into
the Gulf of Mexico. The bulk physil suite of suspended and deep water samples
from both the Caribbean and the Gulf of Mexico consists largely of mica (80 to
95%) with very sharp peaks and varying amounts of chlorite and kaolinite.
Montmorillonite is present in only trace amounts. In the Caribbean Sea the bottom
sediments have a physil suite similar to the suspended physil suite, primarily mica,
though the former has slightly more chlorite, kaolinite, and I/S (Jacobs and Ewing,
1965). As the suites are so similar, it is difficult to determine how much of the
bottom material came from the Amazon and how much from northern South
America. The contrast is more striking in the Gulf of Mexico. Fig. 4-52 shows x-ray

Fig. 4-53. TEM picture showing illite-mica flakes approximately five unit cells thick and D/T ratio of
200 to 1OOO. Thicker flakes are kaolinite. Circles are latex spheres. Bar = 1 pm.
262

patterns of mica-rich deep water suspended material and montmorillonite-rich


underlying bottom sediments from the Gulf of Mexico. The bottom suite is typical
of the Mississippi River sediments, containing only a moderate amount of 10 A
material. Sample V24-7 is only 300 km from the Mississippi Delta. The minor
amount of montmorillonite illustrates the limited input from the Mississippi River.
Jacobs and Ewing believe little mica actually settles in the Gulf and most is
probably carried into the Atlantic by the Florida current. Conceivably some of this
mica could have been deposited in the Atlantic and could account for the increase
in 10 A material in the marine portion of the Mississippi Delta. The contrast
between the suspended suite and bottom sediment physil suite, as has been shown
elsewhere, demonstrates one type of difficulty inherent in interpreting the signifi-
cance of physil suites.
A
The 10 x-ray peak of the suspended material from the Amazon River is
relatively sharp but has a shoulder on the lower angle side. This shoulder is not
present in the Caribbean and Gulf samples. The shoulder is presumably due to the
presence of illite and suggests the illite is dropped not far from the Amazon River,
whereas the mica preferentially stays in suspension.
Igneous and metamorphic mica flakes are much larger than illite; however, this
mica can be cleaved into very thin ( < 50 A), highly flexible sheets (Fig. 4-53).
Diameter/thickness (D/T) ratios can be 1000 or higher, compared to values
commonly on the order of 10 to 50 for kaolinite and illite. These sheets must be an
extremely buoyant and like a stingray can be kept afloat by very minor currents.
Talc, frequently abundant, is present in many of the suspended samples from
south of the Amazon to the center of the Gulf of Mexico. Talc is relatively common
in marine waters. Much of it is believed to have come from aerially-sprayed
pesticides. If this is true, then some of the mica probably has a similar source. For a
further discussion of the origin of talc see pages 323-324.

Pa Delta

The Po River, in northeastern Italy, flows eastward into the Adriatic Sea. Both
the suspended and saltation sediment load of the river is transported directly to the
sea. where the coarser particles settle on the delta front platform and prodelta slope
(Nelson, 1971). The suspended physil suite of the Po River consists of roughly equal
parts of mica (sharp peak) and montmorillonite with lesser amounts of chlorite and
serpentine. (For some reason or other Quakernaat (1968) did not find any
montmorillonite in the bottom muds of the Po River.) The suspended montmoril-
lonite, quartz, calcite, and feldspar decreases when the delta bar is crossed and
montmorillonite is virtually absent 5 km seaward and to the south of the bar at the
limit of the surface momentum current. The physil suite contains approximately
70% mica and chlorite is more abundant than serpentine. The bottom sediments
throughout the Adriatic have a relatively high montmorillonite content, with the
relative amount increasing southward. This increase is presumably. in part, due to
the influx of relatively montmorillonite-rich (30 to 60%) suspended sediments from
263

Fig. 4-54. Distribution of montmorillonite in surface samples of pelagic sediments in the eastern
Mediterranean. Most of the rnontmorillonite in the easternmost area is from the Nile River. After
Venkatarathnam and Ryan, 1971.

the numerous small rivers draining east into the Adriatic (Quakernaat, 1968). The
Adriatic is relatively shallow and sediment concentration near the bottom is high. It
is believed that tidal currents resuspend the bottom fines and the counterclockwise
marine currents preferentially transport the finer montmorillonite to the south. In
any event the contrast between the suspended and bottom physil suites is similar to
that observed off northeast South America and the Gulf of Mexico. As in these
latter areas the 10 A x-ray peak of the suspended material in the Adriatic is sharper
(narrower) than that of the bottom material, suggesting the illitic type material sinks
and the floating material is very thin muscovite.

Nile Delta

Eastward flowing marine currents have smoothed the outline of the Nile Delta
into the classic delta shape. The Nile River muds and delta alluvium are composed
primarily of montmorillonite (60 to 80%) and minor kaolinite and illite (Rateev er
at., 1966). There are little detailed data on the physils in the various deltaic
environments, but several studies show how the Nile physils are distributed within
the Mediterranean.
In the eastern Mediterranean the marine currents flow counterclockwise. Fig.
4-54 shows the distribution of montmorillonite and the bottom sediments in the
Mediterranean. It is apparent that most of the Nile clays are swept east and then
north by the marine currents (Venkatarathnam and Ryan, 1971).
264

Western United States

The Columbia River is the dominant sediment source (99%) on the northwest
coast of the United States. No delta has developed and much of the sediment is
transported across the narrow continental shelf to the continental slope and
Cascadia Basin (Baker, 1973). The suspended physil suite (total sample) from the
Columbia River contains an average of 33% montmorillonite and 17% chlorite; the
Mo/Ch ratio ranges from 1.1 to 3.5. Illite makes up the remaining portion of the
physil suite. The near bottom suspended material from the shelf has a similar
composition, Mo/Ch ratios > 1. Similarly, samples from the upper part of sub-
marine canyons have a ratio of 0.65 to 0.96; suspended samples from the distal
portion of the canyons and on the slope between canyons have lower ratio values,
ranging from 0.29 to 0.48. Thus, there is an apparent loss, presumably settling out,
of montmorillonite seaward. This is further suggested by a comparison of the
Mo/Ch ratio suspended physils in the surface waters (0.43 to 0.81) and bottom
water (1.3 to 2.7). This suggest that much of the shelf bottom waters are funneled
down the submarine canyons, supplying a relatively montmorillonite-rich physil
suite over much of the length of the canyon. The near surface suspended physils,
with a low Mo/Ch ratio, settle on the shelf in areas outside the canyons. This
pattern also suggests that the montmorillonite has formed flocs and the chlorite
occurs as very thin, buoyant plates.
Baker also analyzed some bottom muds from the canyons and slope and noted
that the < 2 pm fraction of the muds were similar to those in the total sample of
overlying bottom water. This illustrates the type of problem that commonly occurs
when physils from different size fractions are used for interpretation. In a study of
the bottom sediments ( < 2 pm) farther offshore, Duncan et a/. (1970) reported
values that indicate the muds from the Astoria Canyon have average Mo/Ch ratio
of 2.1 (52% montmorillonite, 25% chlorite, and 23% illite). The canyon empties into
the Cascadia Abyssal Plain. The Mo/Ch ratio decreases systematically to a low of
0.8 approximately, 450 km south of the canyon mouth. They suggest the increase in
chlorite in a southern direction is due to dilution of the Columbia River physil suite
with a chlorite-rich (51%) suite from the Rogue River in southern Oregon (marine
currents flow predominantly north). Of additional interest are the muds in the
narrow Cascadia Channel which is oriented north-south in the Abyssal Plain. These
muds have a relatively uniform composition (no southward increase in chlorite)
similar to the composition of the muds in the Astoria Channel for several hundred
kilometers. The Cascadia Channel muds are chiefly the upper portion of turbidity
layers and were transported as turbidity currents originating in the Astoria Canyon.
X-ray analyses of core samples showed that there was a relatively abrupt increase
in illite and decrease in montmorillonite (average 40%) at the Holocene - late
Pleistocene boundary. The increase in illite is apparently related to the development
of glacial conditions in the Rocky Mountains (low-grade metamorphosed sediments)
during the late Pleistocene. The portion of the Columbia River draining the
northern Rocky Mountains presently carries a physil suite with 70% illite. I t
apparently carried a relatively larger volume of sediment during the Pleistocene than
265

the lower lying southern tributary, Snake River, which carries a montmorillonite-rich
physil suite. The situation is now reversed.
These two examples from Oregon nicely illustrate a few of the complexities
involved in interpreting the orign of marine muds.
As might be expected, bottom flowing turbidity currents can move large volumes
of physils without causing any appreciable mineral sorting. In areas of slower
deposition, the shallow water currents can deliver suspended physil suites that have
been modified by differential settling and/or the mixing of two or more sources.
Over short periods of time, changing climatic conditions can affect changes in the
source areas and cause a temporal change in the physil suites.
To the south, off the coast of California, a study of a mud core from the Santa
Barbara Basin demonstrated that the effects of river floods can be detected in
marine sediments (Fleischer, 1972). Core samples consist largely of laminated olive
gray mud but also contains numerous gray mud layers, commonly less than 2 cm
thick. The gray layers had previously been considered to be turbidites. The physil
suites ( < 62 pm) of both muds are similar containing roughly equal amounts of
illite and montmorillonite, a moderate amount of kaolinite and minor amounts of
chlorite and vermiculite. The gray layers have less chlorite and vermiculite and more
illite than the olive layers and in that respect, are similar in composition to the
suspended physil suite from the Santa Clara River. On the basis of the physils and
other data, Fleischer concluded that the thin gray layers were the result of periodic
large floods in the Santa Clara River. The estimated average flood frequency is 120
years. Thus, it appears that where the continental shelf is narrow, the results of
major floods can be preserved in offshore basins where marine currents are minimal.

(06/08/19) 1 LAKES

Lakes can be classified on the basis of both their physical and their chemical
properties. The simplest approach is to classify them as clastic or chemical lakes.
Lakes range from small ephemeral ponds to large lakes in structural basins that
cover thousands of square kilometers and exist for millions of years. Hutchinson
(1957) has listed 76 processes which form lakes. Only a few of these form major
lakes: tectonic basins, volcanic activity, glacial activity, landslides, solution activity,
fluvial activity, and wind activity. Though lacustrine deposits are typically thin,
structural basin lakes can contain thick deposits. The Dead Sea, a rift valley lake,
contains more than 4000 m of sediments (mostly precipitates). Many of the thick
Triassic deposits in the eastern United States were deposited in rift valley lakes.

(06/08/19) 1 Clastic Lakes

In the typical clastic lake the nearshore deposits are the coarsest grained and may
be present as deltaic and beach sand and conglomerate deposits. The physils,
commonly along with precipitated carbonates and organic material, are deposited
266

Fig. 4-55. Associations of clay minerals in Recent lakes and lacustrine rocks. There is not a single
association of clay minerals that would be considered typical of lakes. Rather, lacustrine deposits are
characterized by diverse clay minerals that reflect source materials and climate. From Picard and High.
1972. Copyright 1972 Soc. E o n . Paleo. Miner.

towards the center or the deeper portion of lakes. Sediment is normally derived
from rivers and shoreline runoff and erosion. Sedimentation in the deep water is
slow. Where the water is relatively fresh flocculation may not occur, and clay-size
physils may not be deposited unless the water mass is nearly motionless (ice cover)
or the physils are pelletized. Flocculation can occur where solute concentration is
higher than a few parts per million. Fine sediment in the center of large lakes may
contain thin and parallel laminations, which may or may not be disturbed by
burrowing, but in most deep lakes the clayey muds are structureless (Picard and
High, 1972; Sly, 1978).
The physils in most clastic lakes are detrital and reflect the composition of the
source rocks. Fig. 4-55 shows the composition of the physil suites in 40 lake (recent)
and lacustrine (ancient) deposits. The data include some chemical lakes but for
some reason excludes those in which authigenic physils are a major component. The
physil suites in Fig. 4-55 are typical detrital suites and indicate there is no exclusive
lacustrine physil suite. A few examples will illustrate the relation between source
and lake deposits.
The average physil suite of bottom sediments from northeastern Lake Michigan
is 50% illite, 30% mixed-layer physils (I/S and Ch/V), and 20% chlorite. The
average composition of the physil suites in the glacial till from the adjacent shore is
the same as that of the lake sediments. In both the source area and the lake the
physil suites have a relatively uniform composition; however, some sandy samples
267

from the shelf area have more mixed-layer material than the average and the
chlorite is more vermiculitic. The relative increase in degraded physils suggests there
could have been some post-depositional leaching in the porous, coarse samples
(Moore, 1961).
The situation is similar in Lake Ontario (Thomas et al., 1972). The glacially
derived physils have an average composition of 60% illite (54-88%), 18% chlorite
(1-30%), and 158 kaolinite (6-2756). The physil suite is relatively uniform
throughout the lake. There is a positive correlation between clay-sized material,
water depth, and organic content. Lake Erie (Lewis, 1966) and Lake Huron
(Thomas et a/., 1973) have similar physil suites.
Analyses of samples from cores covering the entire postglacial sediment section
of southern Lake Michigan (Gross et a/., 1972) and Lake Superior (Dell, 1971)
indicate a similarity between the lake physils and those in the till deposits of the
surrounding areas. In Lake Superior the average physil suite contains 60% illite and
mixed-layer material, kaolinite, chlorite, vermiculite, and other expandable physils
are generally present in amounts of less than 15%.
Kaolinite and illite are the dominant physils in the ice-covered Stanwell-Fletcher
Lake, 640 km north of the Arctic Circle in Canada (Coakley and Rust, 1968). The
lake is permanently ice-covered except for a short period of time when the margin
ice melts. The kaolinite and illite are derived by mechanical disintegration of
Tertiary-Cretaceous shales in the drainage basins of the rivers draining into the lake.
It is irionic that one of the few lakes in which kaolinite is a major physil occurs
above the Arctic Circle.
Lake Constance, in central Europe, has a physil gradation reflecting the presence
of two source areas. Mica, smectite, and chlorite (minor) are present in the bottom
muds. The smectite/mica ratio systematically increases from 0.3 in the western
portion of the lake to 4-6 in the eastern part. The sediments in the eastern area are
delivered by the Rhine River, which primarily drains alpine igneous and metamor-
phic rocks containing only illite and chlorite. The smaller rivers draining into the
lake derive their physils from volcanic rocks and Cenozoic sediments containing
primarily mica and smectite. Thus, with increasing distance from the Rhine delta
the alpine physil suite (remote source) is progressively diluted by physils from the
local source rocks (Muller and Quakernaat, 1969).
Lake Tahoe, California-Nevada, presents a more complex example of source-
sediment relations (Court et d., 1972). The lake lies high in the Sierra Nevada
Mountains. It is bounded by granite on the south and east and by andesitic volcanic
rocks on the north and northwest. Streams draining the granitic area are trans-
porting vermiculite and “chloritic intergrades” (Ch/S, Ch/V); those draining the
volcanic area are transporting montmorillonite. The surface of the lake bottom
contains two types of sediment, an organic ooze (diatoms and pollen) containing
chloritic intergrades, and non-organic sediment containing mica, vermiculite, and
montmorillonite. Court et al. concluded that the organic ooze is the sediment
presently being deposited and the non-organic sediment was deposited earlier
during times of glacial activity. The latter sediments are kept exposed by bottom
currents.
268

The glacially derived material was primarily from the west side of the lake, where
granitic (vermiculite) and volcanic (montmorillonite) source areas are approximately
equal. Abundant mica, primarily biotite, is nearly completely restricted to the glacial
sediments. Its presence is presumably due to rapid glacial erosion of the un-
weathered granitic rocks. The abundance of chlorite intergrades with minor mica in
the post-glacial sediments reflects the dominance of the granitic source rock (70%)).
The development of chloritic intergrades rather than vermiculite from the biotite
suggests a change in the conditions of chemical weathering. Presumably the
vermiculite is the precursor of the chloritic material.
Thus, the distribution of the physils in Lake Tahoe is influenced by source rock.
currents. time, and climate.
Two studies of lakes located in volcanic terrain showed montmorillonite is the
major physil. The only physil in Crater Lake. in the collapsed caldera of ;I
Pleistocene composite volcano in the Cascade Range of Oregon, is montmorillonite
(Nelson, 1967). The main sources of the montmorillonite are altered volcanic
bedrock, pumice deposits, and possibly ash from past eruptions of intracaldera
cones. The most extensive clay deposits are formed by fumarolic and solfataric
alteration. I t is not clear how much of the montmorillonite is hydrothermal and how
much formed at lake water temperatures. Montmorillonite appears to be thc
principal physil in Lake Nicaragua, Nicaragua, which is surrounded by volcanic
rocks. The x-ray patterns are of such poor quality the interpretation is difficult
(Swain. 1966).
Several studies suggest factors other than source can affect the distribution of
physils in fresh water lakes.
Lake Kinneret, Israel, is fed by the Jordan River, carrying primarily smectite and
kaolinite. Kaolinite is preferentially concentrated near the Jordan River delta and
along the western shore where the river currents are concentrated. Smectite is most
abundant in the center of the lake. The distribution is apparently due to the
difference in particle size which determines the settling rate. A minor amount of
palygorskite occurs along the eastern shore and was apparently derived from a local
source (Singer el ul., 1972).
The Catatumbo River, supplying sediments to Lake Maracaibo, Venezuela.
carries a physil suite comprised largely of illite (74%) and kaolinite (26%). As the
fresh water plume mixes with the more saline lake water (salinity > 2%). the
kaolinite is apparently preferentially flocculated, and the muds undcrlying the
plume contain up to 40% kaolinite (Hyne et al., 1979). Even in lake waters, where
current patterns, tidal effects, and other processes have less of a masking effect than
in estuaries, differential flocculation is not highly efficient.
Studies of glacier-fed Bow Lake, eastern Canadian Cordillera, indicate that
inflow and outflow velocities are such that most of the clay-size material should not
be deposited, yet is abundant in bottom sediments (Smith and Syvitski, 1982).
Analyses of material collected in sediment traps indicate that approximately 60% by
volume of the material collected occurs as ovoidal pellets 1/8 to 1/4 mm in length.
The fecal pellets are likely produced by pelagic copepods. Thus, in lakes as well as
in estuaries, pelletization can be a major factor in clay sedimentation.
269

Table 4-5
Variation of Average Physil Suite with Average Chloride Concentration for Lake Ponchartrain and Lake
Maurepas. (After Brooks and Ferrell. 1970.)
Physils Chloride Content (mg/l)
300 1200 3250
Montmorillonite 47% 54% 61 %
Mite 3% 9% 8%
Kaolinite 50% 37% 31%

LAGO SALINO
Lake Pontchartrain and Lake Maurepas in southeastern Louisiana are examples
of shallow, coastal, brackish water lakes (Brooks and Ferrell, 1970). Physils are
supplied to the lakes by various rivers and bayous (north and west) and by erosion
of the shoreline. Salt water is supplied through passes along the eastern and
southeastern shore. Chloride concentration ranges from 300 mg/l in the western
area to 4500 to 5500 mg/l near the passes. Current and salinity patterns are highly
variable due to frequent and appreciable changes in lake level. In spite of this the
physils show some segregation effect (Table 4-5).
The distribution suggests kaolinite is preferentially flocculated in the low salinity
environment. Montmorillonite and illite are more stable and tend to flocculate in
higher salinity environments. The segregation is not particularly striking.
A modern analog of ancient coal swamps occurs in the Lower Coastal Plain of
South Carolina (Staub and Cohen, 1978). The coastal Snuggedy Swamp contains a
mixture of salt marsh and fresh water vegetation. Peat deposits are as much as 250
cm thick. The physil suite of the swamp muds contains 45-60% montmorillonite,
30-45% kaolinite, and 10%illite. Immediately underlying the peat beds is a zone 30
to 40 cm thick which contains a physil suite composed of 60-80% kaolinite, 5-25%
montmorillonite, and 10% illite. Based on the mineralogy and structural features, it
is believed that the kaolinite formed from montmorillonite by acid waters (pH
4.0-5.0) filtering down from the overlying peat.
Analyses of the low temperature ashes (LTA) of plants and peats from three
swamps (Snuggedy Swamp, coastal South Carolina: Okefenokee Swamp, Georgia;
Mississippi Delta) led Renton et al. (1980) to suggest that physils and quartz formed
authigenically in the plants (wood, bark, leaves, grass) and peats. X-ray analyses of
the LTA indicated that both the plant material and the peat contained what is
apparently an amorphous alumino-silicate (x-ray peak around 4 A) and varying
amounts of quartz, kaolinite, I/S and illite. The silicate minerals are less common
and less abundant in the plant material than in the peat. Kaolinite is more abundant
in the low pH (4-5) than the high pH (6) peats. The authors concluded that the
silicate minerals crystallized both in the plants and in the peat from the amorphous
material present in the plant material.
The average LTA in the plant material was 8%; this compares to an average of
10% for West Virginia coals. Si, Na, Ca and Mg are more abundant and Al less
abundant in plant ash than in peaked coal ash. If the Al is held constant, a
considerable amount of other ions must be removed before the plant material is
270

converted to peat and coal. The average AI,O, content of the LTA of plants from
the 3 swamps ranges from approximately 1.5 to 5% and the SiOz from 7 to 20%. The
authors note that at pH values above approximately 5, the microbial activity would
significantly decrease the organic content, thus increasing the ash values to the level
that the residue would not be classed as a coal. Many of their peat samples contain
more than 80% ash. If this material is all derived from the inorganic content of plant
material, then the microbes must have been active.
The suggestion is interesting, but much more work remains to be done to
establish that the physils in the plants and peat are not detrital, either water or wind
transported.
There is little evidence for the authigenic formation of physils in fresh water
lakes. Most lakes contain diatoms which obtain silica from the dissolution of
volcanic glass and minerals. In spite of the availability of silica, the development of
crystalline alumino-silicates has not been adequately documented (Jones and Bowser,
1978). Muller and Forstner (1973) reported that in Lake Malawi, in southern East
Africa, nontronite has formed authigenically by the reaction of Si-enriched hydro-
thermal waters and ferrous Fe derived by the interaction of the hydrothermal waters
with the sediments. However, the nontronite is hydrothermal; the mechanism of
formation is similar to that in submarine hot springs.

Saline Lakes

Saline lakes are of considerable interest because they produce a wide variety of
exotic minerals, including a wide variety of physils. Lakes are considered to be
saline if they contain more than 5,000 ppm dissolved solids; maximum concentra-
tions are close to 400,000 ppm. High concentrations are obtained by evaporation
exceeding inflow or by the inflow being saline, or both. Saline lakes range from
small ephemeral ponds (playas) to deep perennial stratified brine bodies like the
Dead Sea. The most favorable conditions for their formation are found in rain
shadow basins (Eugster and Hardie, 1978).
Saline lake brines are dominated by a relatively few major solutes: SO,, Mg. Na.
K, HCO,, CO,, SO,, and C1. The relative proportions of major solutes can vary
drastically. The major anion compositions of saline lake brines are very diverse, but
most brines are dominated by a single cation, Na. Eugster and Hardie (1978) divide
brines into four major types: Na-C0,-CI-SO,; Na-CI-SO,; Na-Mg-CI-SO,; and
Ca-Mg-Na-CI. Most lake solutes are obtained by the action of acid rainwater on the
surrounding soils and rocks. These dilute inflow waters are concentrated by subse-
quent evaporation.
In most lake brines the Mg/Ca ratio is relatively low and calcite and/or gypsum
precipitate early. The resulting increase in the Mg/Ca ratio causes Mg-rich minerals
to crystallize: protodolomite, magnesite, and Mg silicates (commonly sepiolite, talc,
kerolite, Mg smectites). Where detrital Al-bearing physils are present they are often
converted to a chloritic physil.
271

The geochemistry of saline lakes has been studied in great detail; unfortunately,
only a few comprehensive studies have been made of the physils. The best known
authigenic silicates in saline lakes are the zeolites, usually an alteration product of
volcanic glass.
Droste (1963) analyzed physils from more than 60 saline deposits from the
western United States and concluded that evaporite sequences of nonmarine origin,
commonly Na and Ca environments, generally are not conducive to authgenic
formation of physils. The physil suites strongly reflect the source rocks surrounding
the basins. However, other studies indicate that under some conditions authigenic,
Mg-rich physils d o form in saline lakes. Droste analyzed the terrigenous clastic
interbeds rather than the salts; this may, in part, account for the absence of
authigenic physils. Terrigenous beds are deposited during the periodic influx of
fresh water, and the physils are not likely to be seriously affected. Authigenic
physils should form after advanced evaporation and be associated with the salt
beds.
The most famous American salina is Searles Lake (Na-Ca playa), California. The
saline beds contain halite, trona, hanksite, borax, and ten other evaporite minerals.
Drost (1961) believe the physils are detrital. Illite and montmorillonite are domi-
nant; minor amounts of kaolinite and/or chlorite are present. There is some lateral
and vertical variation in the composition of the physil suites, which is largely due to
variations in the composition of source rocks surrounding the lake. There is a
suggestion that some of the chlorite has been degraded, but this could have
happened in the source area.
Hay and Moiola (1963) report the presence of authigenic or diagenetic K-felds-
par, analcime, searlesite, and phillipsite in clay and volcanic ash, largely rhyolite,
layers in Searles Lake. The glass in the ash layers altered to the minerals listed
above and not to montmorillonite. In fact, the authors suggest detrital montmoril-
lonite was dissolved and provided the Si and A1 for the formation of K-feldspar,
searlesite, and zeolites (Hay and Moiola, 1963). Later, Hay and Guldman (1986)
found that ash layers deposited in saline nonalkaline environments ranged from
fresh to highly smectitic (with clinoptilolite and analcime). During periods when the
lake waters were highly alkaline, detrital montmorillonite and kaolinite altered to
form K-feldspar and analcime; some montmorillonite was transformed to illite-rich
I/S. The former reaction required about 50,000 years and latter, 85,000 years.
The physil suites in six other saline (Na-Ca) lakes near Searles Lake have similar
physil suites - montmorillonite and illite dominant and minor chlorite/kaolinite
(Droste, 1961). Both layers of unaltered volcanic glass and thoroughly altered glass
(montmorillonite) are present. Though most of the montmorillonite is detrital, some
apparently formed from volcanic glass. Presumably these lakes had a different water
chemistry than Searles Lake.
In a study of 19 other Recent desert lakes (Nevada, California, and Oregon) of
varying salinity (Na-Ca), Droste (1961) again found a close similarity between the
physil suite of the lacustrine sediments and the suite derived from the source rocks.
However, several lakes in areas where Recent volcanism was extensive contained an
abundance of amorphous material and montmorillonite. Droste believes that
272

montmorillonite is presently forming from glass in these lakes, particularly Albert


Lake and Silver Lake, Oregon.
Analyses of the clay fraction of sediments from the Great Salt Lake and its
principal inflow streams have similar physil suites as determined by x-ray analysis
(Jones and Spencer, 1985). The suite is composed of I/S (high S), illite, and
kaolinite. Below a sediment depth of 6 meters the clay fraction contains twice as
much Mg and half the Al of the clay fraction of the modern inflow stream
sediments. The authors suggest that Mg is adsorbed in the interlayer exchange sites
of the detrital dioctahedral smectite and then reacts with pore fluid silica (from
diatoms) to form trioctahedral stevensite interlayers (not positively identified). The
average octahedral occupancy increases from 2.1 for the detrital I/S to 2.6 for the
final product. As no Mg-carbonates are being formed in the lake, the Mg budget is
balanced by the formation of the Mg-smectite.
A similar transformation was observed to occur in saline Lake Abert, Oregon,
sediments (Jones and Weir, 1983). The solute budget indicates a rapid and thorough
depletion of solute Ca and Mg and a longer term loss of K. The Ca is removed as
calcite and the Mg and K are apparently incorporated in the silicates. The detrital
physils, derived from the weathering of adjacent pyroclastic rocks, are highly-charged
dioctahedral montmorillonite and interstratified montmorillonite/intergrade smec-
tite-chlorite. The physils ( < 0.1 pm) in the lake muds contain appreciably more Mg
(up to 20% vs. 4-8% MgO) and K (up to 4% vs. 0.1-1.4% K,O) than the detrital
physils. Table 4-6 shows typical structural formulas for the two types of physils.
It is thought that samples with a high K content (21jb) have developed illite
layers. The samples with a high Mg content (21cb) and a high octahedral population
have developed Mg-rich trioctahedral layers (stevensite). Because of the abundance
of detrital physils the presence of illite and stevensite was not confirmed by x-ray
analysis; however, infrared studies confirm the presence of trioctahedral layer
silicates. Spatially, the amount of Mg in the lake physils increases in the direction of
increasing salinity. The samples with high K values appear to be concentrated in

Table 4-6
Structural Formula for Na-Saturated Weathered and Lake Abert Physils. c 1 p m Fraction. (From Jones
and Weir, 1983.)

Weathered Rock Lake Abert Sediments


R24 R12d * 21jb 21ch
Tetrahedral Si 3.94 3.55 3.75 3.90
Al 0.06 0.45 0.25 0.10
Octahedral Al 1.32 0.68 0.65 0.39
Fe” 0.29 0.75 0.57 0.30
Mg 0.37 0.81 0.93 1.85
c 1.98 2.24 2.15 2.54
Interlayer Na 0.48 0.42 0.36 0.27
K 0.01 0.1 1 0.74 0.1 1
* Contains chloritic interlayers.
273

sediments deposited in fresher water (brackish). Though the lake sediments contain
an abundance of andesitic volcanic ash, there is no evidence (SEM) that it altered to
smectite in the lake environment (Deike and Jones, 1980).
The formation of smectite in Lake Chad in the center of Africa has been
reviewed by Gac et al. (1977). The lake is fed mainly by the river Chari. The lake
proper has the lowest solute concentration. Concentration by evaporation progres-
sively increases from the southern delta area to the northern interdunal depressions
where the salinities can be very high (200 g/1 of dissolved material). These waters
are chiefly enriched in Na, CO,, and HCO,, and are strongly depleted in Ca and
Mg-
As kaolinite is the dominant physil brought into the lake by rivers and smectite is
the dominant physil in the bottom sediments, it is apparent that much of the
smectite is authigenic. In the center of the lake, where solute concentration is lowest,
the smectite is a typical Al-montmorillonite with a relatively high Fe content
(Fe,3,12).The following reaction was postulated:
kaolinite + goethite + H,Si04 + Mg + montmorillonite + H++ H 2 0
In the delta area where solute concentration is, presumably, relatively low,
montronite pellets are present.

( si3.83 A10.06Fe,3,T1 ) (Fe:.~6Mg0.20)010 0.2zNa 0.02

Pedro et al. (1978) suggest that Fe transported by the Chari River is precipitated as
Fe3+ hydroxy-complex which forms oolites. A reducing environment in the oolite
centers allows Si, from solution, to migrate to the center and form a Fe2+-Si
hydroxy complex; later oxidation leads to the formation of a Fe3+-Si hydroxy
complex and eventually nontronite.
In the interdunal depressions the smectite, associated with amorphous Si and Na
carbonate salts, is similar to saponite:
si4 (A10.51Fe0.16Mgl .90l o l O (0H)2Ca0,02Na0.13K0.02

In an attempt to establish the authigenic formation of smectite, Gac et a/. (1977)


evaporated a large volume of Chari River water. They found that Si and Mg
coprecipitated with a final molar ratio of 1.3, approximately equal to that in talc.
No talc was observed, but there did appear to be an increase in smectite content of
the physil suite. Perhaps the smectite was stevensite, which has a Si/Mg ratio
similar to talc.
It is of interest to note that not only did a Mg-rich smectite form under high
salinity and high pH (9.3-9.6) conditions, but Fe-rich and Al-rich smectites formed
under the less saline and lower pH (6.5-7.5) conditions (Tardy et a/., 1974).
Though Mg-rich silicates are the most common neoformed physils in saline lakes,
Stoffers and Singer (1979) have described two African lakes in which montmoril-
lonite has apparently been transformed into illite. Both lakes, Lake Albert and Lake
Manyara, are located in the Great Rift region in eastern Africa. Ultra-basic volcanic
rocks of a potassic nature are the source of much of the dissolved and detrital
material supplied to Lake Albert. The lake waters are alkaline but have a relatively
214

Lake MObutO S c K s e k o I Lake Mobutu sese s e b


Glycolated clay samples Clay samples
Station A1
Fract10n<02 pm
Station A1
Core depth
210cm

6Wcm

Core depth

710 cm

I
15 10 5 2 15 10 5
Degrees 2 9 Degrees 2 0 Cu K LY
Fig. 4-56. X-ray diffractograms of selected glycolated clay samples (fraction < 2 pm and < 0.2 p m )
from Lake Albert showing the conversion of smectite to illite. From Stoffer and Singer, 1979. Copyright
1979 Geologische Rundschau.

low salinity. The K/Na (0.7) and Mg/Ca (3.3) ratios are remarkably high. Fig. 4-56
shows the nature of the physil suite to a depth of 1040 cm. The recent detrital physil
suite is composed largely of dioctahedral smectite with minor illite, kaolinite, and
traces of quartz. The smectite gives way with depth to an I/S phase and to a fairly
well developed illite at 750 cm. Kaolinite decreases and the K 2 0 content increases.
The illitic sample at 750 cm contains 7.23% K 2 0 , indicating it contains approxi-
mately 90% illite layers. A reverse sequence is observed below the illite zone. The
proportion of illite layers and K,O decreases, kaolinite increases, and at 1040 cm a
relatively good smectite is present. The illite-rich zone also contains the maximum
amount of Mg, much of which is present in protodolomitic oolites and shells.
It was postulated (Stoffers and Singer, 1979) that the physil sequence reflects the
evaporation history of Lake Albert. The detrital smectite in the bottom of the core
was deposited when the lake was open and dissolved ionic concentration low. About
25,000 years B.P. the lake level was lowered, it became a closed lake, and ionic
concentration increased; an I/S physil was formed. Between 16,000 and 12,500
years B.P. (dry ice age) the lake level was at its lowest and the K and Mg
concentrations were high enough to convert the smectite to illite. The MgO content
of the clay fraction increased from about 1.5 to 6.6%. Mg presumably replaced Fe3+
in the octahedral sheet and increased the layer charge. During the wet postglacial
period the water level rose and the process was reversed.
215

Lake Manyara is somewhat different. It is a closed basin with internal drainage.


Average water depth is only 3 m, and the water is a highly concentrated Na-bi-
carbonate-carbonate type. Water chemistry fluctuates drastically. The lake is sur-
rounded by basaltic lavas and ashy clays. The present lake waters have a very low
Na/K ratio (0.005). Several cycles of smectite-I/S-illite (the only physils present)
were detected in a 56 m core. In contrast to Lake Albert, volcanic material and
zeolites are abundant. Tuff beds, in addition to unaltered tuff, contain only zeolites
and amorphous material. In intervals where smectite is abundant the fine fraction
( < 0.08 pm) is amorphous; illite is present in both the fine and coarse fraction of
the illitic intervals. The interstitial water in the illitic intervals contains three times
the concentration of K as the smectite intervals.
The authors suggest that when the waters were dilute, smectite, accompanied by
diatoms, formed from volcanic glass. At higher levels of alkalinity and salinity,
various zeolites crystallized from the glass and presumably some illite layers
developed. They also suggest that the early formed zeolite erionite (K, Na) could
have been transformed to analcime (Na), releasing K to the interstitial solutions and
further forcing the conversion of smectite to illite.
It should be pointed out that it is possible that, in addition to the smectite, the
I/S and illite could have been formed in the volcanic soils adjacent to both lakes in
response to climatic fluctuations. This is a less likely scenario. In any event, it is
evident that illite can form from volcanic rocks at surface temperatures. It is
possible that appreciable illite formed under similar conditions during the Pre-
cambrian and early Paleozoic.
Sepiolite is another Mg silicate commonly formed in saline or alkaline lacustrine
environments. Experimental studies by Wollast et al. (1968) indicate that when
stream waters are evaporated and the p H becomes high, sepiolite will precipitate
from solution of the Si activity is lowered by the growth of diatoms, sepiolite could
form in the bottom muds where the diatoms are dissolved.
In addition to Mg, Si is obviously a necessary ingredient for the formation of Mg
silicates. As noted previously, Si may be obtained directly from solution or from the
dissolution of diatoms. Badaut and Risacher (1983) have shown that, in addition,
Mg silicates can crystallize directly on diatom frustules. In a study of ten Bolivian
saline lakes, located in a volcanic-rich area between the Eastern and Western
Cordillera of the Andes, they found that diatoms in the surface layers of five of the
lakes were coated with and partially replaced by a Mg silicate that is apparently
stevensite. In the other five lakes the diatoms were unaffected. Na, 3,590 to 103,300
mg/l, is the dominant cation; calcite and gypsum occur in the diatom mud. Mg
concentrations range from 34 to 3480 mg/l; Ca/Mg ratios range from 4 to 0.4.
There is no relation between Mg concentration and clay authigenesis; however,
there is a relation with pH. Lake waters where stevensite may be forming have p H
values ranging from 8.5 to 9.18. The other five lakes have pH values of 7.2 to 8.15.
Thus, the formation probably is related to pH or a(H+) rather than Mg concentra-
tion a(Mg2+).This confirms the experiments of Siffert (1962), which indicated there
is no direct relation between the crystallization of Mg smectites and salinity and Mg
concentration. Even when Mg concentration is low, the precipitation of Mg smec-
216

tites depends on the concentration of OH- ions in solution when dissolved Si


reaches the saturation point. In lacustrine deposits the Si concentration is increased
by evaporation. Thus, in these lake waters which are saturated in Si, Mg smectites
form only where the pH is high. For some reason or other it prefers to nucleate on
the diatom surface rather than precipitate directly from solution. The latter process
may have occurred and not been investigated.
What should theoretically crystallize often does not. Parry and Reever (1968)
studied six pluvial lake Pleistocene deposits in the southern High Plains, Texas, and
found definite sepiolite in only one deposit, Mound Lake. Guven and Kerr (1966)
found no sepiolite in five playa deposits in the Great Basin; Droste (1961) found
none in 45 playas in the Mojave Desert, California. Most of the physils, primarily
illite, I/S, and montmorillonite, are detrital and were not seriously altered in the
saline environments. In Mound Lake, sepiolite, associated with illite, smectite, and
kaolinite, occurs in dolomite beds. The Na/Mg ratio of present-day Mound Lake
waters ranges from 3.4 to 8.6 and the Mg/Ca ratio from 16 to 152. pH ranges from
7.3 to 8.4. The relatively high Mg content and the higher p H values apparently
favored the formation of sepiolite. Both values may have been higher when the
sepiolite formed.
In a follow-up study McLean ef al. (1972) found both sepiolite and palygorskite
were relatively abundant in Pleistocene playas in the vicinity of Mound Lake. The
two fibrous clays were associated with carbonate layers (sepiolite and dolomite:
palygorskite and calcite) interbedded with detrital material. The fibrous clays
apparently formed during times of periodic desiccation. The two associations
suggest the availability of Mg fluctuated.
Mg silicates are abundant in the Plio-Pleistocene deposits in the Amargosa
Desert, southern Nevada. The area is now partially covered with a playa lake.
Sepiolite beds up to 1.2 m thick occur interbedded with trioctahedral smectite.
Fine-grained dolomite is present in both types of clay. Papke (1972) identified the
smectite as saponite and suggested both physils formed authigenically in highly
saline lake waters (evaporation).
A more detailed study by Eberl et ul. (1982) and Khourty et al. (1982) showed
that the smectite was a mixed-layer kerolite-stevensite with the proportion of
stevensite layers ranging from 0 to 80%. The structural formula for a mixed-layer
sample containing 50% stevensite layers is:

They suggest that variation in the proportion of expanded layers may be due to
differences in the Mg/alkali ratio or water temperature. Both the sepiolite and the
K/S clay beds are characterized by a desiccation texture and intraclasts. The fibers
in the sepiolite beds are long and the sepiolite apparently grew from solution. The
scattered sepiolite that is present in the K/S clay has short fibers and presumably
formed by dissolution and recrystallization of the K/S. Both of the above features
were observed in the lagoonal Miocene palygorskite deposits of Georgia (Weaver
and Beck, 1977). In the latter case the smectite is a montmorillonite and Al-contain-
271

ing palygorskite, rather than sepiolite, formed. It is of interest to note that the small
amount of smectite in the purer palygorskite beds is stevensite.
The area in which the Mg physils formed contains 20 springs, whose waters drain
Paleozoic carbonate rocks. The waters are a Ca-Mg-Na-HCO, type with a pH of 7.4.
The Mg concentration averages 24 mg/l; the Ca/Mg ratio is 2.0 and the Na/Mg
ratio is 3.1; the SiO, concentration is 24.6 mg/l. Isotopic studies of the dolomite
and sepiolite suggest these minerals crystallized from relatively dilute waters rather
than brines. Calculations indicate that the spring waters require only about 10%
evaporation for dolomite and sepiolite to precipitate. Because K/S and sepiolite
have nearly the same Mg/Si ratio, the alkali/H+ ratio presumably determines
which physil will precipitate. Increased evaporation of lake water will increase the
alkali activity and pH; both would favor the formation of K/S. Experimental
studies by Siffert (1962) showed that sepiolite precipitates at pH 8.5 and that at
higher pH values Mg smectite precipitates. At a pH higher than 9 talc precipitates
along with the smectite.
A massive, - 12 m thick, deposit of sepiolite occurs in the semiarid Amboseli
Basin on the Kenya-Tanzania border. The Pleistocene clay bed is underlain by
dolomite which contains veins and pockets of sepiolite. The sepiolite bed contains
dolomite and calcite. The present Lake Amboseli contains water only during the
brief rainy seasons. Much of the basin is bounded by alkaline olivine basalts of
Kilimanjaro and fed by numerous springs and small streams. Mg concentration
ranges from 0.4 to 60 ppm; SO,, from 10 to 107 ppm; Na, 7 to 2,030 ppm; pH, 6.5
to 9.95; Ca values are similar to those of Mg. The waters reflect the chemistry of the
olivine basalt. Evaporation causes the waters to become saturated with respect to
sepiolite. Recent waters have converted some of the sepiolite to kerolite. This is
caused by a decrease in pH. Experiments show kerolite should form between pH 7
and 8 (Stoessel and Hay, 1978).
In southwest Africa, high pH waters in an artesian aquifer ascend a slope where
evaporation and decarbonization close to the surface cause Mg2+ and SiO,, dis-
solved during the ascent, to precipitate as sepiolite (Kantz and Porada, 1976).
The Neogene borate deposits of Turkey were deposited in shallow lakes fed by
hydrothermal solutions related to volcanic activity. The lake waters were rich in
alkali, alkali-earths, Ca-borates, and borate polyanions with a pH around 9. The
Authigenic physils in the borate beds are Mg-montmorillonite, hectorite, vermicu-
lite, Ch/S, and Ch/V (Ataman and Baysal, 1978).
This Page Intentionally Left Blank
279

Chapter V

MARINE TRANSPORT AND DEPOSITION

WATER TRANSPORT

Introduction

Other than those formed from submarine volcanic material, most of the physils
in the oceans are detrital, so, transport mechanisms will be briefly reviewed.
Material is transported to the oceans primarily by streams; however, atmospheric,
and to a lesser extent glacial, transport are significant.

Shelf

At the present time rivers with large sediment loads transport about 7 X lo9 tons
of suspended sediment to the oceans each year. By extrapolation, it is estimated that
the total riverine load entering the oceans is 13.5 X lo9 tons. However, it is not
known how much of this material actually reaches the open ocean. Much of it is
deposited in the lower alluvial plains, deltas and estuaries. For example, in the
Yellow River of China 33% of the sediment is deposited on an alluvial plain, 43% is
deposited in the delta region and only 24%reaches the ocean. The 13.5 x l o 9 tons
value takes this into account (Milliman and Meade, 1983).
Analyses of surface suspended samples from both sides of the Atlantic and the
western Pacific (Emery and Milliman, 1978) indicate that total suspended matter
concentrations exceeding 0.25 mg/l usually extend 50 to 200 km offshore. Maxi-
mum extent is on the order of 500 km. The non-combustible material (clays, silts
and siliceous skeletal debris) generally comprised more than 50% of the total
suspended matter, particularly near river mouths. They concluded that the vast
portion of fluvial sediment introduced to the oceans is concentrated near land,
primarily along the western sides of oceans, and only during major storms or floods
does much sediment escape the nearshore area. As discussed elsewhere (Chapter
IV), most of the fluvial physils are deposited in deltas, estuaries and marshes;
however, there are physils out there.
Holocene sedimentation rates are generally highest adjacent to the continental
coasts. Sedimentation rates as high as 1000 cm/1000 years are reported in the
Mississippi Delta and 500 to 600 cm/1000 years for the Rhome and Orinoco deltas.
280

Rates on the shelves are commonly less than 10 cm/1000 years. Rates in the deep
ocean are on the order of 0.1 to 1 cm/1000 years (Lisitzin, 1972).
Once the suspended material escapes the rivers and estuaries it is exposed to
various marine currents which determine how much is trapped on the shelf and how
much escapes to the deep ocean. In locations with moderate wave activity the
coastal zone down to a depth of about 10 m is covered by sand and the mud zone
( < 63 pm) occurs seaward of the sand. Sands, commonly relict glacial deposits, also
occur seaward of the mud zone. The location of the mud zone depends on the
degree of coastal exposure. In some areas, commonly adjacent to large deltas,
coastal mud flats extend into the tidal flats (Louisiana). Off some deltaic areas, such
as the Mississippi, muds blanket the whole shelf (McCave, 1972).
Beyond the shelf the sediments are mainly fine-grained. They are transported
either in suspension or by turbidity currents. Turbidites are important components
of the continental rise and parts of the abyssal plains.
The old concept of a seaward-fining of sediments (sand on the shelf, silt on the
slope and clay on the rise) was based on limited sampling. More detailed sampling
indicates the mudline, newly defined as the depth at which the proportions of clayey
silt, silty clay and clay reach a near maximum value and the proportions no longer
increase significantly with depth (Stanley and Wear, 1978), has a variable relation to
the shelfbreak. Stanley et a/. (1983) describe four general types of mudline-shelfbreak
relations :
Type I: The mudline occurs at considerable depth on the upper- and, in a few
cases, mid-slope (300 to > 1000 m).
Type 11: The mudline occurs below the shelfbreak on the uppermost slope (to
approximately 200 to 400 m).
Type 111: The mudline is nearly coincident with the shelfbreak (the latter ranging
from about 100 to 175 m in many instances).
Type IV: The mudline occurs considerably shoreward of the shelfedge ( < 100 m).
Fig. 5-1 shows the relation of the mudline to the shelf break along the continental
margin of northeastern United States. The position of the mudline is controlled by
shelf width, sediment supply and magnitude of fluid energy. Type I prevails under
conditions of high energy and low sediment supply. Type I1 develops under
conditions where energy levels control deposition more effectively than does sedi-
ment supply. Type I1 occurs where more sediment is supplied than fluid processes
can effectively erode. Type IV results when there is a large sediment supply but
insufficient energy to transport it offshelf (Western Gulf of Mexico). In many cases
the mudline serves as an energy-level marker that helps define the boundary
between erosion and depositon.
The concentration of suspended sediment in surface waters decreases roughly
exponentially in a seaward direction. The order of magnitude of these concentra-
tions changes from 10-100 mg/l nearshore to 0.1-1 mg/l in the deep ocean. The
average value for the World Ocean is 1 mg/l or 1 ppm (Lisitzin, 1972). The
concentration values refer to the total suspended material, which can contain from
+
less than 10% biogenous material (CaCO, SiOZAm(,rph + COrg)in the nearshore
areas to > 50% in the open ocean. Emery and Honjo (1979), based on optical
281

-
MEAN SAND
w
Y///A <1
0
10 - 20
20-30
___ '30

36" 79O 78" 35O 77O 34" 76O 750 330 740
Fig. 5-1. Continental margin off northeastern United States showing distribution of sand on the outer
shelf and slope (percentage mean sand data from Keller et al., 1979; modified with permission). Arrows
depict examples of mudline types discussed in text: Type I (Cape Hatteras margin), Type I1 (most of the
margin between Norfolk and Hudson canyons), Type 111 (shelfedge at head of Hudson Canyon). Depth
in meters. From Stanley et al., 1983. Copyright 1983 Soc.Exon. Paleo. Miner.
282

Fig. 5-2. Concentrations of suspended matter in surface waters of the North Atlantic during spring.
autumn and winter. From membrane ultrafiltration data. From Lisitzin, 1972. Copyright 1972 SOC.Econ.
Paleo. Miner.

observation, concluded that mineral grains comprised only 0.4 to 2.5% of the surface
suspended matter in the ocean off the east coast of Africa. Most others report
higher values. Fig. 5-2 shows a typical concentration pattern for the North Atlantic.
Superimposed on these patterns are plumes of higher concentration, extending from
the mouths of major rivers, across the shelf, to the continental slope. Plumes of
suspended sediment can also occur where there is a convergence of coastal currents,
creating a current across the shelf (McCave, 1972).
283

The absolute mass of suspended material under each square meter of ocean
surface can be calculated from the weighted average concentration of the water
column and water depth. Over shelves the absolute mass is usually less than 1000 g
because of small water depths. The highest values, 4000 to 10,000 g, occur over
continental slopes. In some open ocean localities the amount of suspended material
ranges from 3000 to 6000 g. Absolute suspended mass values correlate well with
climatic zones, being greatest in the humid zones and least in the arid zones
(Lisitzin, 1972). Lisitzin goes on to calculate that the World Ocean contains
1.370 X lo'* tons of suspended material. The annual discharge of suspended and
dissolved substances from land is 1.270 X 10". Thus, the suspended stock in the
World Ocean is equivalent to 80 to 100 years of total world terrigenous sediment
discharge, provided the entire dissolved discharge is transformed into particulate
matter by organisms. He also notes that the transport between oceans is 1.679 X 10"
tons per year, approximately as much as is eroded from the land surface each year.
It is surprising the ocean bottom physil suites are not more homogenized.
On the continental shelf suspended sediments are moved by processes of eddy
diffusion and of advection (McCave, 1972). The diffusion process is generated by
tidal motion and wind-generated waves and transports sediment from areas of
higher concentration to one of lower concentration, seaward. Advective transport
means movement of suspended material by a net horizontal water movement, such
as is created by the upwelling of colder, more saline waters onto the shelf,
estuarine-type circulation caused by the shoreward movement of bottom waters in
response to a seaward flow of lighter surface water, deltaic plumes, residual currents
formed from adjacent oceanic currents, and longshore currents generated by wind
and waves. Though these processes can transport material in opposite directions
there is a net seaward transport and diffusive escape from the shelf system.
Residual currents account for most mud transport; examples include the Amazon
mud stream along the Guiana coast (Fig. 4-50) and the Mississippi River mud
deposited in the western Gulf of Mexico.
Curray (1965) suggests that with the recent decrease in the rate of sea level rise,
muds have begun to spread across the shelf floors. He notes that these blanket muds
rarely extend farther than 30 to 40 km from shore, where they pinch out into the
underlying relict sands. Though some mud is deposited in the outer shelf, the rate is
so slow that much of it is worked into the sand by burrowing organisms and current
surges. These sands are characteristically muddy and mottled.
Shelves have been classed as autochthonous or allochthonous (Swift, 1973).
Autochthonous shelves are those on which the sediments on the shelf, largely
Pleistocene deposits, are reworked. Tidal currents and wave activity winnow physils
from the sands and deposit them in the sheltered low area. The mud distribution
tends to be patchy. An excellent example is the storm-dominated shelf off the east
coast of the United States. The low mud content of this area is in part due to the
rapid post-Pleistocene rise in sea level, which caused much of the fine sediment to
be trapped in the estuaries (drowned river mouths) and marshes.
Allochthonous shelves receive their sediment primarily from rivers. In addition to
what has been said about sediment-shelf transport, another factor should be
284

--
Ice Temperate Arid Tropical humid
humid
-v

Fig. 5-3. Dependence of grain size composition of biogenous and temgenous sediments upon climatic
zonality. From Lisitzin, 1972. Copyright 1972 SOC. Econ. Palm. Miner.

mentioned. The seaward transport of the majority of suspended fines is more often
periodic than continous. During periods of calm weather much of the fine material
can settle to the bottom only to be resuspended during storms. Scruton and Moore
(1953) noted a hundred-fold variation in turbidity in areas near the Mississippi
Delta, between calm and stormy weather.
Once physils are deposited as muds, electrostatic attraction, van der Waals
attraction, pelletization, and binding by organic material makes erosion relatively
difficult. With increased consolidation the shear stresses required for erosion,
increase. McCave (1972) calculated that mud should be deposited when current
velocities are less than 17 to 25 cm/sec at 1 m above the bed. At sea, this
corresponds to a surface velocity of a little less than 50 cm/sec (1 knot). He points
out that virtually any area of the shelf can, at times, have velocities < 25 cm/sec
and can, at least temporarily, be potential sites of accumulation.
Both the total amount and grain size of terrigenous material on the shelves can
change sharply in different climatic zones (Fig. 5-3). Sands and silts are predomi-
nant in the ice and cold humid zones. The amount of pelitic material ( < 10 pm) and
physils is at a minimum off the desert areas and reaches a maximum in the tropical
humid zone where chemical modification is highest (Lisitzin, 1972).
285

Slope, Rise and Abyssal Plain


Sediment / Distribution
By the various mechanisms discussed in the preceding pages, some fine-grained
suspended material reaches the edge of the shelf. Where it goes from there is the
next question. McCave (1972) made the following calculations. As the average rate
of sedimentation, on a carbonate-free basis, for the deep sea is about 0.4 g/cm2/103
years and the deep sea covers 270 X lo6 km’, the annual deposition is about
1.1 X lo9 tons. This is a small fraction of the estimated fluvial suspended sediment
supply of 18 X lo9 tons (Holeman, 1968) (a more recent estimate is 13.5 x lo9 tons
by Millman and Meade (1983). McCave assumes that about 50% of the fluvial
supply stays in coastal areas and on the shelf. Thus, about 9 X lo9 tons (or 7 x lo9
tons) escapes the shelf edge of which 8 X lo9 tons (or 6 x l o 9 tons) should be
deposited on the continental slope and rise. The calculated amount deposited on the
rise, based on 5 g/cm2/103 yrs and 20 X lo6 km’, is only another 1 x lo9 tons.
McCave suggests the remaining 7 X lo9 tons (or 5 x lo9 tons) occurs as great fans
and cones of sediment seaward of the major rivers. Thus, the bulk of suspended
sediment must be moved in advective plumes off the major input points.
However, for the Yellow River of China, which has the second highest sediment
discharge in the world, only 24% of the reported discharge (the gauging stations are
upstream from the river mouth) actually reaches the ocean (Milliman and Meade,
1983). This suggests as much as 70 to 80% of the fluvi’al suspended sediment is
deposited shoreward of the shelf slope break. Using a shelf escape value of 25% and
13.5 x lo9 tons of fluvial material, then only 1.3 X lo9 tons need be accounted for
in fans and cones. The truth is probably somewhere in between.
One of the main problems in depostion beyond the shelf is how physils ever get
to the bottom. For the time being we will assume that they do get there. Both
sedimentation rates and the absolute amount of sedimentation are greatest on the
slope and rise, with values increasing towards the base of the slope. For example,
the thickness of Holocene layers on the shelf off eastern North America ranges from
20 to 35 cm; at the base of the continental slope it is 300 to 500 cm; farther out in
the abyssal plain it gradually diminishes to 30 cm and less. Sedimentation rates are
usually much higher near the mouths of large rivers, with values as high as 1000
cm/103 years reported for the Mississippi Delta (Lisitzin, 1972).
The distribution of the major sediment types is shown in Fig. 5-4. Terrigenous
material covers much of the Ocean floor but sizable areas are covered by biogenous
material. In general the proportion of muds and physils in the terrigenous facies
increases from shallow to deep water. The pelitic fraction ( < 10 pm) usually makes
up more than 70% of pelagic sediments, with physils being the major component.
Fig. 5-5 shows the distribution of the < 10 p m fraction in the Pacific Ocean. The
values for the Atlantic Ocean are probably lower. The sediments on the Atlantic
continental slope are approximately 60% mud (Shepard, 1959).
Transport Mechanisms from the Shelf
After crossing the shelf edge the fine grained sediment arrives at its depositional
site by gravity settling and turbidity or lutite flow. These turbid currents commonly
286

Fig. 5-4. Grain size of the bottom sediments of the Atlantic Ocean. From Lisitzin, 1972. Copyright 1972
SOC. &on. Paleo. Miner.

move down submarine canyons to the relatively flat sea floor of the continental rise
or the abyssal plain. Successive deposits will build a deep-sea fan. Coalescing fans
cover much of the continental rise. Large, composite fans are called cones. Abyssal
cones occur seaward of many large rivers, including the Mississippi, Amazon,
Congo, Ganges and Indus. These deposits contain appreciable sand as well as mud
and usually show graded bedding. Sands and pebbles are most abundant in the
main channel nearest the base of the slope with the mud content increasing radially
outward from the mouth of the canyon.
The Mississippian Fan or Cone has been studied rather extensively. It consists of
a lobe-shaped prism of muddy sediments extending for about 600 km from near the
287

Fig. 5-5. Percentage distribution of pelite fractions ( < 1Opm) in Pacific sediments. From Lisitzin, 1972.
Copyright 1972 Soc. Econ. Palm. MIner.

Mississippi Delta to the abyssal plain (Stuart and Caughey, 1976). The fan covers an
area over 300,000 km2 and has a sediment volume of over 290,000 km3. During the
Pleistocene the Mississippi Trough, on the shelf edge, apparently channeled detritus
across the outer shelf and upper slope and acted as a point source of fan sediments.
Near the outer shelf-slope area the Pleistocene sediments are approximately 4600 m
thick. They thin to the south and are over 300 m thick in the lower part of the fan.
The fan is composed primarily of gray-green clay with thin beds of silt and sand.
The sediments are largely turbidites and pelagites (settling). Holocene muds are
relatively thin. The lower Pleistocene sea level allowed much of the Mississippi
River detritus to be transported across the shelf to deep water environments.
Following the rise in sea level much of this material has been trapped in the coastal
areas and transported laterally by marine currents.
288

The Amazon Cone, which is on the continental slope and rise of the western
flank of the Atlantic Ocean, is similar in many respects to the Mississippi Fan
(Danuth and Kumar, 1975). It is 650 to 700 km from the shelf edge to the Abyssal
Plain; the cone is 380 km wide along the shelf and 600 km wide near the base.
Thicknesses range from 9 to 11 km at the head to 1 to 5 km near the base of the
cone. Sediments were delivered by the Amazon River via the Amazon submarine
canyon during periods of low sea level. The sediments are primarily dark-gray to
olive-gray clay with a lesser amount of interbedded sand and silt. Holocene
sediments, marl and ooze, are less than a meter thck. Much of the sediment was
deposited during the Pleistocene, but the cone began to form in the Miocene.
These two examples are typical of the sedimentation pattern along most coasts.
During periods of relatively high sea-level stand, muds, along with silts and sands,
tend to be trapped on the continental shelf. With a eustatic lowering of sea level
much of the mud is transported to the deeper water slope and rise. Thus, the loci of
mud, and physil, deposition can oscillate between shallow and deep water environ-
ments over relatively short periods of time.
A note about mud fabrics. O’Brien et al. (1980) found, from SEM studies, that
the physils in hemipelagic physilites were well oriented, whereas those in turbidites
possessed a random orientation (Fig. 10-3). Preferred orientation is normally
produced when the concentration of physils in suspension is low (flocculation is at a
minimum) and deposition is slow, the situation that prevails in a hemipelagic
environment. The random orientation in the turbidites may be due to rapid
sedimentation but more likely indicates the clays were in a flocculated state. I t is
likely that the physil concentration in turbid bottom flows is high enough that
flocculation can occur. O’Brien er al. (1980) found the physil suites of the hemi-
pelagites and turbidites were similar. Other studies have shown the physil suites of
these two types of beds can be significantly different. Whether or not there are
differences would depend on whether more than one source was supplying physils
and on the current patterns.
Though turbidity currents were an important transport mechanism during the
Pleistocene and for many ancient deep sea deposits, many of the Holocene (past
11,000 years) fine-grained deposits show no evidence of turbidite deposition. Much
of this material in these deposits apparently settled from the nepheloid layer (Stokke
et al., 1977).

Nepheloid Layer and Deposition


Near the bottom of large areas of the ocean floor, particularly areas containing
cones and fans, there is a thick layer of fine-grained suspended material called the
nepheloid layer. This layer may be more than 2 km thick. Sediment concentrations
can vary widely but are typically much lower (0.01 to 0.3 mg/l) than in estuaries
and nearshore waters. The sediment is maintained in suspension by relatively strong
bottom currents (Pierce, 1976). The suspended sediment concentration is de-
termined indirectly, using a light-scattering nephelometer. Though there are enor-
mous volumes of sediment in the nepheloid layer, calculations indicate that if it
were all deposited at once and compacted to a density of 2 g/cm3 (dense mud) the
289

47OOC
Fig. 5-6. Thickness of the bottom nepheloid layer (BNL) on the Nitinat Fan, Washington. From Stokke
et al., 1977, Geol. Soc.Amer. Bull., 88, 1586-1592.

layer would be from 15 to 150 pm thick. This would be a thin layer, but over
geologic time could account for thick deposits of shale.
Stokke et al. (1977) reported on the character of the nepheloid layer associated
with the Nitinat Fan. The Nitinat deep-sea fan is located in the northern Cascadia
Basin, offshore from the state of Washington. Fig. 5-6 shows the thickness of the
bottom nepheloid layer (BNL) overlying the fan. The cross-section in Fig. 5-7 shows
the relation of the nepheloid layer to bottom topography and the systematic increase
in suspended sediment concentration with depth. Thickness and concentrations
( < 50 to > 200 pg/cm2), which are closely related, are greatest in the northern part
of the area adjacent to three submarine canyons (Barkley, Nitinat, Juan de Fuca)
and northwest-southeast aligned Cascadia Channel. The bottom low-density turbid
290

BNL suspended
sediment load (wg/cm2)
0 Top of BNL.
loom] 110mg/crn2.yr
-22.0 150-200 0 L.H. ACC. rates
50-100 >ZOO km
0 E.H. ACC. rates V.E. = 3.7x
100-150

Fig. 5-7. Cross section through the bottom nepheloid layer (BNL) and Nitinat Fan, showing the
relationship between the top of the nepheloid layer, the underlying bathymetry, and late Holocene and
early Holocene accumulation rates. Suspended sediment load is integrated from the top of the nepheloid
layer down, so that the total load at a given station is indicated by the load immediately above the sea
floor.Stokke et al., 1977, Geol. Soc.Amer. Bull., 88,1586-1592.

waters evidently emanate from the three canyons and preferentially move along the
Cascadia Channel. Silt-sized material is deposited primarily in the channels and
clay-sized material in the interchannel areas. Medium grain size of the interchannel
muds is approximately 2 pm. However, the size determinations were made on
dispersed samples and it is unlikely the size values are those of the original
sedimented particles (or aggregates).
The nepheloid layer remained relatively constant over a period of four years,
indicating it is not episodic, as are normal turbidity flows. The accumulation rates,
< 2.5 to > 12.5 mg/cm2/yr, are closely related to the thickness and suspended
sediment load. This indicates the nepheloid layer material is not primarily resus-
pended bottom sediments and that there is a net downward flux of particles within
the nepheloid layer.
The size of the particles in the nepheloid layer is not well known. A clue to the
size of the suspended physils is provided by a study of suspended and bottom
material from the offshore Washington area. Baker (1973) found that the physil
suite of the bottom suspended load (total sample) more closely resembled the physil
suite of the < 62 p m fraction of the underlying bottom muds than it did the < 20
and < 2 pm fractions. This suggests much of the suspended material occurs as
relatively coarse flakes and/or aggregates. In these samples chlorite and illite are
concentrated in the coarse fraction and montmorillonite in the fine fraction.
Fig. 5-8 shows the size spectra of a series of suspended samples collected on the
Nova Scotian continental rise (McCave, 1983). In the nepheloid layer, the size
distributions are unimodal with a peak at 4 pm. A second mode at 16 is present in
291

e
1.75 1.80 1.85
, , , , I

480C

300

4900

zoo
cn E
0
P c
n 9
al al
3 B
5,000 %
t cn
L
al
I00
r

510C

Pm
Fig. 5-8. Vertical profile on Nova Scotian continental rise showing the potential temperature (left). light
attenuation coefficient (c, right) and particle size spectra from the indicated levels, in the lower 350 m of
the water column. Numbers under each spectrum give the total measured particle volume (ppb,
mm3/m3). The top two spectra are from 500 to 1000 meters above bottom. From McCave. 1975.
Copyright 1983 Amer. Geoph. Union.

the shallower samples. Much of the material in the fine mode is aggregates and that
in the coarse mode is single solid particles. Measurements were made using a
Coulter Counter which, unfortunately, only counts particles in the 1.26 to 32 p m
range. The regional and temporal distribution of the material in the nepheloid layer
suggests a substantial amount of the detritus was from lateral injection of material
resuspended from the continental rise or from rough topography within basins
where there are locally fast currents.
Fig. 5-9 shows the concentration of suspended material in the nepheloid layer of
the Atlantic Ocean. The high suspended loads in the Western Atlantic are believed
to be caused by the western boundary currents, which diminish in intensity
equatorward (Biscaye and Eittreim, 1977).
There are differences of opinion on how physils get to the bottom of the deep
ocean. McCave (1975) states that most of the particles suspended in seawater have a
292

Fig. 5-9. Distribution of suspended material in the nepheloid layer in the Atlantic Ocean. Note the high
concentrations found only in the western portions of the ocean overlying the continental rise. From
Biscaye and Eittreim, 1977. Copyright 1977 Elsevier.

diameter ( 2 pm, and over much of the ocean floor the < 2 pm fraction is
dominant. It needs to be pointed out, again, that we have no idea of the deposi-
tional grain-size of the physils in muds and physilites. The process of measuring the
293

grain-size requires that the original texture be disrupted and the measured size, for
muds, is normally finer than the depositional size; the reverse may be true for
physilites. The extreme case is smectite. Well dispersed smectite flakes are normally
< 0.2 pm in diameter and should virtually never settle to the bottom, yet large areas
of the ocean floor are covered with smectite.
Simple particle settling is considered by many to be the normal depositional
mechanism (Jacobs et al., 1973); however, the inhomogeneity of the ocean floor
physil suites and the close association of bottom physil suites with nearby input
suites indicates settling is commonly more rapid than can be accounted for by single
particle (flake or plate in the case of physils) settling. Accelerated sinking is required
(McCave, 1975).
Jacobs et al. (1973) collected samples from clear water and nepheloid water from
the Atlantic Ocean and obtained an average suspended sediment concentration
+ +
value of 0.1 0.1 mg/l for clear water and 0.5 0.1 mg/l for water from the
nepheloid layer. Based on partial count they found that non-opaque mineral grains
(presumably largely physils) comprised 85 to 91% of nepheloid layer suspensions
and 76 to 86% of the clear water suspensions (deep water from near Mid-Atlantic
Ridge). They further reported that 85 to 96% of the particles in nepheloid water
were < 2 pm; the value for the clear water was 76 to 87%. Aggregates comprised
from 1 to 20% of the samples. The problem with the size values is that the samples
were collected on filter paper and repeatedly washed with distilled water. This
would disperse, or partially disperse, many of the physil aggregates and provide
erroneous size data. Pierce (1976), using a SEM, found that 38 to 100% of the
particles in the nepheloid layer were aggregates, though many were relatively small.
Abundant large physil-organic aggregates, 0.03 to 0.26 mm in diameter were found
off the coast of Monaco (Kane, 1967). The quantity of surface aggregates decreases
seaward. The physil-organic complexes are colonized with bacteria which tends to
cause an increase in density; the aggregates 10 to 100 m below the surface are more
distinct than surface aggregates. Presumably some form of consolidation occurred.
Settling, suspended sediment was collected in a sediment trap, at a depth of
3200 m in the Sargasso Sea, every two months for a period of two and one-half
years (Deuser et al., 1981). The biological material and the physils (based on A1
analysis) displayed a pattern of variations similar to the annual cycle of primary
production in the Sargasso Sea. The authors concluded that the physils were largely
present as fecal pellets and other organic aggregates created by filter feeders. The A1
values indicate physils comprise on the order of 20% of the < 37 pm fraction and
less of the total sample. Others report values as high as 70% were found in the Gulf
of Mexico (Manheim et al., 1972). The physil flux, at 3200 m, ranged from 105 to
470 g/m2/day. The atmospheric deposition values are similar, suggesting the
atmosphere was a major source of physils. Calculations indicate that the bulk of the
sediment moved from the source to 3200 m in less than 60 days.
Numerous analyses of particle size have been made using a Coulter Counter
(Carder et al., 1971). These studies all demonstrate the relative abundance of
fine-grained particles. Unfortunately, the technique excludes particles > 100 pm in
diameter. A spherical aggregate 10 p m in diameter, with 50% porosity, could
294

Fig. 5-10. Aggregate composed of physils, bacteria (rods), and excreted fibril mucilogenous appendages.
Bacteria approximately 1.5 pm. Courtesy H.W. Paerl.

contain approximately 2,600 1 p m plates (D/T = 10); a 100 p m sphere could


contain 2,600,000 plates. A Copepod fecal pellet with a diameter of 80 p m and a
length of 250 pm (typical) and 50%porosity could contain approximately 6,000,000
1 pm plates. Even if these large particles contain only 25 to 50% physils (and
quartz), they can account for a large part of the mass of suspended physils.
Studies of near surface suspended material in Lake Tahoe indicated the surfaces
of the particles, both organic and inorganic, commonly contain living fungi and
bacteria ( < 0.5 pm). Extensive “webbing” was observed associated with both
microorganisms (Fig. 5-10). Particles are trapped by these web-like structures and
increase in size (Paerl, 1973). In a follow-up study of suspended material from other
lakes and the Pacific Ocean, Paerl (1975) found that bacteria had a variety of
attachment mechanisms. “Among these are adhesive stalk formation, capsular
secretions, fibrillar appendages which serve as anchors, attached webbing on which
cells are located, and sorption of bacterial cells to particles without the aid of
cellular appendages or secretions.” He also established that the particles need not
contain organic material in order to attract the microflora and that at a depth of
> 100 m microbial attachment showed a marked decrease.
Organisms and organic matter apparently supply the binding to make physil
aggregates cohesive and allow them to settle through moderate currents. These
295

Table 5-1
The egestion rate of Tigriopus californicus for various mineral suspensions indicating the number of
pellets counted (N), the number of pellets per day produced by one copepod (NPD), reaction time (RT)
and mean pellet size (MPS). The reaction time equals the maximum time some particles would reside in
the digestive tract assuming an even rate of pellet ejection. From Syvitski and Lewis (1980).

Mineral N NPD RT (hours) MPS ( r m )


Montmorillcmitte 1900 19 1 140
1300 13 2 125
Mite 1200 12 2 120
700 7 3 110
Clinwhlore lo00 10 2 120
Tremolite 790 8 3 120
Vermiculite 700 7 3 115
Muscovite 600 6 4 110
600 6 4 110
Kaolinite 420 4 6 105
Microcline 300 3 8 100
Quartz 210 2 11 100

aggregates typically have a density range of 1.2 to 1.6 g/cm3. The relatively high
organic content of these aggregates indicates large amounts of organic material must
be destroyed after deposition, possibly by bacterial action.
Recent studies leave little doubt that most of the physils deposited in the deep
sea, other than by turbidity currents, arrive there in the form of relatively large
physil-organic aggregates. The aggregates may be formed by filter feeders (fecal
pellets), by a microflora webbing, or by physil-organic “flocculation.”
Particle count studies overestimate the importance of fine particles and in most
instances do not detect the large particles.

Pellets
The importance of biogenic activity in estuaries and the open ocean in con-
centrating suspended physils and other material into pellets has been discussed.
Pelagic filter-feeders pelletize fine material in suspension, commonly near the
surface, thus enabling physils to be deposited in areas where currents are weak to
moderate. Not only do the filter-feeders aggregate the physils, they can cause serious
chemical and mineralogical changes.
Studies of zooplankton raised in captivity indicate they can strongly modify the
ingested physils (Syvitski and Lewis, 1980). The copepod Tigriopus californicus was
maintained in an aquarium and allowed to feed on a variety of minerals. The
copepods showed a definite mineral preference. Table 5-1 shows the rate of pellet
production, the maximum time the particles remained in the digestive tract and the
pellet size. Montmorillonite was the preferred feed. They ingested particles ranging
in size from 0.5 to 50 pm (flocs).
Before and after x-ray patterns indicate some of the minerals were significantly
altered by the zooplankton (Fig. 5-11). The mineral changes are interpreted from the
published x-ray patterns and differ somewhat from the authors’ interpretation. The
296

Sample ( 550.C 1

Standard ( 5 5 0 ~ )

3.56 z10

Sample (300.C 1
17

Standard glycalated 14
A

3
Standard (20.C)

3.55
w h
7.1

Sample glycalated
Fig. 5-1 1. X-ray patterns of starting montmorillonite with some interlayer materia! (standard) and pellite
residue (sampp) which appear: to be composed of kaolinite (7.1 A and 3.55 A), vermiculite (14 A),
brucite (4.74A), and illite (10A). From Syvitski and Lewis, 1980. Copyright 1980 Soc. &on. Paleo.
Miner.

montmorillonite appears to be altered to a mixture of kaolinite, vermiculite and


brucite (4.74 A). The tremolite was partially altered to a chlorite. The biotite-
vermiculite feed lost all crystalline reflections and the Mg/Al, K/Al and Fe/Al
values were decreased from 26 to 61%. The muscovite peaks broadened and the
001/002 ratio increased, indicating the loss of K and the formation of some
expanded layers. The odd order reflections of the clinochlore sample decreased. This
was accomplished by a decrease in the Mg/Al ratio and an increase in the Fe/Al
ratio. Presumably some Mg was replaced by Fe. These are drastic changes. If
zooplankton make significant contributions to the bottom sediment, a lot of
reinterpretation will be required. It may be that the lack of organic coatings on the
297

laboratory physils allowed chemical reactions much more severe than would occur
under natural conditions.
Many benthos or bottom dwellers are also effective pelletizers. These bottom
dwellers may be filter feeders or mud-eaters (deposit feeders).
In the Dutch Waddensee, mussels and cockles occur in numbers averaging more
than l,000/m2, with values as high as 100,000 cockles per square meter. These
organisms feed by pumping 1.4 1 of sea water per hour through their feeding
systems, for 12 hours per day. In the process they remove 420 mg of material from
suspension daily. Assuming a concentration of only 20 individuals per square meter,
the mussel Mytilus edulis annually produces 25 million kg dry weight of fecal pellets
in the western Waddensee. This is equivalent to a layer 0.25 mm thick extending
over 600 km2 (Verwey, 1952).
Strakhov (1969) reported areas where the sea bottom contained an almost
continuous layer of mud-eating worms (Tubificidae). In a single day they pass
through their intestines from 3 to 12 kg of mud, or 4 to 6 times the bulk of the entire
biomass. Approximately half the weight of worms from the Bay of Naples is due to
mud in the alimentary canal.
In the Firth of Clyde (Moore, 1936) and Cape Cod bays (Johnson, 1974) pellets
form more than 408 of the sediment. Fecal pellets are present in nearly all samples
off the Nigeria1 coast; in some samples they comprise almost 100%of the sediment
(Porrenga, 1966).
Some of the Nigerian pellets have a physil suite similar to that of the matrix
mud; however, the physils in most of the pellets have been altered to an Fe-rich I/S
with 70 to 80% smectite (glauconite) or to a poorly crystallized Fe-rich chlorite
(chamosite). Montmorillonite and kaolinite are the main detrital physils. The
alteration is apparently facilitated by the abundance of detrital Fe oxide produced
under the tropical weathering conditions that prevailed.
A more specific study of pellet formation and accompanying mineralogic changes
was made by Pryor (1975) in an area off the coast of Georgia and the northeastern
Gulf coast. Two major filter feeders in these areas are the Callianassa major
(burrowing “ghost shrimp”) and Onuphis microcephala (a polychaete burrowing
worm) that inhabit shallow marine and lagoonal-estuarine environments. The action
of the Callianassa is highly visible. They live primarily in burrows in sandy
beach-foreshore zones where they feed by occupying the upper part of vertical
tunnels and setting up in currents of water with a rapid motions of their swim-
merets. The suspended detritus is sifted out, by hairy maxillipeds, and passed into
the mouth. They feed for an average of 18 minutes and then for 3 minutes discharge
fecal pellets, at the rate of 40 pellets per minute. Their place is then taken by
another shrimp. The pellets are approximately 2 mm long 0.75 mm in diameter, with
densities of 2.35 to 2.53 (Fig. 5-12). They are composed of 90 to 98% minerals,
largely physils, with some fine quartz, and approximately 3% organic carbon (the
organic materials are largely bacteria, diatoms, and algal cells). They are firm and
resistant to abrasion.
An average of 2,480 pellets per day are produced by a Callianassa major. In the
Sapelo Island (Georgia) beach-shoreface, the Callianassa population alone pelletizes
298

Fig. 5-12. Callianassamajor burrrow entrance and recently discharged fecal pellets. Pellets approximately
2 mm long. Sapelo Island, Georgia offshore bar surface. From Pryor, 1975, Geol. SOC. Amer. Bull., 86,
1244-1254. Courtesy W.A. Pryor.

12.3 metric tons of suspended material per year. This is enough to cover the area to
a depth of 4.5 mm in 1 year. The equivalent value for the Onuphis population is 3.7
mm per year. In back island areas, where the population density is greater, the
production is 141 mm of fecal pellets per year. The pellets are moved along the
bottoms and deposited in ripple troughs and swash zones on the beach. Because of
differences in size and density selective sorting occurs. Callianassa pellets are
transported and deposited with medium sand-sized quartz, and Onuphis pellets with
fine sand. On Sapelo Island portions of the lower beach face are covered by as much
as 30 cm of Callianassa fecal pellets; farther along the shore, in a lower energy
environment, Onuphis pellets accumulate in thicknesses up to 30 cm. These accumu-
lations can contain 5 to 10% organic material.
The Georgia offshore area is covered with relect sands and few if any physils are
being deposited by inorganic means. Physils are being deposited primarily in the
form of fecal pellets. When these pellets are compacted slightly, water is expelled
and they lose their integrity; their recognition in sedimentary rock is difficult or
impossible. Thus, the presence of thin physilite layers or beds in high energy coastal
299

K I CMMx I CMMx I CMMx

T”LJ

I’.J
Suspended i
I
Suspended Suspended

”J

Collionorra Collionossa Collionarro


Dauphine Ponorno City

H-12

Collionossa Onuphir
Pcnsocolo

F-5

M issirsippi Sound Florida Gulf SOpelo Island


Fig. 5-13. Representative X-ray diffraction patters of suspended clays and related fecal pellet clays. From
Pryor, 1975, Geol. Soc.Amer. Bull., 86, 1244-1254.

sandstones can not arbitrarily be interpreted to indicate periods or areas of low


energy deposition.
Pryor also established that the physil suite in the pellets were different from those
suspended in the overlying water and that the pellets of different organisms had
slightly different physil suites (Fig. 5-13). In general chlorite appears to be de-
stroyed, the “crystallinity” of illite and kaolinite is reduced (x-ray peaks are
broadened) and the montmorillonite and I/S appears to be slightly modified. The
published x-ray patterns suggests some of the expanded layers contracted, presuma-
bly by fixing K. The difference in the physil suites of the suspended and pelletized
material points out another complication that can arise when comparing suspended
physils with bottom muds and then attempting to interpret the differences in terms
of multiple sources and currents.
It is not known what goes on inside the organisms but the potential is there. “...
mouth regions are nearly neutral, stomachs and midguts are acid, and the hindguts
300

are alkaline. Intense physical reduction of ingested material takes place in the large,
powerful gastric mills of the midguts. In addition to these harsh and alternating
conditions of acidity, there are numerous digestive enzymes (proteases, lipases and
amylases) and large numbers of intestinal bacteria whose effects on clay minerals
are largely unknown but probably important.” (Pryor, 1975).
It is most likely that pelletization and the associated modification of the physils is
the beginning step in some glauconitization processes (Chapter VI). Shortly after the
pellets are discharged the particulate organic material is rapidly converted by
bacteria and the pellets become infested by diatoms. The decay of the bacteria and
organic material presumably creates a microreducing environment, which reduces
some of the Fe, increasing its mobility, and begins the construction of Fe-rich
physils such as glauconite and chamosite.
Pryor noted an additional process. Mullets, crabs, shrimp, etc. were observed
vigorously eating the Callianassa fecal pellets (coprophagy). Repeated coprophagy
presumably wreaks havoc with the physils.

ATMOSPHERIC TRANSPORT

Eolian

Eolian and volcanogenic pyroclastic materials (aerosols) are transported in


similar ways but have different origins and will be treated separately. On the
average aerosols are removed from the atmosphere at about the same rate as they
enter it. World-wide production of natural aerosols, exclusive of volcanic input,
amounts to approximately 21 X lo4 g/year, of which 24% is windblown dust, of
which approximately half is finer than 5 p m (Wallace and Hobbs, 1977). Garrels
and MacKenzie (1971) estimated that about 0.54 x l O I 4 g/year of wind-storm-
transported material falls on the ocean. Slowly settling dust from the stratosphere,
which is assumed to have cm3 of dust/cm3 of air (equivalent to 10,000 physil
particles 1 pm square by 0.1 pm thick), falling on the ocean, amounts to 0.074 x l O I 4
g/year. Thus, a total amount of atmospheric dust that reaches the ocean is
approximately 0.6 X l O I 4 g/year (6 X lo7 tons/year). This is 20-30% of the rate of
sedimentation of fine-grained materials on the abyssal plains. More recent estimates
of deposition rate are an order of magnitude higher, 5 to 8.5 X l O I 4 g/year (Table
5-2). The higher value is equal to the estimated global annual deposition rate for
detrital clay minerals in recent pelagic sediments (Prospero, 1981). The annual
atmospheric flux is equivalent to 4 to 6% of the total river sediment discharge.
Aerosols are distributed in three general ways: local fallout, tropospheric fallout
(maximum elevation of 11 km), and stratospheric or global fallout (above 11 km).
Local fallout is influenced by temperature inversions which serve as “lids” which
prevent the upward movement of aerosols. Tropospheric transport occurs at consid-
erable heights and is primarily latitudinal. Transport range depends on wind height
and force, size, specific weights of particles and atmospheric precipitation.
301

Table 5-2
Mineral Aerosol Deposition - Estimates for Various Ocean Regions (after Prospero, 1981).
Ocean Region Deposition Rate
x g/cm2/yr x 10’’ g/yr
North Atlantic north of trades 82 12
North Atlantic trades - 100-400
South Atlantic 85 18-37
Pacific 31 66
Indian Ocean 450 336
All Oceans (minimum-maximum) 532-851

The conditions most favorable for the capture of aerosols occur in the arid zones
where the air is dry and constant tradewinds exist. As a result, the bulk of the
distant dispersal material migrates within the arid zones, mostly from east to west.
In the humid zone aerosols are rapidly washed from the air.
The greatest transport distances occur in the stratosphere, where jet stream
velocities are 300 to 500 km/hour and fine particles may circle the earth for years.
Studies of stratospheric radioactive fallout indicate the most favorable areas for
deposition are the arid zones. Thus, the stratospheric aerosols accumulate in narrow
bands of latitude rather than uniformly over the earth (Lisitzin, 1972). The winds
are capable of keeping clay size material in suspension for long periods of time;
deposition from the air is mainly by rain or snow.
The oceanic areas receiving the greatest mineral aerosol fallout are those adjacent
to major deserts and arid regions; however, there are large seasonal and annual
changes in aerosol concentration. The frequency of occurence of haze is commonly
used to monitor the distribution of atmospheric dust. Haze is caused by a suspen-
sion in the air of extremely small, dry particles invisible to the naked eye. Fig. 5-14
shows the changes in haze distribution between winter and summer over the
Atlantic Ocean. The seasonal oscillations of the dust and haze plumes is due to
large-scale shifts in the wind circulation patterns and changes in rainfall, which
scrubs the dust from the atmosphere. The drastic change in haze conditions in the
Indian Ocean is due to a shift in wind direction. During the high haze season the
winds come from the northeast desert regions and during the low haze season the
winds are from the southwest.
Atmospheric transport and deposition is analogous to river transport and deposi-
tion, where much of the suspended sediment is de ivered to the sea during periodic
floods. The flow of aerosol out of Africa into t l!e Atlantic is a relatively steady
day-to-day phenomenon that is periodically augmented by large-scale dust storms
(Prospero, 1981).
The production of aerosol dust can be influenced by long-term climatic changes.
A significant increase in aerosol transport over the tropical North Atlantic in the
early 1970’s coincided with a major drought in North Africa. During much of the
Pleistocene, arid, desert regions were much more extensive than they are today. As a
consequence eolian deposits, particularly loess, were widespread and presumably
302

Fig. 5-14. Frequency of haze over the Ocean produced by wind-bourne dust from the continents.
Arrhenius, 1963. Modified by Turekian, 1976. Copyright 1963 John Wiley and Sons, Inc.

large amounts of fine aerosol material were deposited in the oceans (Prospero,
1981).
The amount of dust transported and the distances involved can be impressive. In
March and April of 1960 a dust storm in the North Caucasus removed soil to a
depth of 1 m from an area of more than 4 X lo6 hectares. The total weight of the
dust was 960 to 1280 X lo6 tons, which is larger than the annual sediment discharge
of the Amazon River. The dust was transported thousands of kilometers (Lisitzin,
1972). The calculated flux of Saharan dust across the coast of West Africa, between
15"N and 25"N, was 150 x lo6 tons over a 90-day period (summer of 1974) of
measurements. The amount of material reaching the area of Barbados (Caribbean
Sea), 5000 km west, was about 40 X lo6 tons. Thus, approximately 100 X lo6 tons
was presumably deposited in the tropical Atlantic trade wind belt. The annual mass
of carbonate-free sediments deposited in this area is estimated to be 30 x lo6 tons
303

(Prospero, 1981). This is a large discrepancy, particularly considering the atmo-


spheric flux was for only a 90-day period. Sedimentation rates in the ocean are
calculated in terms of thousands of years; thus, the 90-day flux value may not be a
reasonable average value for the past 10,000 years.
Though, on most continents rivers transport much more mineral matter to the
ocean than do winds, Prospero (1981) calculated that for Africa the dust flux may
be the major means of mechanical denudation of the continent.
Some of the meteorologic aspects of dust transport are illustrated by the Saharan
air layer (Prospero, 1981). Intense heating of the surface of the Sahara Desert causes
the formation of a mixing layer (thermals and rolls) that extends to altitudes of up
to 5 to 7 km. When this hot, dry, dust-laden air mass emerges from the coast,
moving west, it is undercut by the cool, moist northerly coastal winds, thereby
generating a sharp inversion (temperature increases with height rather than decreas-
ing, as is normal), or “lid”, at an altitude of 1 to 1.5 km. A second inversion
(subsidence) creates an upper “lid” at 5 to 7 km.
This westward-moving dust layer is known as the Saharan air layer (SAL). Dust
is transported from within the SAL to the underlying moist layer by convective
erosion of the base of the SAL and by large particles settling through the lower
“lid”. Some of the finer physil particles are transported out of the arid zone by the
lower elevation winds and are removed by precipitation in the fringing humid zones,
particularly in the intertropical convergence zone (ITCZ) near the equator (see Fig.
5-24). The layered structure is maintained for at least as far west as Barbados, West
Indies (5000 km), where dust concentrations within the SAL are three times greater
than those in the surface layer. The travel time to Barbados is from 5 to 6 days.
Other synoptic (large-scale atmospheric motions) situations are more complex.
For example, dust from the midwestern United States can ascend to high altitudes
and be transported out over the Atlantic where a branch may descend, be incorpo-
rated into the easterly trade winds and be transported back towards the United
States. Some of the dust will remain at a high altitude and move on to Europe and
some will be deposited in the Atlantic Ocean. Analyses of dust from one of the
midwest dust storms, collected from Savannah, Georgia, showed it was composed of
90% clay size material of which 80 to 90% was illite. This particular storm
transported approximately one million metric tons of dust across the east coast of
the United States. If all the dust was deposited in the Atlantic it would amount to
one-quarter of the estimated average annual sediment rate (Windom and Chamber-
lin, 1978).
The mineral composition of aerosols for specific areas is discussed in the next
chapter; only a few general observations will be made here. Physils are the most
abundant minerals in aerosols. Quartz is the most abundant non-physil, commonly
comprising 10 to 20% of the mineral suite. Minor amounts of feldspar and
carbonate minerals are present, generally in amounts less than 10%. The quartz
content of aerosols over oceans is appreciably less than the quartz content of the
suspended load of rivers. Coastal muds, and shales, composed of river detritus
commonly contain about 30% quartz. However, the presence of quartz in pelagic
muds is considered to indicate an eolian origin.
304

Table 5-3
Average Mineral Compositions per Outbreak at Sal Island, Barbados, and Miami (after Glaccum, 1978).
Sample Mica Kao. Chl. T. Clay Qtz. Mcr. Plg. Cal.
Sal Island
Mean 53.8 6.6 4.3 64.6 19.6 2.2 5.4 8.2
Std. Dev. 4.2 1.2 0.8 3.8 2.1 0.7 1.1 2.9
95% conf. interval 2.3 0.7 0.4 2.0 1.4 0.4 0.6 1.5

Barbados
Mean 64.3 8.3 4.1 71.3 13.8 1.5 4.1 3.9
Std. dev. 2.7 1.6 0.3 2.2 1.9 0.4 0.6 0.9
95% conf. interval 1.4 0.9 0.2 1.2 1.0 0.2 0.3 0.5

Miami
Mean 62.9 6.3 4.2 13.4 14.2 1.1 4.5 6.9
Std. dev. 3.9 2.0 0.5 2.4 2.6 0.4 0.8 1.3
95% conf. interval 2.8 1.4 0.4 1.l 1.9 0.3 0.6 0.9

The non-physil minerals are generally coarser than most of the physils and tend
to be deposited earlier. Table 5-3 shows the change in mineralogy of Saharan dust
samples collected on Sal Island, Cape Verde Islands, 500 km off the coast of Africa,
Barbados, West Indies and Miami, Florida.
There is a significant decrease in quartz and an increase in illite-mica with
distance. The mass median diameter at Sal Island was 6.1 pm, while at the two
western sites it was 2.1 pm. Samples collected at three different times had very
similar mineral suites, indicating the source was homogeneous. It should, however,
be noted that other analyses from the same localities show a higher kaolinite
content. The mica/kaolinite ratio in Table 5-3 ranges from 8 to 10. Other studies
report ratios ranging from 1.3 to 2.0 with kaolinite comprising 25 to 43% of the
physil suite. Based on the composition of the deep Atlantic bottom sediments, the
latter values are perhaps more reasonable.
Table 5-4 illustrates how the quartz concentration increases as a function of size.
The Barbados material is dust samples. The two Zep samples are bottom muds from
east of the Mid Atlantic Ridge which contain more quartz in all but the less than
2 pm fraction than do the Barbados dust samples. The muds’ values are believed to
represent the early stage of dust deposition.

Table 5-4
Quartz as a Function of Particle Size (after Delany et a/.,1961).
Size Barbados Barbados Zep 23 Zep 25
(P) 1 - 2
<2 5.6 8.0 7.0 6.2
2-4 12.8 13.5 22.3 20.1
4-8 17.0 19.5 25.0 30.5
8-16 23.6 21 .o 39.2 39.6
16-40 * * 56.6 46.8
* Insufficient sample for analysis
305

Typical size distributions


lo5
1

lo4-

lo3-
h
FI

L
FI 102-
-5
(L

H
.
.
n
B
100 -

I
10-11 I I I I
10-2 lo-’ 100 10’ 102
Radius (pm)

Fig. 5-15. Two representative fitted log normal size distribution curves for atmospheric crust. The upper
(solid) curve is typical of measurements made under conditions of heavy aerosol loading; the lower
(dashed) curve is typical of measurements made under conditions of light aerosol loading, From
Patterson and Gillette, 1977, Jour. Geoph. Res. Copyright 1977 Amer. Geoph. Union.

Various reports of the size distribution of atmospheric dust, over the oceans,
indicate the generally half or more of the particles are less than 2 pm in diameter.
Chester et al. (1972) reports that in dust, collected at ship level, off the west coast
of Africa 60 to 90% is less than 2 pm and only a trace is present in the 16-32 pm
range. On Barbados only 40 to 50% of the dust is less than 2 pm (Delaney et al.,
1967). As is the situation in the ocean, wind currents’ velocities are usually sufficient
to prevent gravitational settling of less than 2 pm particles. Fine particles are
cleaned from the atmosphere primarily by precipitation.
Size distribution plots indicate that many tropospheric aerosols, with a relatively
high content of soil-derived material, are trimodal (Fig. 5-15). The finest mode, 0.02
to 0.5 pm, is not related to the soil but is background aerosol commonly composed
of secondary aerosols generated by industrial processes. The mode between 1 and 10
pm is primarily a measure of the amount of physils present and the coarse mode, 10
to 100 pm, consists primarily of quartz, commonly with small physil particles on the
surface.
Windom (1969) analyzed dust samples collected from permanent snowfield and
glaciers to evaluate the contribution of atmospherically transported solids to recent
marine sediments. Fig. 5-16 shows the mineral size frequency distribution of typical
samples from Mount Olympus, Washington. The finer mode is presumably due to
material derived from global sources (distant) and the coarse mode to local fallout.
Most of the global dust is c 10 pm with an average value near 4 pm. A relatively
!rrC,,
306

Quartz
c
c
U

c
a 20
0
FeldsDar
c

8
d
Mica
+
c
U

f
a

.v

8
x
q;\,
20

‘1 4 10
~,
Chlorite

40 100
Size (microns)
I 4 10
A,
40 100
Size (microns)
Fig. 5-16. Mineral size frequency distributions in samples from the surface and at a depth of 2 m in the
Mount Olympus Snow Dome, Washington. From Windom, 1969, Geol. SOC.Amer. Bull., 80, 762-782.

small amount of the material is < 2 pm. Table 5-5 summarizes the mineral
composition of the global dust from the Northern and Southern Hemispheres and
the winds that transport the dust. Mica is the dominant physil at all locations. The
average concentration is 66% (of the physil suite). This is similar to the average
concentration of illite-mica in shales and clays.
The accumulation rates (Table 5-5) suggest that in the North and South Pacific
and Central Atlantic 25 to 75% of the detrital material in deep marine sediments
could be from atmospheric dust fallout.
One example will be given to illustrate that the composition of the physil suites in
dusts is variable and the physils reflect the composition of the source soils. The
composition of the soils is, in general, related to climate. Fig. 5-17 shows the
variation in physil composition of a series of dust samples collected, from ship-
board, along the west coast of Africa. The most significant trend is the increase in
kaolinite and decrease in illite in the humid equatorial region where chemical
weathering is at a maximum. Chlorite increases to the south where weathering is less
intense. The distribution of kaolinite, illite and chlorite is similar to that in the
underlying marine muds; however, this does not prove that the marine physils are
primarily eolian in origin. The African rivers presumably carry much the same
physil suites as the winds. However, the distribution of quartz and dolomite and the
fallout data, near the equator, suggests much of the physil material in the marine
sediments had an eolian origin.
307

Table 5-5
Global dust components, after Windom (1969).
Wind System Sample Location Accumulation Weight Percent
Rate Quartz Feld- Amphi- Mica Chlo- Kaoli- Mont-
(m/ spar bole rite nite moril-
10' yr.) *
lonite
N. Hemisphere Site 2, Greenland 0.14 23 16 5 37(66) 11 8 -
Polar Easterlies Camp Century, 24 23 6 27(57)11 9 -
Greeland
N. Hemisphere St. Elias Range, 0.1 1 21 4 3 57(81) 10 4 -
Westerlies Yukon Ter.
Mount Olympus, 0.21 23 7 2 56(82) 10 2 -
Snow Dome, Wash.
N. Hemisphere Mount Orizaba, Mex. 0.09 56 ? 11 16(52) 3 6 6
Trades Mount Popocatepetl, < 0.01 53 ? 8 18(47) - 11 9
Mexico
S. Hemisphere Tasman Glacier, 0.80 29 19 - 41(79) 7 4 -
S. Hemisphere New Byrd Station < 0.01 Present ? - - - _ -
* The accumulation rates were calculated for the accumulation of a deposit having a density of 1.5 g/cc
which is taken as an average value for the density of marine sediments. ( ) % of physil suite.

%I

:I*,;,-cw
10
0
20'" 10"N 0 10°S 20"s 3C

Fig. 5-17. The variation of clay minerals with latitude in dusts off the west coast of Africa. The
distribution of clay minerals (as a percentage of the total < 2 pm clays) is plotted against latitude; all
values are averaged for each 5' of latitude. From Chester et al., 1972. Copyright 1972 Elsevier.
308

Another minor but annoying source of atmospheric dust is crop dusting. Talc is
widely used as a carrier and diluent for pesticides (Windom et a/., 1967). When
sprayed, much of it never reaches the ground and, presumably, a portion of that
which does can be wafted back into the atmosphere. Talc, as a minor component, is
commonly present in dust, glacier, and suspended ocean samples. Much of it is
presumed to be derived from manufactured agricultural dusts, but this may not be
true (see pages 324).

Volcanogenic Sedimentation

The reason for including volcanogenic material in this discussion is that much of
it alters to physils, primarily smectites, when deposited in the marine environment.
Both subaerial and submarine eruptions are relatively common and have indirectly
been responsible for large volumes of physils.
The total number of volcanoes that have been active since 1500 A D has been
estimated to be about 450. They have produced 330 km3 (equivalent to 0.75
km3/year) of pyroclastic material and only 50 km3 of lava (Butze, 1955, in Lisitzin,
1972). The production of pyroclastic material is enough to cover the deep sea floor
at the rate of 2 pm/year or 2 m/million years. On another scale the amount of
pyroclastic output per year, 1500 X 106 tons, is approximately 7 times the annual
suspended load of the Mississippi River, or approximately 10% of the total sus-
pended load delivered by all rivers. Estimates of the yearly injection rate vary
widely. Based on the sedimentation rate in the southern Pacific Ocean, Griffin et al.
(1968) calculated that only 150 x lo6 tons of volcanic dust was mobilized to the
atmosphere annually.
Active subaerial volcanoes are very unevenly distributed. Approximately 64% are
situated in the Pacific Ocean, mainly in the western part. Thirteen percent are in the
Atlantic Ocean. Thus, volcanic material is most abundant in the western Pacific and
adjacent to island arcs. The number of active underwater volcanoes is poorly
known. More than 10,000 volcanoes occur in the Pacific Ocean, but most are old
and inactive (Lisitzin, 1972).
Pyroclastic material ejected into the air is dispersed in the same manner as eolian
material. Though the maximum recorded height of ejected pyroclastic material is 50
km, only rarely does it reach heights greater than 10 to 15 km. Tropospheric fallout
is characterized primarily by 1 to 50 pm ash particles. The finer particles can reside
in the atmosphere for as long as a month. When fine material is ejected into the
stratosphere it can be transported for years and be distributed globally.
Local ashfalls are those that are deposited within a few hundred kilometers of the
crater; tropospheric ashfalls can be deposited a few thousand kilometers; global
ashfalls can be distributed over much of the earth.
The amount of fine material in volcanic ash varies with the type of eruption and
distance from the volcano. The grain size and amount of fallout ash decreases with
distance from the volcano. There may also be some differential settling, based on
mineralogy and chemistry. Local ashfalls contain an appreciable amount of rock
309

fragments, heavy minerals and pumice, whereas tropospheric and stratospheric


fallout is predominantly volcanic glass (Eaton, 1964; Lisitzin, 1972).
Eaton (1964) estimated that eruptions which ejected ash to heights of 5 to 12 km
occur once every two years. In terms of geologic periods this can amount to a lot of
ash. Individual eruptions can produce as much as 50 km3 of ash, though values of
0.2 to 5 km3 are more typical. The ash from the larger eruptions is commonly
distributed over areas larger than 200,000 km2. The Ordovician ash beds (K-be-
ntonites) of the eastern United States are present in an area in excess of 5,000,000
km2.
Ash distribution is largely controlled by the wind. The fallout area is ellipsoidal
with the long axis corresponding to the wind direction. As a consequence of wind
transport most extensive ash falls are distributed meridianally. Whereas most local
and tropospheric ash settles due to gravity, the fine (0.2 to 1 pm) stratospheric ash is
largely washed down by snow, rain, and fog.
Submarine volcanoes on the ocean periphery commonly are explosive. The hot
pyroclastic glass rapidly hydrates to form palagonite (amorphous, hydrated glass).
Mid-ocean volcanism is more commonly in the form of lava flows. Lisitzin (1972)
believes that not only is submarine ash rapidly altered to palagonite but that the
subaerial ash that falls into the ocean alters relatively quickly to palagonite. The
palagonite, in turn, alters to smectite or zeolite (phillipsite). The widespread distri-
bution of these minerals in parts of the Pacific Ocean suggests they formed from
subaerial ash rather than more localized submarine ash.

ICE TRANSPORT

The sediment load of glaciers and icebergs deposited in the sea has been
estimated to be 20 X l O I 4 g/year (Garrels and Mackenzie, 1971) or about 8% of the
total sediment flux. Lisitzin (1972) suggests the value may be an order of magnitude
higher. The distribution of ice-rafted debris is related to the distribution of ice and
icebergs. Though glacial gravels are abundant, Lisitzin (1972) describes iceberg
sediments off the coast of Antarctica that contain 30 to 50% < 1 pm material. The
physically weathered polar physils are characterized by a relative abundance of
chlorite, a physil that is extremely sensitive to chemical weathering.

DISTRIBUTION OF PHYSILS IN THE OCEANS

General

I will first review the global distribution of physils in the oceans and then discuss
the individual oceans in more detail in order to obtain an understanding of the
various depositional processes. As most of the oceanic physils are detrital, the
distribution of the various types should be related to the climatic zones (weathering).
310

PERCENT
>50
40-50

Fig. 5-18. Kaolinite concentrations in the < 2 pm size fraction of sediments in the World Ocean. Data
from Griffin et al., 1968 and Goldberg and Griffin, 1970. From Windom, 1976, Chem. Oceanog. 5.
Copyright 1976 Academic Press.

Local anomalies are caused by alteration of submarine volcanics to smectite.


Actually, the anomalies can be quite extensive in island arc regions. The distribution
of physils in the world ocean has been reviewed by Biscaye (1965), Griffin ef al.
(1968), Rateev et al. (1969), and Windom (1976).
Kaolinite (Fig. 5-18) and chlorite (Fig. 5-19) best illustrate the effects of climate
zonality. The kaolinite content of the world ocean sediments ( < 2 pm) ranges from
< 5% to 60%. Highest values generally occur in the low latitudes, 30"N to 40"S,
where chemical weathering is most intense (tropical humid climate). On a global
scale kaolinite systematically decreases towards the poles and from near-shore to
deep ocean. The pattern clearly indicates a continental source and a latitudinal,
climatic control.
Gibbsite, another mineral formed by intense chemical weathering, has a distribu-
tion pattern similar to kaolinite and is localized in the tropical and subtropical
latitudinal belts.
In contrast to kaolinite, chlorite (Fig. 5-19) is most abundant in the polar regions
and decreases towards the equator. Chlorite is, perhaps, the most easily weathered
physil and is a significant detrital component only where chemical weathering is
severely restricted. The cold climate and marked topographic relief (rapid erosion)
of the high latitude land areas provides that relatively unweathered physils will be
transported to the oceans. The main source rocks are metamorphics and shales.
Much of the physil material is scoured and mechanically ground by glaciers (rock
flour) and transported to the ocean by ice rafting. The importance of ice rafting is
indicated by the relative abundance of silt-sized chlorite. The sharpness of the
311

PERCENT

Fig. 5-19. Chlorite concentrations in the < 2 pm size fraction of sediments in the world Ocean. Data
from Griffin et al., 1968 and Goldberg and griffin, 1970. From Windom, 1976, Chem. Oceanog. 5.
Copyright 1976 Academic Press.

chlorite x-ray peaks decreases (development of expandable layers?) towards the


equator. This reflects the increase in intensity of chemical weathering.
Concentrations of smectite-I/S range from < 20% to > 70% (Fig. 5-20). Zonal
patterns are less well developed than for kaolinite and chlorite; however, the highest
concentrations mostly occur in the southern hemisphere (below 20°N latitude), with
the highest concentrations occumng in the South Pacific. Some of the smectite is
detrital, but it appears that most of it has altered from submarine volcanics or

PERCENT
270

Fig. 5-20. Montmorillonite concentrations in the < 2 pm size fraction of sediments in the World Ocean.
Data from Griffin et al., 1968 and Goldberg and Griffin, 1970. From Windom, 1976, Chem. Oceanog. 5.
Copyright 1976 Academic Press.
312

PERCENT
250

Fig. 5-21. Illite concentration in the < 2 pm size fraction of sediments in the world Ocean. data from
Griffin et al., 1968 and Goldberg and Griffin, 1970. From Windom, 1976, Chem. Oceanog. 5. Copyright
1976 Academic Press.

atmospheric transported volcanic ash. High concentrations commonly occur in areas


where there is minor terrigenous input.
Much of the smectite is actually an I/S with a significant number of illite layers.
Various authors have referred to the variation in “crystallinity” of the marine
smectites (Biscaye, 1965). The published x-ray patterns indicate that the variations
in the character of the major “smectite” peak is likely due to the presence of
interstratified illite layers, in amounts ranging from 30% to 60%. It is not clear
whether the I/S is detrital, smectite has “fixed” K from the sea water, or some K
has been stripped from detrital illite in the marine environment. T h s will be
discussed in more detail in the following section.
The concentrations of illite (actually illite-mica) range from < 20% to approxi-
mately 80% (Fig. 5-21). The distribution (Weaver, 1959) and the high K-Ar apparent
ages (Hurley et al., 1963; Krylov and Silin, 1963) leave little doubt that most of the
illite is detrital. In a general way the illite distribution is the opposite of that of
montmorillonite. It is most abundant in the northern latitudes (temperate and cold
zones), where the continental land masses are concentrated and are a major source
of detrital sediments. Along with chlorite, it is abundant in the Antarctic region.
Illite, as well as chlorite, is derived largely from loess, till, shales and metamorphic
rocks and soils developed on these materials. Though most of the illite is trans-
ported by large rivers, much of the illite in the Pacific muds was wind-transported.
Table 5-6 lists the average physil concentrations in the major ocean basins.
Though these are not weighted averages, a comparison with the distribution maps
suggests the average values for the world oceans are reasonable. Illite and smectite
are present in approximately equal amounts and, together, comprise 70% of the
physil suite. The obviously detrital physils - illite, chlorite and kaolinite - comprise
313

Table 5-6
Average Composition of Physils in the Major Ocean Basins (after Windom, 1976).
Area Kaolinite Chlorite Smectite Illite
North Atlantic 20 10 16 56
Gulf of Mexico 12 18 45 25
Caribbean 24 11 27 36
South Atlantic 17 11 26 47
North Pacific 8 18 35 40
South Pacific 8 13 53 26
Indian Ocean 17 12 41 33
Bay of Bengal 13 15 41 30
Arabian Sea 9 18 28 46
Total average 14.2 14.0 34.7 37.7

63% of the physil suite. Both detrital and authigenic smectites are present. If it is
assumed that none of the smectite in the Gulf of Mexico is authigenic, all that in the
South Pacific is authigenic, and half that in the other ocean basins is authigenic,
then the average amount of authigenic smectite (and I/S) in the world ocean is
approximately 18%. It should be noted that as much of the “smectite” is actually
I/S, which has a lower x-ray diffraction intensity than smectite, the amount of
“smectite” is probably greater than calculated by the various authors.
Other physils present in significant amounts are I/S, palygorskite and sepiolite.
The distribution of these physils is described for the individual oceans.
Fig. 5-22 shows the concentrations of quartz on a carbonate free basis. Most of
the quartz is detrital and was transported by river, wind, and glacier. The distribu-
tion is similar to that of illite, which is also largely detrital. The average quartz

PERCENT

Fig. 5-22. Quartz concentrations in sediments of the World Ocean on a carbonate free basis. Data taken
from Goldberg and Griffin, 1964, Beltagy and Chester (in Riley and Chester, 1971, p. 333), Ellis, 1972,
Griffin and Goldberg, 1969 and Windom (unpublished data). From Windom, 1976, Chem. Oceanog. 5.
Copyright 1976 Academic Press.
314

concentration is approxinmately 9%, considerably less than for shales (30%). This
reflects the paucity of deep ocean muds-shales in the geologic record.

Atlantic Ocean

The kaolinite/chlorite ratio (Fig. 5-23) of the physil suite ( < 2 pm) in the bottom
sediments of the Atlantic Ocean is probably the best example of symmetrical
latitudinal zonation. The ratio is in excess of 10 near the equator and decreases
towards both poles. Where ratio values are > 10 the kaolinite content is in the range
of 50 to 60%. The distribution of gibbsite is similar to that of kaolinite.
Off the eastern coast of South America the area of abundant kaolinite is
relatively small. This is because only small rivers, particularly the Sao Francisco,
drain the area containing heavily weathered tropical soils. The bulk of sediment
from South America is transported by the Amazon River which derives most of its
detritus from the cool highlands of the Andes Mountains (Gibbs, 1967) (see
p. 195-197). The situation is analogous to that in the southeastern United States
where the small coastal plain rivers transport a physil suite (montmorillonite)

Fig. 5-23. Distribution of kaolinite/chlorite ratio in < 2 p non-carbonate fraction of Atlantic bottom
sediments. The distribution corresponds to weathering intensity differences due to latitude o n the are
continents. Generalized from Biscaye, 1965. From Turekian, 1976. Copyright 1967 Prentice-Hall.
315

strikingly different from that of the larger rivers (kaolinite) (see p. 202). The physils
carried by small rivers reflect the composition of the rocks and the climate on the
landmass adjacent to the immediate offshore area, whereas larger rivers commonly
transport material from a more diverse terrain and wider climatic range and reflect
regional conditions. Thus, the size of the rivers must be taken into account when
interpreting the significance of detrital physils in ancient sediments.

Kaolinite
The high kaolinite concentration off the west coast of Africa (Fig. 5-18) reflects
the presence of major rivers, the Niger and Congo, draining the African tropical
region and the presence of trade winds. As the coastal equatorial counter current in
the area of the Niger and the Congo deltas flows to the south, it is unlikely the rivers
are a major source of the kaolinite in the area north of the Niger Delta. On the basis
of the location of the kaolinite “high”, the systematic westward decrease of
kaolinite, and analysis of dust in the northeast Trades, most authors suggest much
of the kaolinite is eolian, derived from the Sahara Desert. The pattern of frequency
of haze off the coast of northwest Africa (Fig. 5-14) is very similar to that of the
kaolinite distribution.
Though it is generally considered that much of the fine-grained material, particu-
larly kaolinite, off northwest Africa is eolian, the importance of atmospheric
transport may be overestimated, particularly in the “nearshore” area. Saharan dust
off the coast of Africa contains 25 to 35% kaolinite (reviewed in Windom, 1975), yet
the marine sediments contain up to 30 to 60%. Second, the kaolinite “high” is not
off the coast of the Sahara Desert, but occurs immediately south of the southern
boundary of the desert (Fig. 5-24). The climate of the coastal area is humid tropical
and the soils are classed as latosolic soils. This area is drained by many small rivers.
It seems likely that the kaolinite “high” is largely due to river transport from the
tropical coastal area rather than eolian deposition. Further, atmospheric scientists
believe that quartz in fine-grained marine sediments is a measure of eolian input.
The quartz “high” ( > 15% in the < 2 p m fraction) in equatorial North Africa (Fig.
5-24) lies north of the kaolinite high and is directly offshore from the Sahara Desert
(Windom, 1975). More recently Sarnthein and Koopman (1980) suggested that the
silt-sized quartz (10-12%) was deposited from the low level trade wind but that the
kaolinite and other lateritic weathering products were transported from the semi-arid
Sahel zone (on the southern flank of the desert) by the 700 mb (millibar) winds. The
distribution of silt-quartz indicates that during glacial times, at 1800 years B.P., the
arid belt was wider, both to the north and to the south. The size of the quartz and
distance of transport indicate wind velocities were higher.
Analyses of core samples off the coast‘ of the Sahara and immediately to the
south show that the physil suite remained essentially constant for the past 0.7 m.y.
+
(through the glacial cycles). The K Ch/I ratio (chlorite is negligible) of the Sahara
coast is 1 and, in the southern well, 3. The difference is similar to that found in
recent samples from the same area (Parkin and Padgham, 1975).
Fig. 5-25 shows some of the data concerning deposition in the eastern Atlantic.
The kaolinite concentration in the near-shore marine sediments closely follows the
316

40'

30'

20'

10'

0'

10'

20

30

I I I I I I I I I I I I
90' 00' 70" 60" 5O0 4O0 30° 20- 10° WOO€ 10' 20' 30-
Fig. 5-24. Kaolinite distribution in the < 2 pm size fraction of Equatorial North Atlantic sediments.
Arrow and shaded area indicate trajectory and dust veil of dust storm studied by Prosper0 et al., 1970.
Dashed line outlines high ( >15%)quartz lobe. Data used in figure preparation are from Griffin et al.,
1968. Modified from Windom, 1975. Copyright 1975 SOC. Econ. Paleo. Miner.

precipitation curve for the coastal area, though it appears to be relatively high
offshore from the Sahara. This may reflect a contribution from the 700 mb winds.
Samples of atmospheric dust, from ship level, show a symmetrical distribution of
kaolinite, centered very near the equator (Fig. 5-17); the illite distribution is the
opposite. High values occur offshore of both the Sahara and the Namib Deserts.
Illite is commonly the most abundant physil in ancient desert deposits. The
distribution of illite in the coastal marine sediments is similar to that of the
atmospheric samples. In this area the physil suite of the airborne dust appears to be
more closely related to the climate zones than is the marine physil suite. The
disparity between atmospheric and marine kaolinite values in the 0" to 15"N
interval is presumably due to the input of coastal rivers.
The kaolinite distribution pattern in the mid-Atlantic (Fig. 5-24) is similar to that
in the coastal atmosphere. This similarity, along with numerous other data, indicates
much of the deep ocean kaolinite is eolian.

Chlorite
As the kaolinite/chlorite ratio map (Fig. 5-23) and the chlorite distribution map
(Fig. 5-19) indicate, chlorite increases towards the poles. In the regions poleward of
approximately 50"s and 50"N the chlorite values are in the 20 to 30%range; locally
31 7

300t
4001
0
\
10
,
20 30 40 50
Percent physils; cm precipitation
60

Fig. 5-25. Distribution of some physils in the air and marine sediments off the west coast of Africa. Mean
annual precipitation, in cm, is also shown. Marine physil data from Biscaye, 1965. Atmospheric data
from Chester et al., 1972.

values are > 30%(Antarctic) and as high as 47% in the Arctic, north of Greenland
(Berry and Johns, 1966). These are high values. The average shale has approximately
10%chlorite or less. The high chlorite values not only reflect conditions of mild to
no chemical weathering, but source rock with a relatively high chlorite content -
low-grade metamorphic rocks (i.e., slates and greenschist).
Chlorite is not only abundant in the < 2 pm fraction but also in the silt-sized
material (Griffin el al., 1968). This is further evidence that much of it was derived
from metamorphic rocks.
The sharpness of the x-ray peaks tends to decrease with decreasing latitude,
which is presumably due to increased weathering in the temperate regions. The
relative intensity of the 001 peaks suggests the chlorites are iron-rich (Biscaye, 1965).
This technique is not reliable, and the most that can be said is that the chlorites do
not have a very high Mg content. They are apparently similar in composition to
chlorites in shales and low-grade metamorphic rocks (Fe/Mg 0.4-2.0). -
The decrease in chlorite around 50"s latitude closely coincides with the northern
boundary of ice sediment transported from Antarctica by ice (Lisitzin, 1972).
Chlorite and illite were literally scraped off the fresh outcrops and transported as
much as 2000 km north by icebergs. The high chlorite concentration in the
northwest Atlantic is due primarily to the input from the St. Lawrence River, which
contains 30 to 35% chlorite in its suspended load.
318

Vl5-130/top A180-73/10crn
58‘s - 60‘W OOD-230W
v/p = 0.18 -
v/P = 0.17
I

_$”/
lb

Fig. 5-26. Low-angle diffractograms (V14-88/top, equatorial Atlantic; V15-130/top, South Atlantic;
A180-73/10 cm, southwestern Indian Ocean) showing variations in the v/p (“crystallinity”) ratio of
< 2 pm, glycol-expanded montmorillonite. Degrees 20 (Cu K a ) are marked below the base line.
Symbols are v - valley” depth: p - peak height; M - montmorillonite; C - chlorite; ML - mixed-layer
or palygorskites; I - illite. From Biscaye, 1965, Geol. Soc. Amer. Bull., 176, 803-831. Physils probably are
mixed-layer illite/smectite with ratios of 3:7, 6:4, and 63:37 (left to right).

In the area between Greenland and Norway the chlorite content increases from
an average around 30% to 43% near the North Pole. This distribution is also
presumed to be due to ice rafting.

Expandable Physils
The expandable physil in the ocean sediments has generally been referred to as
montmorillonite with varying degrees of crystallinity (Biscaye, 1965; Griffin ei al.,
1968). Chemical analyses indicate most of the expandable material is montmoril-
lonitic and much has a relatively high Fe content, particularly in the Pacific Ocean.
Fig. 5-26 shows the types of x-ray peaks afforded by the expandable physils in the
Atlantic. Biscaye devised a “crystallinity index” (v/P) based on the discreteness of
the 001 x-ray peak. Various explanations have been given to explain these variations
in peak shape (i.e., particle size, lack of periodicity, chemical degradation, poor
orientation) (Griffin et a/., 1968). If these x-ray patterns were obtained from a shale
the expandable material would be considered to be I/S clays. The following
interpretation would be made of the expandable physils in Fig. 5-26: v/p = 0.70 =
I/S ( - 3:7), v/p = 0.18 = I/S ( - 6:4), v/p = 0.17 = I/S ( - 63:37). Actually, the
319

interpretation is probably more complex, as vermiculitic and chloritic layers are


probably present in many of the samples. Biscaye does identify a mixed-layer phase
that has peaks between 10 and 14 A. These spacings would indicate that the I/S
would have more than 70% illite layers. Some peaks in this interval do not shift
when the sample is glycolated and are apparently due to palygorskite (10.5 A) and
sepiolite (12 A).
Expandable physils can be detrital from the continents or formed by alteration of
submarine volcanic and atmospheric transported volcanic material (see “Authigenic
Marine Physils”, Chapter VI). There are a few questions to be answered: Is the I/S
material detrital? Has some of it formed from montmorillonite in the marine
environment? Has it formed directly from basaltic material in a marine environ-
ment?
The Atlantic Ocean contains less expandable material than the other oceans. The
highest concentrations are in the central South Atlantic where values are in the 30 to
50% range and higher. Because much of the “montmorillonite” is actually I/S the
true values are probably higher than the calculated values in the literature. The
origin of the high concentration of expandable physils in the South Atlantic is still
not known.
The physil suite of the Niger River contains 25 to 40% expandable physils, but
the current pattern suggests it would not be transported to the South Atlantic.
(Kaolinite is not the only physil produced in tropical regions. If drainage is poor,
abundant montmorillonite can form in the soils.) Similar reasoning can be applied
to the Congo River. The South Atlantic dusts contain less than 20% expandable
(Chester et af., 1972) and are not a likely source. Some of the expandable material
may be due to the submarine weathering of volcanic material from the Mid-Atlantic
Ridge, but why is there not an equivalent concentration in the North Atlantic?
A wide variety of silicate minerals has been found associated with basalts in the
Mid-Atlantic h d g e and Trench, including trioctahedral (Fe, Mg) smectite, cel-
adonite, talc, chlorite, corrensite, serpentine, sepiolite and palygorskite. Most of
these physils were formed under hydrothermal conditions created by the extrusion
of molten basalt and waters associated with the basalt. Studies of the Cenozoic and
Mesozoic sediments of the North Atlantic indicate that the predominant physil is an
Fe-montmorillonite, derived directly or indirectly from oceanic basalts. The low
content of smectitic material in the North Atlantic indicates that, at least during the
Quaternary, physils derived from marine basalts are relatively insignificant. In the
central South Pacific where the influx of terrigenous material is less, it is quite likely
a significant portion of the smectite physils were derived from basalts. The surface
alteration zone of pillow-lava from the center of the Rift Valley in the Mid-Atlantic
Ridge is from 6 to 350 pm thick and is composed largely of palagonite and
manganese with minor “poorly crystallized montmorillonite” and chlorite. At a
distance of 4 to 6 km from the center of the rift the basalts are older, 130,000 to
482,000 years, and have similar but thicker (130 to 1,955 pm) zones of alteration
(Hekinian and Hoffert, 1975). The rate of formation of montmorillonite from
submarine volcanoes would appear to be too slow to make a major contribution to
Holocene sediments, at least in the North Atlantic.
320

Griffin et al. (1968) suggested that volcanic ash from the South Pacific and/or
South America could have been transported into the area by the westerly winds and
altered to montmorillonite after deposition. The relatively high concentration off the
northeast coast of South America apparently reflects the input of the Amazon River
(Gibbs, 1977) and the effects of differential settling. The amount of “montmoril-
lonite” in the shelf sediments increases from 27% at the mouth of the Amazon, to
40% 1400 km to the northwest.
Between latitudes 20”N and 55”N expandable physils comprise < 20% of the
physil suite and are largely detrital. Above 55”N values increase to an average of
33% near volcanic Iceland, then decrease northward to 20% near the pole (Berry and
Johns, 1966). This suggests a volcanic influence, either subsea alteration or, more
likely, from the soils developed on the Iceland volcanic rocks. The high concentra-
tion off the northwestern coast of Antarctica may also be related to volcanic islands
in the area of the Palmer Peninsula.
The v/p values have a distribution similar to that of kaolinite. Values are highest,
0.50 = I/S with I < 40%, in the high latitude areas and decrease to 0.0 = I/S
-
with I 60% in the central Atlantic, where kaolinite is most abundant. The
distribution and the composition of the expandable material, along with other data,
presented later, indicate it is most likely that the expandable physils with relatively
few illite layers were derived from the weathering of volcanic material, either subsea
or subaerial. The material with a high proportion of illite layers probably is
subaerialy degraded illite, mica, and chlorite. Some of these layers that retained a
high layer charge could have been rejuvenated (adsorb K or Mg hydroxide) in the
marine environment (Weaver, 1958). This degraded material is more likely to form
in the tropical and temperate climate zones than in cooler climates. Berry and Johns
(1966) suggest that in the marine environment illite and chlorite can be degraded to
a mixed-layer phase. There is no direct evidence that this occurs.
In addition to a poleward increase in v/p, the v/p generally increases, but to a
lesser extent, from the central part of the Atlantic towards both the east and west.
Thus the expandable physils are becoming more illite towards the center of the
basin where the sedimentation rate is very slow. There is the possibility that some
illite layers are forming under these conditions. K-Ar values (Hurley et al., 1963)
indicate most of the illitic material is Paleozoic in age and therefore detrital;
however, the uptake of a small amount of K could be masked by the large amount
of K in the more abundant detrital illites. The K-Ar age of the expandable physils
needs to be determined.
Our studies in the southeastern United States leave little doubt that a full
spectrum of montmorillonitic-I/S physils are being supplied to the Atlantic Ocean
(Fig. 5-27). In general the Coastal Plain rivers are transporting a relatively good
“montmorillonite” (I/S with 80 to 90% smectite), obtained directly from the
Tertiary clays. The proportion of expandable layers increases southward. The larger
rivers draining the Piedmont carry mixed-layer expandable physils with a low v/p
ratio (I/S with 60 to 40%, and less, smectite). Similar material is probably
transported from the north by longshore currents. Many of these low v/p samples
do not collapse to near 10 A when heated and presumably contain chloritic, and
321

Fig. 5-27. X-ray patterns of detrital I/S from the southeastern coast of the United states; South Carolina
(left) to North Florida (right). The proportion of illite interlayers decreases from about 60%(left) to 20%
(right).

probably vermiculitic, layers. The Piedmont contains a variety of slates, schist


gneisses and granites. Where the soil profile or saprolite is well developed, kaolinite
is abundant; however, where weathering is less intense the chlorites and micas alter
to vermiculite, smectite and a variety of mixed-layer products (e.g., I/S, Ch/V,
Ch/S, I/V/S etc.). Some of these degraded materials may be partially rejuvenated
by extracting K and/or Mg (hydroxide) from the sea.
Biscaye notes that some “obvious” (x-ray peak between 10 A and 14 A) mixed-
layer material is present. This material should be primarily I/S with 70 to 90% illite
layers. The mixed-layer physils are most abundant in the North Atlantic along the
west side of the Mid-Atlantic Ridge. From north to south the main mixed-layer
peak decreases in intensity and shifts from 10.7 to 13.3 A (not e.g.). This is
equivalent to an increase in the smectite layers from approximately 20 to 40%.
Farther north (Norwegian Sea to Arctic Ocean), the amount of mixed-layer material
decreased from south to north (Berry and Johns, 1966). Biscaye also found the
mixed-layer material was present in the 2-20 pm fraction. The size and areal
distribution suggest the mixed-layer physil is detrital and reflects weathering condi-
tions on the North American continent. In the Arctic regions there is little chemical
weathering and little mixed-layer material is formed. Southward, the climate be-
comes more temperate; K is stripped from the illites and micas, creating an I/S
with a high proportion of illite layers. Farther south, where weathering is more
intense, the micaceous material is degraded to the extent that it only contains a
small percentage of 10 A layers and is classed as montmorillonite rather than as
mixed-layer material.

Illite
Illite (Fig. 5-21), in both the < 2 pm and 2-20 pm fractions, is the most
abundant physil in both the North and South Atlantic and is most abundant in the
322

former. Windom's map of the northern portion of the North Atlantic is generalized
and only indicates illite is present in an amount > 50%. Rateev et a/. (1969) report
values ranging from 60 to 70% for most of the area, with slightly lower values, 40 to
60%, in a tongue offshore from the St. Lawrence River. Griffin et ul. (1968) report
general values of 60 to 70%, with the offshore St. Lawrence River area containing 50
to 60% illite. They also show a large area off the coast of Spain that contains > 70%
illite.
The high illite content generally reflects the moderate climate and the abundance
of Paleozoic and metamorphic rocks in much of the coastal area fringing the North
Atlantic. Most of these rocks are in mountain ranges and probably supply a
disproportionate share of physils to the ocean.
The suspended physil suite of the St. Lawrence River contains an average of 64%
illite (D'Anglejam and Smite, 1973), a value similar to that in the marine lobe. What
is surprising is that the lobe maintains its identity for a distance of approximately
2000 km. Apparently much of the suspended material from the St. Lawrence is
transported by turbidity currents along the Laurentian Cone and by the eastward-
flowing Gulf Stream as far east as the Mid-Atlantic Ridge. Griffin et al. (1968)
suggest the high concentration of illite off the coast of Spain is due to eolian
transport from the North American continent, and perhaps Asia, by the jet-stream
westerlies. River input in this area is minor.
From 30"N latitude southward, along the east side of the Atlantic, the distribu-
tion of illite closely follows the climatic zones. Illite concentrations are relatively
high offshore from the Sahara and Namib Deserts and decrease towards the equator
(Fig. 5-25). The physil suite on the shelf off the west coast of south Africa contains
65 to 75% illite and 25 to 37% kaolinite (in addition to glauconite). The adjacent
rivers have a similar physil suite (Birch, 1973). The distribution is similar to that in
dust samples from the same area (Fig. 5-17). In the tropical region much of the illite
is converted to kaolinite and/or montmorillonite. The Niger River muds contain on
the order of 20 to 30% illite, whereas the Saharan dust contains on the order of 50%.
The physil suites in Arctic fjords commonly contain 80 to 90% illite-mica.
The numerous lobes (50 to 60% illite) extending off the coast of South America
suggest the high illite concentrations are due to detrital input from the various
major rivers; however, there is a problem. For example, sediment off the northeast-
ern coast of Brazil is supplied by the Amazon River. Gibbs (1977) reported that the
muds at the mouth of the river contain 30% illite ( < 2 pm) and the amount in
bottom sediments systematically decreases to 20%, 1400 km north, in the area of
Trinidad. A study by Eisma and van Der Mare1 (1971) in the same general area
reports illite values in the 20 to 25% range. Actually, Biscaye reports illite in the
range of 25 to 49% in this area. These values seem reasonable. The high values off
the northern coast of Venezeula appear to be real but the source is open to question.
Jacobs and Ewing (1969) found that the surface physil suite (total sample) at the
mouth of the Amazon contained a high ( > 50%) content of illite-mica. Millimar et
ul. (1975) found that the I/K ratio of suspended physils in the mouth of the
Amazon River (2.3) was considerably higher than that in the underlying mud (1.04).
They suggested this material was transported to the Caribbean Sea, apparently
323

bypassing the Brazilian-Guiana coast, where some was deposited. It is also likely
that considerable illite was derived from the mountains of northern Venezuela. Illite
from the Sahara dust presumably makes some contribution, but the dust samples
collected at Barbados only contain an average of 46% illite, so, it is unlikely this
material would account for a major portion of the samples containing > 50% illite.
The samples from the middle of the Atlantic Ocean, where illite values are in the 40
to 50% range, are more likely to reflect appreciable eolian input. Actually, the belt
of sediments lying adjacent to latitude 20"N between northwest Africa and Barba-
dos has a physil suite very similar to the Barbados dust: illite, 46%; montmoril-
lonite, 18%; kaolinite, 36%; chlorite, 11%(recalculated to exclude quartz).
The high illite lobe extending off the coast from Buenos Aires is presumably due
to the high illite-mica input from the La Plata River, which, near its mouth, carries a
physil suite containing 44% illite in the clay fraction and even more in the silt
fraction (Depetris and Griffin, 1968), The northeastern orientation of the illite lobe
is probably due to the convergence of the northward-flowing Falkland Current and
the southward-flowing Brazil Current, producing an eastward-flowing current.

Other Physils
Biscaye (1965) found traces of pyrophyllite in samples from off the northern
coast of Africa and the northeastern coast of South America. It was presumably
derived from metamorphic rock. Pyrophyllite-bearing rocks are present on both
northwest Africa and northeast South America.
Talc is commonly a minor to major component of the suspended physil suite in
many areas of the Atlantic, the Caribbean and the Mediterranean (see Poppe et a/.,
1983). It is relatively rare in bottom sediments. Though talc is present in the
Amazon River (Gibbs, 1967), in northern South America, and in the Appalachian
Mountains, among other places, most of that in the oceans is thought to be
anthropogenic, but this may not be true.
Talc is widely used in many commercial products including paint, ceramics,
rubber, roofing, paper, toiletries and pesticides and finds its way to the world
oceans primarily in industrial wastewaters and pesticide dusts (crop dusting).
Windom (1969) found talc in every snowfield he sampled in the Northern Hemi-
sphere. Windom et al. (1967) found talc in a variety of snow, rain and atmospheric
dust samples. Risebrough et a/. (1968) established the presence of dust samples
containing talc. Thus, there is little doubt that an appreciable amount of the talc in
the ocean waters was transported by atmospheric winds after being reduced to a
fine size by industrial processes. Analyses of suspended samples from three depths
on the Continental Shelf and Upper Slope, northwestern Atlantic, indicated the
percentage of water samples containing talc decreased with depth (Poppe et al.,
1983). They concluded its absence (not detected) in bottom samples was because it
had been introduced to the atmosphere so recently that there had not been time for
a detectable amount to accumulate in bottom sediments.
Talc is relatively abundant in the swash zone of beach sands of the northeastern
Gulf of Mexico and along the Atlantic coast of Florida (Griffin, 1963).
324

Suspended Physils
One example will be given that demonstrates the complexities and uncertainties
involved in studying the transport and deposition of physils in the open ocean. In
the offshore areas between Virginia and northern Georgia four water masses can be
identified on the basis of their hydrographic properties. These are the Virginia
coastal water (north of Cape Hatteras), Carolina coastal water (North Carolina shelf
south of Cape Hatteras), Gulf Stream water (outer edge of shelf and surface water
layer west of the shelf break), and Carolina slope water (underlying the Gulf
Stream). Analyses of suspended material from these four water masses indicate they
carry distinctive physil suites (Table 5-7) (Pierce et af., 1972). X-ray patterns are
shown in Fig. 5-28.
The predominant physils in the Virginian coastal water are well crystallized illite
and chlorite along with minor kaolinite and talc. T h s material has apparently been
transported from a considerable distance to the north, possibly from glacial deposits
on the Continental Shelf. The Carolina coastal water contain “montmorillonite”,
kaolinite, illite and no chlorite. This suite is similar to that in the Carolina estuaries
and much of it was locally derived. The Gulf Stream, characterized by high
temperatures and high salinities, suspended suite is characterized by an abundance
of talc and minor amounts of chlorite and kaolinite. The authors did not speculate
on the source of the talc. If the talc is not due to contamination, then it was most
likely deposited from the atmosphere and/or transported a long distance. The
suspended physils in the southern equatorial water off the northeastern coast of
Brazil have a talc/mica ratio of 6 (Milliman el al., 1975).) It is difficult to imagine a
dust, 100 km from land, composed almost entirely of talc (the minor amount of
chlorite is probably associated with the talc flakes). However, Risebrough et a/.
(1968) show an x-ray pattern of dust collected over the Coral Sea, Pacific Ocean, in
which talc is by far the major physil. It is possible that much of the talc in the ocean
waters and in atmospheric dust crystallized in the upper, evaporatic film of ocean
water. The Carolinian slope water, underlying the Gulf Stream, is believed to be
composed primarily of Caribbean water. The physil suite is similar to that in the
Gulf Stream; talc is dominant. However, the suite is not like that in the Caribbean
and Gulf of Mexico, where mica dominates the suspend physil suite and talc is a
minor component (Jacobs and Ewing, 1969).

Table 5-7
Summary of Mineralogy of Different Water Masses (after Pierce er ol., 1972).
Water Mass Mineralogy
Virginian coastal Chlorite, illi te, talc, amphibole, quartz, feldspar, dolomite,
rare traces of kaolinite
Carolinian coastal Montmorillonite, illite, kaolinite. quartz, calcite
Gulf stream Minor chlorite, talc, kaolinite, quartz
Carolinian slope Illite, talc. kaolinite. quartz, feldspar, calcite,
dolomite
325

I 3.5 4 4.5 5 6 7 10 14 A

Fig. 5-28. X-ray diffractograms of samples from different water masses. (a) top, Virginian coastal water;
(b) Carolinian coastal water, (c) Gulf Stream; and (d) Bottom, effluent from Savannah, Georgia harbor
showing close relationship to carolinian coastal water but with addition of talc to suite. From Pierce et
al., 1972. Copyright 1972 Dowden, Hutchinson and Ross.

Nevertheless, there are at least three distinctive physil suites in this area. Two
coastal suites, one containing physils transported a considerable distance along
shore (Virginia), a second containing a locally derived suite (Carolina), and a third
east of shelf edge that contains global and/or marine physils. The above discussion
demonstrates the problems involved in trying to relate marine suspended physil
suites to source.
A study of a series (seven traverses) of closely spaced suspended samples
extending 30 km off the coast of Georgia illustrates further the inhomogenity of
suspended marine physil suites (Bigham, 1972). The suspended physil suite ( < 2 pm)
changes composition seaward. Kaolinite is relatively uniformly distributed ( 30%). -
Illite values increase from -
30% in the inner 8 to 13 km to 70% seaward;-
conversely, montmorillonite values decrease from -
30 to 40% in the inner shelf
area to zero, 8 to 15 km offshore. The trend is the opposite of that noted off the
mouth of the Amazon. Chlorite-vermiculite is present in only a few samples. Minor
talc and pyrophylLite are present in about half the samples.
The distribution indicates the montmorillonite, locally derived, is not transported
very far seaward and most is probably flocculated and trapped in the estuaries and
marshes, whereas the kaolinite resists early flocculation. Much of the kaolinite may
have been introduced by the Savannah and/or Altamaha Rivers in fresh water
plumes extending out into the shelf. The illite has a distant source.
Thus in addition to the four water suites described by Pierce et al. (1972), there is
a fifth suite in the coastal waters of offshore Georgia. The innermost physil suite
along the Georgia coast is similar to the Carolina coastal water suite (montmoril-
326

lonite, kaolinite and illite). The physil suite of the midshelf is composed predomi-
nantly of illite ( - 60%) and kaolinite and chlorite ( - 40%). The physil suite and
x-ray patterns are similar to those of the suspended suite in the Caribbean and Gulf
of Mexico (Jacobs and Ewing, 1969). The North Equatorial Current moves parallel
to the equator in the Northern Hemisphere where it is joined by that portion of the
South Equatorial Current that is shunted towards the South American coast. The
flow splits into two masses; the Caribbean Current that passes through the Yucatan
Channel into the Gulf of Mexico and the Antilles Current that flows along the
Atlantic side of the West Indies. These currents reconverge as the Caribbean
Current exits the Gulf of Mexico between Florida and Cuba and form the Florida
Current, moving north.
Apparently, on the basis of the physils, these two currents do not mix initially.
The Caribbean Current, characterized by a predominance of illite-mica, stays to the
west and moves over the continental shelf. The Antilles Current, talc-rich, stays to
the east, along the shelf edge. Thus, the talc-rich water mass moves from the
equatorial region where conditions would be favorable for it to crystallize in the
zone of surface evaporation. The Caribbean portion of the current moves along the
South American coast, where the talc is diluted by the influx of illite-mica.
It is important to note that not only does the amount of suspended material
decrease seaward but the proportion of mineral grains also decreases. So, even
though a global current may carry a unique physil suite into a coastal area, it may
have only a slight influence on the bottom physil suite.
Bottom samples (offshore Georgia) were collected at each site where a suspended
sample was collected. There is considerable difference between the two physil suites.
Table 5-8 summarizes the data for the suspended samples and the bottom sediment
samples in the inner (8 to 13 km) and outer (8-13 to 30 km) shelf.
The bottom sands of the inner shelf have slightly more montmorillonite and less
illite than the mid-shelf sands. This is similar to the distribution of suspended
physils. The differences between the bottom and suspended physil suites are greater
in the mid-shelf area. Apparently much of the physil material in the Gulf Current
bypasses the shelf. T h s is not surprising. The entire shelf is covered with sand and
appreciable ( < 10%) clay and silt size material occur only in the inner shelf area.
The clay and silt are believed to be Holocene sediments, which the physils tend to

Table 5-8
Composition of Suspended and Bottom Physil Suites from Offshore Georgia (data from Bigharn. 1972).
Average Percent Physils
lllite Kaolinite Montmorillonite Chlorite
Inner Shelf
Suspended 32 28 34 7
Bottom 17 36 45 3
Mid-Shelf
Suspended 59 21 0 15
Bottom 27 32 31 5
321

confirm. The few mid-shelf bottom physils were apparently deposited during low
water stages of the Pleistocene, as was the sand, and reflect a local origin.
Farther south, the inner shelf carbonate muds along the east coast of the Florida
Keys contain a complex physil suite (Manley, 1973). The following physils were
identified: montmorillonite, chlorite, sepiolite, I/S (possibly palygorskite?) and
traces of illite, talc, kaolinite and corrensite. Sepiolite is enriched to the north and
illite and chlorite to the south. Insoluble material comprises 50 to 70% of the
carbonate muds in a shoal area extending approximately 1.6 km seaward. The
source is not known but all of the physils, with the exception of talc and corrensite,
are present in the Cenezoic carbonate rocks of Florida.
Hathaway’s (1972) study of bottom physils, primarily from the Continental slope,
shows a systematic increase in illite from the Florida slope (absent) to New York
(50 to 60%) (Fig. 4-29). The trend on the shelf is similar, though illite is more
abundant on the southern portion of the shelf. The illite in the physil suite increases
from 20% off the coast of southern Georgia, to 25% off the coast of northern South
Carolina (Weaver), to 50 to 80% off the coast of Maryland and New Jersey (Miller
et al., 1978), and remains high on up to the New England shelf (Hathaway, 1972).
Pevear (1972) reported a similar trend, zero illite in Florida to 21-50% in northern
North Carolina, in coastal and nearshore sediments. Chlorite increases from ap-
proximately 5% in the south to 15 to 45% (Maryland and New Jersey), to 40% (New
England). If the illite in the south was transported from the north, more chlorite
should be present. It is likely much of the illite in the S.E. shelf was transported by
the northward-flowing Florida Current.
Hydrothermal palygorskite and sepiolite are locally abundant. Their distribution
and origin are discussed in the following chapter.
It is apparent from this review that we know very little about the transport and
deposition of physil suites in the open marine environment. Much more data are
needed on the identification and distribution through time of suspended physil
suites.
The distribution and origin of Atlantic Ocean physils since the development of
the Proto-Atlantic Ocean is discussed in Chapter IX.

Pacific Ocean

North Pacific
The Pacific Ocean can be treated as two separate oceans - the North Pacific, in
which illite is the dominant physil, and the South Pacific, where montmorillonite is
dominant (Fig. 5-20 and 5-21).
One of the most striking features in the North Pacific is the band of high illite
concentration (60 to 80%) (Rateev et a/., 1969). This band also contains a relatively
high content (10 to 15%) of eolian quartz (Rex and Goldberg, 1958). Rateev et a / .
(1969) believe the illite in this band came from the Hwang Ho and Yangtze Rivers,
which drain large areas of illite-rich loess. Griffin et a / . (1968) believe the illite and
328

quartz concentration is due to jet stream transport from the European-Asian arid
land areas.
The illite (2M) content of the bottom shelf sediment physil suite of the East
China Sea averages 65% (25% chlorite, 6% kaolinite, 4% I/S) (Chen, 1978). Samples
10" longitude seaward of the shelf break contain 60 to 70% illite (Aoki and Oinuma,
1974). Illite decreases in abundance but remains the dominant physil as far south as
Borneo. To the north, in the Sea of Japan, the illite content averages 54% (24%
chlorite, 12% kaolinite, 11% montmorillonite) (Saburo and Kaoru, 1973). In the
Philippine Sea (east of the Philippine Islands) the 50% contour occurs at about 20"
latitude (Kolla et a/., 1980). The > 50% smectite contour (increasing to the south)
ranges between 10" and 20"N latitude. Thus, the western end of the North Pacific
illite is broader than shown in Fig. 5-21 and extends west to the mainland of China.
The zone of 50 to 70% smectite can be extended westward to the southern end of
Luzon Island.
As we have seen from the discussion of the physils in the Atlantic Ocean,
illite-mica preferentially stays in suspension and can be transported a long distance
by marine currents (Amazon River to the Gulf of Mexico and farther). The current
direction and net water transport contours for the North Pacific have a distribution
that is very similar to the percent illite contour pattern (Sverdrup, 1947). I t is
plausible that much of the illite and associated physils in the illite band, at least in
the western portion, was current transported and derived from the major rivers
draining central Asia.
One of the major arguments for the importance of atmospheric transport is the
abundance and size of the quartz in the pelagic sediments (Rex and Goldberg,
1958). The quartz, in one pelagic sample, had a size range of approximately 1.0 to 30
pm, with a strong mode between 3 and 10 pm. The authors contend that quartz of
this size could not be transported to the mid-Pacific by marine currents. On the
basis of an oxygen isotope study of quartz from the Pacific sediments and from
various source areas, Clayton et al. (1972) concluded it was eolian, probably derived
from loess deposits. Windom (1969) compared the mineral composition of atmo-
spheric dust samples to the composition of bottom samples and concluded that
about 50% of the < 2 pm material and essentially all of the > 2 pm material in the
central North Pacific sediments could be attributed to atmospheric transport, giving
a total value of 75% for the eolian contribution (approximately 50% of the bottom
material is < 2 pm).
The illite concentration systematically decreases away from the area of maximum
concentration, except to the west. The decrease is largely due to an increase in the
montmorillonite content of the physil suite; chlorite is a significant component
along the northern flank of the illite-high. The relative decrease in illite to the north
and east is probably due, primarily, to dilution with water transported physils from
the continents. The southern decrease is probably related to the authigenic forma-
tion of montmorillonite from marine volcanic material.
Fig. 5-29 shows a series of x-ray patterns of bottom samples that extend from
near the west coast of the United States to near Taiwan. The near-coast samples on
both sides of the Pacific are composed predominantly of I/S ( - 3:7). The relative
329

28'35" 118'42'W

31'05'N 135.24'W

29.24'N 153.06'W

16-N 17e'E

u
21'58'N 151O19'E

20°N 124'E

Fig. 5-29. X-ray diffraction tracings of the ethylene-glycol-treated ( < 2 pm) clay fraction in the surface
sediments in a traverse across the North Pacific Ocean. After Griffin and Goldberg, 1963. Reproduced
with permission from The Sea, 3. Copyright 1963 John Wiley and Sons, Inc.

amount of illite increases and I/S decreases towards the central Pacific. The illite
content of the I/S phase systematically increases to -
60%. The chlorite/kaolinite
ratio is relatively constant, except for a very low kaolinite value for the westernmost
sample, and both increase as the illite content increases. The 10 A/7 A ratio is
relatively constant and both peaks appear to increase in sharpness towards the
mid-Pacific.
The I/S, with few illite layers, is largely river transported and derived from the
volcanic soils and smectite-rich Cenozoic sedimentary rock on both sides of the
Pacific Ocean. The big question is, is the increase in the proportion of illite layers in
the I/S phase due to authigenic transformation or some other mechanism? The
x-ray patterns of the mid-Pacific clays (center of Fig. 5-29) are essentially identical
to those of samples from the East Chma Sea, which, in turn, are similar to those of
the loess deposits of China. A reasonable conclusion is that the mid-Pacific illitic
material is detrital and was transported mainly from China. Whether the physils
were transported by air or water, or both, is not resolved. There is little doubt some
was transported by air.
330

The most strilung change northward in the North Pacific is an increase in chlorite
(Fig. 5-19). Values are generally > 20% and values higher than 50% occur in the
Gulf of Alaska. This chlorite is a well crystallized IIb variety (high temperature)
(Hayes, 1973), derived from metamorphic rocks along the western coast of Canada
and the southern coast of Alaska. Rivers draining the southern coast of Alaska have
physil suites containing 25 to 35% chlorite (40 to 60% illite, 0 to 12% kaolinite and
trace to 21% expandables) (Naidu and Mowatt, 1983). Glacial input is also signifi-
cant.
The chlorite lobe ( > 20%) extending east from the Japan-Taiwan area presuma-
bly represents input from the Asian Rivers. The chlorite in shelf sediments of the
East China Sea and the Sea of Japan average approximately 25% (Oinuma and
Kobayshi, 1966; Saburo and Kaoru, 1973; Chen, 1978).
The chlorite lobe is superimposed on the western portion of the high illite band,
suggesting an appreciable portion of the illite was also transported by water rather
than by air.
A study of the physils on the Alaskan continental shelf (Naidu and Mowatt,
1983) confirms the abundance of chlorite ( - 53%) and illite in the central Gulf of
Alaska. The high chlorite is related to the occurrence of glacial flour. The currents
transport these physils westward where they are diluted by the influx of expandable
physils derived from the volcanic rocks of the Alaska Peninsula. Expandable clays
are the dominant physil (30 to > 60%) in much of the Bering Sea. Illite is generally
dominant in the coastal areas and to the north in the Beaufort and Chukchi Seas
(Arctic Ocean). The illite is largely locally derived, whereas the expandable physils
are transported northward through the Being Sea by marine currents.
Moberly et al. (1968) found that detrital halloysite and amorphous material, from
weathered basalt, were abundant in the near shore muds of the Hawaiian Islands
and that illite, chlorite and montmorillonite increased seaward. That interpreted this
to indicate the latter suite of physils formed authigenically from the former. More
likely this is an example of dilution of a local water transported source by a long
distance wind transported source.
Several studies have demonstrated a major change, through time, in the physil
suite of North Pacific sediments.
Analyses of core samples offshore from Oregon and in the Gulf of Alaska
(Hayes, 1973) generally confirm the results of previous studies. The Holocene and
Pleistocene muds contain a physil suite of detrital illite, chlorite, vermiculite and
I/S (referred to as montmorillonite in earlier studies). Chlorite is more abundant
( - 45%) and vermiculite and I/S less abundant in the Alaska samples than in the
Oregon samples, reflecting less weathering in the source area. Older samples,
Pliocene to Miocene, differ from the younger samples in that they contain more
I/S, with a high smectite content (Fig. 5-30). Hayes suggested the I/S in the older
samples was formed from submarine volcanics when the sites were at a considerable
distance from land and that the upward increase in detrital minerals is due to the
eastward and shoreward movement of the Pacific plates.
Griffin and Goldberg (1963) found that the physil suites were relatively uniform
with depth until Tertiary sediments were encountered. The latter sediments contain
331
Cloy m i n e r a l s t r a t i g r a p h y a t S i t y 1 7 8 , G u l f of Alaska

002- c

h 001-c
(minorv) I

Middle ~ i o c e n e
(44-2-90-92)
d

Early Miocene (?)


ML (57-1- 125 -127)

1 . 8 . . . . ....I
I . 1 I . ... l . . d
n.eA 3A 3.5A 4A 5AWA 15A ZOA
Fig. 5-30. X-ray diffraction patterns showing relative changes in physil mineralogy upward through time
at Site 178, Gulf of Alaska. From Hayes et al., 1973.

a montmorillonite-rich (I/S) physil suite. The montmorillonite is believed to reflect


the intense volcanic activity that occurred in the western United States during much
of the Tertiary. Airborn ash presumably altered to montmorillonite after deposition
in the Pacific.
A more regional study of core samples from more than 40 wells in the equatorial
Pacific confirmed that the Quaternary physil suite, in much of the area, is drastically
different from that of the Tertiary sediments (Heath, 1969). Fig. 5-31 shows the
variation in quartz content of the 2 to 20 pm fraction of samples from the North
and South Pacific, between 15"N and 10°S latitude, as a function of age. In the
North Pacific there is an abrupt increase in quartz in the Quaternary samples. In the
northwest the increase begins in the late Miocene. There is no equivalent increase in
quartz in the Quaternary sediments of the South Pacific. Based on previous studies
the increase in quartz is assumed to be due to glacial activity and the generation of
332

0
0
12 -
0

00
10 - 0

0 0
8-
c

8 Q
x 6-
c
N

0 .a
O 0 00 . @ a 0.
4- 0
0 .a@
0' a*
0 0 0 .
2-
0 0 0

"0 "0.
0 I I 1

Approx. age (M.Y.)


Fig. 5-31. Quartz content of 2- to 20 p m fractions. 0 = samples taken south of the equator. 0 = samples
taken north of equator, east of 160"W.; 0= samples taken north of the equator, west of 160"W. From
Heath, 1969, Geol. Soc. Amer. Bull. 80, 1997-2018.

enormous quantities of rock flour and loess. The Quaternary sediments of the
mid-North Pacific are essentially loess that was deposited on the sea rather than on
land.
Fig. 5-32 shows the distribution of montmorillonite and illite ( < 2 pm) as a
function of age. From the Middle Eocene to Late Miocene times, montmorillonite
was the dominant physil. These sediments contain little quartz and pyroxene, and
frequently contain zeolites. This mineral suite is the product of oceanic volcanism
(oceanic suite).
During the Late Miocene or Early Pliocene the amount of chlorite, kaolinite and
pyroxene increased in the southwestern Pacific sediments. The geographic distri-
bution and composition suggest the minerals were derived from the andesitic island
arc rocks of the western Pacific (island arc suite).
In the North Pacific significant deposition of the continental suite, illite, quartz
and alkali feldspar, began at the end of the Pliocene. This material appears to have
been mainly transported as wind-blown dust, but some may have been introduced
to the western Pacific by ocean currents.
If the continental suite is assumed to contain one-thrd quartz (based on analyses
of glacial samples), then it contributed < 10%to Middle and Late Eocene deposits
and > 50% to the Quaternary deposits north of the equator. South of the equator,
333

40 -
0
00

c 0
$ 30- 0
a
k m
0

-
r
-
20- 0,”
0
0

0
t
0

o o 0 %:a
0
0 0
000 0.QO :
10- ooo o% om
ooo:.o 0
08 o r ’
r.
0 0
0 ,o ,
Approx. age (M.Y.) Approx. age (M.Y.)
Fig. 5-32. Montmorillonite and illite content of < 2 pm fractions. 0 =samples taken south of the
equator; o = samples taken north of the equator, west of 160”W.; 0 = samples taken north of the
equator, west of 160”W. From Heath, 1968. Copyright 1969 Geol. SOC. Amer.

the concentration through all of Cenozoic time is on the order of 10 to 20% (Heath,
1969).
The only diagenetic changes observed in the physils was a broadening of the
chlorite peaks in x-ray patterns of samples older than Pliocene. However, as in the
present-day Atlantic Ocean, this “degradation” may be due to the formation of
some vermiculitic layers by weathering on land during warmer pre-glacial times. The
other change observed was an increase in median grain size with age. This is due to
the dissolution of opaline tests and the precipitation of the silica as cement.
No systematic change in the “crystallinity” (I/S ratio) of montmorillonite was
found, though all samples older than about 30 m.y. had good “crystallinity” (l/S
with high smectite content). Thus, over a period of 50 m.y. there was no increase in
the proportion of illite layers in the I/S phase; in fact, the reverse situation was
observed. The change more likely reflects a source-climate effect.

South Pacific
The South Pacific bottom sediments are characterized by an abundance of
smectite (Fig. 5-20). Whereas, in the North Paclfic a lobe of detrital illite extends
eastward from central Asia, in the South Pacific a lobe of authigenic smectite
extends westward from northwestern South America. Maximum smectite concentra-
tions are in excess of 70% and in many localities near 100%.
To the south, towards Antarctica, the amount of illite and chlorite increases and
becomes predominant in the region where ice rafting is prevalent. The same physil
suite predominates off the west coast of Australia. Analyses by Windom (1969)
indicate that approximately 40% of the sediment is probably eolian in origin,
334

derived from Australia. Part of the sediment may have a glacial origin and part may
have been transported by rivers from New Zealand.
Other areas with a relatively high chlorite concentration occur around the
Hawaiian Islands and off the northeast coast of Australia. Heath (1969) found
samples from the western equatorial Pacific that had anomalously high concentra-
tions of chlorite. At least in some of the areas the chlorite is fine grained ( < 2 pm),
as opposed to the silt-sized eolian transported chlorite. The associated sediments
contain volcanic glass and pyroxenes, suggesting the chlorite formed from volcanic
material. Whether it formed on land or in the ocean is not known. Carroll (1969)
found that in a 200 cm thick core sample (red clay) from the North Pacific, poorly
crystallized, detrital chlorite became better crystallized and more Fe-rich with depth.
Swindale and Fan (1967) suggested that, in the Hawaiian area, detrital gibbsite was
altered to chlorite in a marine environment. Neither of these studies firmly proves
that chlorite formed in marine sediments.
Kaolinite is relatively abundant only along the east coast of Australia (Fig. 5-18).
The relatively high values in the northern portion of this double-lobed concentration
are believed to be caused by runoff of coastal rivers, and those in the southern
portion to eolian material from the arid interior of Australia (kaolinite 50 to > 90%)
(Griffin et al., 1968). Rateev’s et al. (1969) data for this area are considerably
different. They show a band containing 40 to 60% kaolinite (twice the values of
Griffin et al. (1968)) extending from the East Indies, east by southeast, to the center
of the South Pacific. Reqardless, the equatorial South Pacific contains considerably
more kaolinite than the North Pacific and most of it must be derived from East
Indian and northern Australian soils, formed under tropical humid conditions.
Though detrital physils are shed from the Andes Mountains. They are not
transported very far into the Pacific. In part, this is due to the presence of the
Peru-Chile Trench. Qualitative analyses of samples in and adjacent to the trench
(Zen, 1959) showed that mica-illite was present in all samples. Chlorite, mostly
poorly crystallized, and I/S, with a variety of ratios, are present in the great
majority of samples. The kaolinite distribution is more irregular and in general it is
more common in the samples off the northern coast of Chile. The analyses are not
quantitative enough to establish whether any lateral trends exist. Zen observed
volcanic glass fragments replaced by illitic material. Whether this alteration oc-
curred on the land or in the sea is not known.
The Panama Basin, situated off the west coast of Panama and northwestern
South America, has been studied in enough detail to show the seaward transition of
the detrital physil suite into the authigenic suite (Fig. 5-33). The basin is bounded,
and essentially enclosed, to the northwest by the Cocos Ridge and to the south by
the Carnegie Ridge. The southern portion of the basin contains the Galapagos Rift
zone. The amount of non-biogenous material decreases from 80%, in the area
adjacent to the continental shelf, to < 10%in the western portion of the basin. Most
of the non-biogenous material is < 2 pm in size (Heath et al., 1974).
Fig. 5-34 shows the location of many of the physiographic features discussed in
the following few pages. The distribution of illite and chlorite are similar. The
distribution pattern conforms to the pattern of bottom-water movement. A north-
335
30 85

SMECTITE KAOLlNlTE

90 85 80

ILLlTE CHLORITE

Fig. 5-33. Distribution of physils in surface of the Panama Basin sediments. From Heath et al., 1974;
simplified by Rateev et al., 1980.

ward-flowing current swings west through a gap in the Coiba Ridge. The relatively
high chlorite content reflects the presence of metamorphosed basic and ultrabasic
rocks in western Columbia. The illite has a point source and is derived almost
entirely from several closely-spaced rivers near the Columbia-Ecuador border. The
kaolinite distribution pattern is somewhat different from that of illite and chlorite,
suggesting its transport and settling characteristics are different. The quartz distri-
bution pattern is similar to that of illite and chlorite. Values range from 10 to
> 20%in the nearshore area and decrease to < 1% in the western portion of the
basin, near the Galapagos Islands. Though the basin has a rugged topography, the
physil distribution does not reflect the topography. This indicates the physils are not
transported by density current or bottom-water flow but are transported by shal-
lower water currents. Further, the distribution patterns indicate physils can be
dispersed by marine currents over a distance of at least 1500 km.
The patchy smectite pattern reflects the intermingling of terrigenous and authi-
genic smectite. The area outlined by the < 70% contour is similar to the area in
336

Fig. 5-34. Major physiographic units of the northern Nazca plate. Isobaths in metres. Highs and lows
shaded. Panama Basin is depression east of Galapagos Rift zone. From Heath and Dymond, 1977, Geol.
Soc.Amer. Bull., 88, 723-733.

which illite, chlorite and kaolinite are concentrated, whch suggests most of the
smectite in this area had a terrigenous origin. The presence of two types of smectite
is suggested by differences in ease of expansion. All the smectites expanded when
treated with glycerol, but the smectites in the samples containing > 80% smectite
showed additional expansion when treated with ethylene glycol. The latter response
normally indicates the presence of some highly charged layers. As the sorption
energy for glycol is larger than for glycerol, it will normally expand higher charged
layers than will glycerol. The three areas containing > 80% smectite contain an
abundance of volcanic debris, suggesting the highly-charged smectite is authigenic.
Analyses of samples from the East Pacific Ridge (spreading center), approxi-
mately 1200 km from the western edge of the Panama Basin, confirm the prevalence
of authigenic smectite (Fe-rich) (Rateev et af., 1980). The Quaternary and Pliocene
section, resting on basalt, is composed primarily of Fe-montmorillonite, actually
I/S (10 to 14% Fe,O,, 7 to 13% Al,O,, 1.1 to 2.2% K,O). A typical structural
formula is:

Microscopic examination shows the Fe-montmorillonite is a product of devitrifi-


cation of basaltic glass (authigenic minerals, formed in a marine environment, are
337

Q
M I I

- 5 10 20 30A
Fig. 5-35. X-ray diffraction spectra for untreated < 2 p m fraction core samples from the northeastern
Pacific; top sample is no. 1 and bottom No. 44. The sampling interval is 10 cm. The total length of the
core is 430 cm. Samples 23 to 44 are identical to 22. Ch: Chlorite, M: Montmorillonite, I: illite, K:
kaolinite, Q: quartz, F: feldspar, Cp: clinoptilolite, C: cristobalite. From Aoki et al., 1974. Reprinted
with permission from Deep-sea Res. 21,865-875. Copyright 1974 Pergamon Journals, Ltd.

sometimes referred to as hydrogenous minerals; detrital minerals are called lithoge-


nous minerals). The smectite occurs as nodules or globules, frequently around
volcanic fragments, as cavity fillings in radiolarian, and replacing coprolites. A
similar Fe-montmorillonite (well crystallized with few illite layers) occurrence, to the
north, was described by Aoki et al. (1974). In both occurrences the amount of
detrital admixture (illite, kaolinite and chlorite) increases upward in the section (Fig.
5-35), presumably reflecting an increase in wind and/or water currents during the
Quaternary and a decrease in the availability of volcanic material. Gorbunova and
Shirshov (1976) and Perry (1976) reported a similar increase in authigenic (or
hydrothermal) I/S and the proportion of smectite layers in the I/S, with depth, in
several wells (Pleistocene to Cretaceous) located between the tip of South America
and Antarctica. The crystallinity of the montmorillonite (I/S) was best developed in
horizons immediately above the basalt. Chemical data indicate the lower samples
are Fe-rich (Drever, 1976) and the smectite is presumably an Fe-montmorillonite.
338

Oxygen isotope data and other information indicate the smectitic material on top of
the basalt was formed from the submarine weathering of volcanics; terrigenous
physils dominate the overlying sediments. The transition between the two physil
suites can be relatively abrupt.
Analyses of cores off the coast of Antarctica indicate the change in the physil
suite begins in Upper Miocene sediments. Jacobs (1974) believes this is due to early
glaciation on the Antarctic continent.
It appears that throughout much of the Pacific Ocean there is a major change in
the physil suite near the end of the Pliocene, and in some areas earlier, with
continental detritus becoming predominant. The change is in part due to a decrease
in volcanic activity and an increase in glacial activity and atmospheric transport.
Studies of sediment mounds near the Galapagos Spreading Center (GSC) indi-
cate Fe-rich physils apparently formed in a hydrothermal environment (Rateev et
al., 1980; Hoffert et a/., 1980; Dymond et al., 1980). Small mounds adjacent to the
GSC are composed primarily of small black granules of Mn oxides and physils and
dark green to black physil-rich mud. The physils have a higher iron content (27 to
30% Fe,O,) than the Fe-montmorillonites and are classed as nontronites. The x-ray
patterns and chemical data (2 to 4% K,O) indicate the physils are mixed-layer
celadonite-nontronite. The Ce,” apparently formed by reactions between basalts
and hydrothermal brines rich in potassium. An Fe-montmorilline phase may have
formed as the original precipitate and later altered to Ce/N.
There is no doubt that volcanic material can alter to smectite and I/S in the
marine environment (for discussion, see the following chapter), but is it presently
forming on the sea floor and is it volumetrically important? Occurrences of
Fe-montmorillonite in the Pacific Ocean bottom sediments have been described by
Bonatti (1963), Griffin et al. (1968), Sayles and Bischoff (1973), Heath and Dymond
(1977), Hein et a/. (1979), etc. Regional studies of the Nazca plate (0” to 25”s
latitude and 80” to 115”W longitude) (Heath and Dymond, 1977) and the DOMES
area west of the East Pacific k s e (EPR) (10” to 20”N latitude and 125” to 155”W
longitude) (Hein et a/., 1979), indicate “authigenic” Fe-montmorillonite is abundant
in Pacific Ocean bottom sediments.
The Nazca plate is bounded on the west by the EPR. The Bauer Deep and
Central Basin lie between the EPR (spreading center) and the Galapagos Rise, near
the center of the plate. Fe-rich smectite is present in both the rise and basin
sediments but is most abundant in the latter. The EPR sediments contain on the
order of 30% Fe. The source of the Fe is the “hydrothermal” reaction that occurs
when hot, reducing solutions emanating from newly formed basalt react with cold
oxidizing sea water. The Fe is rapidly leached from the basalt and forms amorphous
hydroxide flocs. A relatively small portion of the Fe, and other adsorbed elements,
react with biogenic silica to form Fe-rich smectite. Deep water currents carry some
of the Fe-hdyroxide flocs and smectite into the basins where the surface-active
hydroxide reacts with biogenically deposited silica to form more Fe-rich smectite.
About two-thirds of the hydrothermal Fe in the Bauer Deep and Central Basin
samples is in the form of smectite. On the basis of x-ray and chemical analyses of
smectite from the Bauer Deep, Cole and Shaw (1983) called the material nontronite.
339

The material contains very little tetrahedral A1 and only 50% of the filled octahedral
positions contain Fe3+.Generalized formula:

(Fe:.&AI 0.44 Mg0.54 0.1 1 ) (si3.WA1 0.03 (Ca 0.l o N a 0.40 0.12 >OI0 (OH),
The smectite has a composition intermediate between that of montmorillonite and
nontronite and probably should be called Fe-montmorillonite (Fe-M).
In the DOMES area Mn nodules are present in silicious ooze and red clay. Fe-M
comprises approximately half the physil suite, illite 30 to 40% and the remainder
chlorite and kaolinite. On the basis of 6 0 l 8 data Hein et al. (1979) concluded the
Fe-M was authigenic. (Detrital smectites have 6 0 " values < +22%0, whereas
authigenic smectite formed at 0°C in the deep sea has a value of +31%0.) On the
basis of the absence of volcanic material and the high percentage of some trace
elements, the authors suggest the Fe-M formed by the low-temperature reaction of
Fe oxyhydroxides and silica in a deep ocean-floor environment. They propose that
the Si was biogenic and the Fe hydroxide was transported from the East Pacific
Rise, 4000 to 5000 km to the east. There is a problem with the source of Al. One
possible source is Al adsorbed on biogenic silica. Another possible source is eolian
gibbsite. They calculate that over an area of approximately 2x106 km2 the mean
clay content is 30% and that authigenic Fe-M comprises at least 10% of the total
sediment.
Thus, it is apparent that a significant, if not major, portion of the smectitic
material in large areas of the Pacific Basin is authigenic, probably hydrothermal-
authigenic; the reaction rate between cold basalt and cold ocean water is apparently
too slow to produce appreciable smectite during a short time period. The relative
abundance of smectite in the South Pacific maybe due to the presence of the
spreading center (EPR) and the paucity of detrital physils. However, a word of
caution: The physil suite of atmospheric dust samples from the vicinity of the
Galapagos Islands contains as much as 63% smectite (Prosper0 and Bonatti, 1969).

Indian Ocean

Southern Indian Ocean


On the basis of the physil distribution, the Indian Ocean can be divided into
three areas--southern, northern and middle. Mite is the dominant physil in the
southern ocean. Illite and chlorite are present in a wide band parallel to the coast of
Antarctica and were derived from that continent. The illite-rich sediments off the tip
of southern Africa do not contain much chlorite, suggesting little, if any, Antarctic
contribution. The illite is presumably derived from the abundant shales and low-
grade metamorphic rock in southern Africa.
It is generally assumed that most of the Antarctic sediments are delivered to the
ocean by ice-rifting. A detailed study of the southeast Indian Ocean, between
Australia and Antarctica (Wilkes Land) indicates a variety of transport mechanisms
are operative (Moriarty, 1977). The two continents are separated by the Southeast
340

WilkesLand

Fig. 5-36. Complex distribution of marine physils in the area between Australia and Antarctica. Black
dots show location of talc. M = montmorillonite. Data compiled from Moriarty, 1977.

Indian Ridge. Physils are abundant on the flanks of the two continents and scarce
on the southeast Indian ridge where opaline silica is abundant ( > 40%).
Fig. 5-36 is a generalized map showing the complex distribution of the physils i n
the area between Australia and Antarctica. The complexity is increased when the
physils are further characterized on the basis of Fe content and crystallinity. The
physil distribution indicates the presence of numerous source areas. The ridge area
is characterized by the presence of abundant Fe-montmorillonite. It is flanked on
the north and south by high illite zones. The illite in the southeastern area has a
broad 10 A peak and a high 10 A/5 A
ratio suggesting it is an I/S physil with a
high illite content. A chlorite rich zone is present south of New Zealand and a
kaolinite zone off the southwest coast of Australia.
The I/S in the southern basin, and a variety of other physils, were derived from
the Wilkes Land; however, studies of bottom sediments indicate the major transport
mechanism was turbidity currents rather than ice-rifting. The illite and kaolinite,
along with gibbsite and I/& along the southern coast of Australia were apparently
largely wind transported from the deeply weathered Western Australian desert. The
physils in the high chlorite zone. with well crystallized illite-muscovite, were
presumably mechanically eroded by glaciers on the South Island. New Zealand.
Greenschist rocks and glaciers (and large trout) are abundant on South Island.
Though several types of montmorillonite are present, abundant well-crystallized
Fe-montmorillonite is the dominant physil associated with the opaline sediments o n
the southeast Indian Ridge. Moriarty suggests this material was transported from
Antarctica. In view of what we have learned about the origin of Fe-montmoril-
341

lonites in the Pacific Ocean, it is more likely that the smectite on the southeastern
Indian ridge is authigenic, formed by the reaction between Fe released from the
ridge basalts and opaline silica. Talc is commonly present in samples located on the
mid-ocean ridge, in areas of low or negative accumulation rates of physils. Most
likely it was produced by hydrothermal activity. There is relatively little crop
dusting in Antarctica.

Northern Indian Ocean


Perhaps the most striking feature in the northern area is the concentric distribu-
tion of montmorillonite adjacent to the Indian Continent (Fig. 5-21). In the coastal
areas montmorillonite is in excess of 70% and decreases seaward to 30 to 50%, and
lower. It is derived largely from soils formed on the basaltic Deccan Traps and
transported by rivers to the coast (Goldberg and Griffin, 1970).
East of India, in the Bay of Bengal, montmorillonite is diluted by the physils
transported by the Ganges and Brahmaputra Rivers, which contain abundant illite
and chlorite in their suspended load. Both rivers have their headwaters in the
Himalayas and presumably derive much of their physil material from the moun-
tains.
The most striking feature of the Arabian Sea is the north-south band of illite-rich
sediments. Near the equator the band assumes an east-west alignment. Much of the
Arabian Sea is underlain by the Indus Cone, indicating an appreciable portion of
the sediment was derived from the Indus River. The headwaters of the river are in
the Himalayas and for a considerable portion of its length it flows through desert
country. Illite is normally the major physil in both these types of source areas. The
elongate pattern is presumably due to dilution on both flanks, primarily by
montmorillonite from India and Africa. To the north and west the Arabian Sea is
flanked by major desert regions and, on the basis of dust concentrations, Goldberg
and Griffin (1970) calculated that on the order of 15 to 20% of the sediment could
be eolian. In dust samples collected from the edge of the desert in northwest India,
illite was the major physil (47 to 62%). The high illite and accompanying chlorite in
the equatorial band are believed to have been transported by winds from the arid
regions to the north and west.
Chlorite, in concentrations greater than 30% ( > 40% in a small area), is present in
the extreme northwestern portion of the Arabian Sea (Gulf of Oman) (Fig. 5-19).
Chlorite is not particularly abundant in the Arabian Gulf (Persian Gulf) or in the
Indus River. The source rocks were most likely the ophiolites and other volcanic
rocks in southern Iran and perhaps the Oman Mountains in Saudi Arabia (southern
flank of the Gulf of Oman). Dust samples from off the coast of Irar. (west of
Goldberg and Griffin samples) have an average composition of 54% illite, 30%
chlorite, 10% kaolinite and < 5% smectite. The physil content (based on Al
concentration) is among the highest found for marine regions (Chester, 1985). Thus
it is likely that a large portion of the physils in the sea west of the Indus River are
eolian. But, keep in mind the Indus River discharges 100 X l o 6 tons/year of
sediment to the Arabian Sea and regionally must be the dominant source.
Palygorskite is relatively abundant, and in some areas the dominant physil, in the
342

Fig. 5-37. Distribution of palygorskite/illite peak-height ratios in the Arabian Sea sediments. Most of the
samples with less than 0.4 ratios have almost no palygorskite. From Kolla et al., 1981. Copyright 1981
Soc. Econ. Paleo. Miner.

northwestern Arabian Sea (Kolba et al., 1981; Goldberg and Griffin, 1970), the
Arabian Gulf (Al-Bakri et al., 1984) and the Red Sea and Gulf of Aden (Heezan et
a/., 1965). Palygorskite is abundant in Cenezoic rocks and soils of the Middle East
and northern Africa (Weaver and Beck, 1977; Callen, 1984) and is the obvious
source of the palygorskite in the various bodies of water. The distribution and
character of the associated minerals indicate much of it was wind transported. The
distribution in the Arabian Sea (Fig. 5-37) indicates a predominantly western source
from the deserts of the Arabian Peninsula and Somalia. Palygorskite is abundant in
the same general area of the Arabian Sea in sediments ranging in age from
Pleistocene to Cretaceous. In the younger sediments the concentration increases
from southeast to northwest, as the Arabian Peninsula is approached (Weaver and
Beck, 1977).

Middle Indian Ocean


Smectite is the predominant physil over much of the middle Indian Ocean area.
As suspended samples from this area contain an average of 14% smectite (Chester et
d., 1974) most of the smectite was presumably formed authigenically, probably
hydrothermally, from volcanic ash and basalts in the rise areas, which comprise
343

much of the ocean floor. The kaolinite distribution is interesting. The kaolinite
concentration in the vicinity of Madagascar is due to run-off of material from
tropical soils, as would be expected. The lobe of high kaolinite extending from the
west coast of Australia largely reflects transport by southeasterly winds of material
from the western Australian desert. Though illite is usually associated with desert
climates, the Australian desert soils commonly contain from 60 to 90% kaolinite. In
pre-Recent times the area was one of intense lateritic weathering (Griffin et al.,
1968) and the kaolinitic soils are still well preserved. This latter sediment-soil
scenario demonstrates one of the problems in using clays to determine paleocli-
mates. The physil-climate association is as it should be but there is a time lag of
millions of years. Theoretically, one should be able to examine Deep Sea Drilling
cores from the offshore area and by determining the age at which a significant
increase in kaolinite occurs, date the beginning of lateritic weathering, assuming the
wind direction has not changed. The DSDP cores (Vol. 26 and 27) indicate the
Cenezoic stratrigraphic record is highly incomplete but does show that a relatively
kaolinite-rich physil suite has been deposited since about Middle Miocene time.
This Page Intentionally Left Blank
345

Chapter VI

“AUTHIGENIC MARINE” PHYSILS

In the previous chapter, I discussed the distribution of detrital physils in the


oceans and glossed over the physils which actually form in the oceans or peri-marine
environments. The previous chapter was concerned primarily with physical processes;
this chapter is concerned primarily with chemical processes which modify detrital
minerals and/or create new minerals. I have used the term “authigenic” to include
physils formed or modified at ocean water temperatures and those formed at
elevated temperatures (hydrothermal).

EXCHANGE REACTIONS
Most of the ions in the ocean have been transported there by rivers. In addition
to the ions in solution easily exchangeable cations are also transported by the
physils. The cations adsorbed on the physils do not add to the ionic populations of
the oceans but by exchanging the adsorbed river ions for those concentrated in the
ocean they modify the ocean chemistry. The exchange capacity of the river sus-
pended suite depends on the physil suite, plus organics and amorphous material,
carried by the rivers. Kennedy (1965) found that the CEC of the clay fraction
( < 4 pm) of streams from the eastern United States range from 14 to 28 meq/100 g;
streams from the central and western United States have values ranging from 18 to
25 meq/100 g. The high exchange capacity values are characteristic of streams with
a high proportion of montmorillonite and/or vermiculite. The CEC of the silt
fraction ranged from 4 to 34 meq/100 g. Most of this exchange capacity is due to
the presence of shale fragments and clay aggregates. In most rivers the majority of
exchangeable cations are carried by the physils but organic material can account for
an appreciable proportion of the exchangeable cations in the river load.
The ratio of cations adsorbed on suspended sediment to cations in solution is
extremely variable (0.0007 to 2). In general, when the suspended sediment con-
centration is high the ratio tends to approach one and can exceed one in some
western streams (Kennedy, 1965).
As physils transported by fresh water streams enter marine waters they readjust
to the change in water chemistry. Various investigators have suggested that K and
Mg are “permanently” incorporated between expanded physil layers to form illite
and “chlorite” respectively. The magnitude of these reactions has not been de-
termined. However, a major change does occur in the composition of the adsorbed
cation suite.
346

Table 6-1
Exchangeable ions on fluvial sediments, meq/100 g and percent of total (modified from Sayles and
Mangelsdorf, 1977).
Ca Mg Na K
Before sea water exposure 2.75 0.99 0.54 0.38
59 21 12 8

After sea water exposure 1.18 2.22 2.14 1.06


18 34 32 16

A number of investigators have measured the change in composition of the


adsorbed cation suite as clays are mixed with sea water (Carroll and Starkey, 1960)
and as river clays are deposited as marine mud (Russell, 1970). Sayles and
Mangelsdorf (1977) demonstrated earlier results were, to a large extent, an artifact
of the laboratory technique. Washing with distilled water, H,O, treatments etc.
modifies the adsorbed cation suite. They developed a method that eliminates the
washing procedure and obtained results that are more realistic. Table 6-1 shows the
median values obtained when 32 fluvial sediments from the Atlantic seaboard were
exposed to seawater. The major difference from previous studies is a high N a value
for the sea water treated samples. Note that the total cation suite increases from
4.66 to 6.60 meq/100 g.
Table 6-2 shows the cation suite of several physils after they were equilibrated
with artificial river water (mean world river). The cation suites are similar to those
of natural samples.
The exchange composition of some physils and natural muds, equilibrated with
sea water are summarized in Table 6-3. The top four samples were first equilibrated
with river water (Table 6-2).
These data indicate that for most of these samples Na is the dominant cation
comprising close to 50% of the exchangeable cations. Mg is the next most abundant
ion, ranging from 22 to 47%. The Arizona montmorillonite is the only sample with
Mg higher than Na. In general Arizona bentonites have about twice as much Mg in
the octahedral layer as the other montmorillonites used. Note also that illite has the

Table 6-2
Exchangeable ions on river physils, meq/100 g and percent of total (modified from Sayles and
Mangelsdorf, 1977).
Ca Mg Na K H -L
Arizona 48.8 26.8 1.6 1.o 1.3 79.5
Montmorillonite 61 34 2 1 2
Texas 57.6 18.4 2.3 0.6 2.0 80.9
Montmorillonite 71 23 3 0.7 2
Wyoming 43.0 19.4 2.6 0.3 3.4 68.7
Montmorillonite 63 28 4 0.4 5
Bath 3.2 1.1 0.4 0.05 1.5 6.2
Kaolinite 51 18 6 0.8 24
347

Table 6-3
Exchangeable ions on physils and muds equilibrated with sea water, meq/100 g and percent of total
(modified from Sayles and Mangelsdorf, 1977).
Ca Mg Na K z:
Arizona 19.1 37.7 19.4 3.2 78.1
Montmorillonite 24 47 24 4
Texas 15.7 18.1 44.3 2.7 80.8
Montmorillonite 19 22 55 3
Wyoming 1.o 28.2 39.2 2.7 71.1
Montmorillonite 1 40 55 4
Bath 1.5 2.0 2.4 0.4 6.3
Kaolinite 24 38 38 6
I/S - 14.2 21.6 2.7 38.5
- 36 56 7
Fithian 1.8 4.1 7.9 2.8 16.8
Illite 11 24 47 17
Pacific Sediment' 18 29 39 14
Black Sea Sediment ' 22 28 37 13
' Zoytseva (1966), as reported by Sayles and Mangelsdorf (1977).

highest exchangeable K value. The anomalous values for these two samples suggests
that for some physils mineral selectivity may have an influence on the composition
of the adsorbed cation suite.
The exchangeable cations were measured by analysing the decrease in the amount
of cations in the seawater in which the physils were immersed. Thus, the exchangea-
ble cations on the physils were not directly determined. It is likely that some
relatively high charge sites were present in the illite and these sites preferentially
attracted, and perhaps fixed, K. The high-Mg montmorillonite may contain a few
scattered brucite islands that acted as a nucleus or template and preferentially
attract Mg.
Sayles and Mangelsdorf (1977) calculated, assuming an average C.E.C. of 25
meq/100 g for river suspended material, that the physils adsorbed 2.3 X 1015
meq/yr Na+ or nearly 40% of the Na released by continental weathering. Released
Ca2+increased the flux of river-borne Ca2+by slightly less than 10%.
Following their initial study, Sayles and Mangelsdorf (1979) analysed a series of
suspended samples from the Amazon River before and after exposure to sea water.
Their results are, in general, similar to those they obtained using pure physil
samples. Analyses of the exchangeable cations from samples collected along the
main stream gave the following values: Ca2+= 79-8696, Mg2+= 12-19%, K + =
7-16% (12%more typical value), Na+ = 0.3%. The values for H + range from 5 to
22% but H + was not included in the calculations to determine the percentage
composition of the exchangeable suite. The Ca/Mg ratio of the exchangeable
cations is linearly related to the Aca/AMg activity ratio of the solution in equi-
librium with the sediment. The CEC ranges from 13 to 31 meq/100 g. The CEC
increases with decreasing grain size. Values range from 40 meq/100 g for the
c 2 pm fraction to 8 meq/100 g for the > 74 pm fraction.
348

I I
I I
I

-Ti
I
I
I
8 1 I
1 L - l -
T I
I
I
I
L--

-- Mg I
I
I
L-l
I I I
I I I I I
20 30 40 50 ( 0
" I o of exchangable cations
Fig. 6-1. Histogram showing the distribution of exchangeable Na and Mg, as percent of total exchangea-
ble cations, for physils exposed to sea water. Data from Sayles and Mangelsdorf, 1977, 1979.

When the river samples are equilibrated with seawater Na and Mg become the
dominant exchange cations. They are present in roughly equal proportions (40%
each) in most samples; however, Na/Mg ratios range from 0.4 to 2.2. Ca values
range from 0 to 20% with most values being near 15%. K values are relatively
constant at 8%. Fig. 6-1 shows the distribution of the exchangeable Na and Mg
values for samples equilibrated with seawater by Sayles and Mangelsdorf (1977,
1979). The modes for both histograms are relatively broad and the overall range of
values is large. This suggests there is not one uniform exchangeable ion suite
determined by the composition of seawater, but that mineral selectivity is a factor.
For the samples collected during the river flood stage the concentration of
exchangeable cations is small relative to the dissolved cation concentrations. How-
ever, for both Mg2+ and Ca2+the concentrations are close to 10% for each. When
the average annual cycle of the Amazon is considered the influence of cation
exchange on the dissolved cation suite delivered to the ocean is less than 10%for all
species. The low value is, in part, due to the low suspended so1id:dissolved-solid
ratio of 1.7. The world average ratio is considered to be 4.
Though Sayles and Mangelsdorf's (1979) experiments showed that the total
amount of cations adsorbed from seawater was equal to the cation exchange
capacity of the physil, they did not establish if the adsorbed seawater cations were
exchangeable. Weaver (1958) demonstrated that high charged (> 150 meq/100 g)
349

expandable physils were able to “fix” K from sea water and contract to 10 A.
Materials that have the potential to do this are largely micas and illites that have
had much of their K removed during weathering but still retain much of their layer
charge. These stripped physils may resemble vermiculite or montmorillonite in
regard to their swelling characteristics. Roberson (1974) soaked several soil
vermiculites and a soil montmorillonite in seawater for 11 weeks and found that
only 53 to 56% of the cations adsorbed from seawater were exchangeable (NH,Cl
and BaCl were used to replace the adsorbed cations). None of the K was exchangea-
ble. X-ray analysis confirmed that the K had collapsed some of the layers to 10 A
and formed an I/S physil. Roberson’s soil vermiculites had a lower layer charge
than Weaver’s, explaining why all the layers did not contract to 10 A. Vermiculite is
commonly present in the rivers along the east coast and Gulf coast but is not
observed in the estuaries or bays. It is likely that much of this material has “fixed”
K and reverted to its mica-illite precursor.
The other possible reaction is the precipitation of Mg(OH), in the interlayer
space of expandable clays to form a chlorite-like physil. This can be done in the
laboratory quite easily but normally at p H values higher than those encountered in
estuaries. It has not been satisfactorily demonstrated that this reaction occurs in a
normal marine environment. Drever (1971) found that Mg was depleted in the
shallow cores from Banderas Bay and believed this Mg was incorporated in the
octahedral layer of montmorillonite, replacing Fe. His samples appear to contain a
minor amount of chlorite, so it may be that the Mg replaces Fe in the brucite layer
of the chlorite.
Various studies of estuarine physils have established that an expandable phase is
A
present that does not collapse to 10 when heated at moderate temperatures. The
assumption is usually made that Mg(OH), has precipitated as islands in the
interlayer space and formed a chlorite-like mineral. This is possible, but another
possibility is that organic material is present in the interlayer space.
Montmorillonites transported by the Mobile and Pascagoula Rivers to Mobile
Bay apparently contain some interlayer humic material (Milne and Shott, 1958).
Temperatures of 350°C to 450°C are required to collapse the montmorillonite to
10 A. The organic material is less abundant in the marine muds but is present. The
role of organic matter has not been fully evaluated but in some rivers is a major
factor. McCrone (1967) found that in the Hudson River Estuary the organic
material (5 to 6%) in the clayey silts accounted for 65 to 80% of the cation exchange
capacity. Hydrogen ions occupy approximately 2/3 of the exchange sites.
And, as discussed in the section on estuaries, there is evidence that much of the
inorganic material in the brackish and marine environments contains a thin organic
and/or metal (Fe, Mn) oxide surface coating. These coatings have been used to
explain the adsorption behavior of such metals as Cd, Cu, Pb (Lion, et al., 1982).

PHY SIL DISSOLUTION AND REPRECIPITATION


In addition to the processes of cation exchange and cation fixation, physils
dissolve, to some extent, upon contact with seawater. When placed in sea water,
350

physils rapidly release Si to sea water. Except for kaolinite, they release within 10
days approximately 50% of the amount they release over a period of several years.
Concentrations range from 2 to 20 ppm. The amount released is largely related to
surface area. Thus, montmorillonites provide the higher values and kaolinite, illite,
and chlorite, the lower values. A value of 20 ppm corresponds to a loss of 0.04% of
the mass of the solid (Mackenzie and Garrels, 1965; Lerman et al., 1975). The
amount of soluble A1 is much lower and is not normally measured. The amount
present in sea water is on the order of 0.001-0.002 ppm.
Mackenzie and Garrels (1966) proposed that authigenic physils begin forming in
sediments and overlying waters shortly after river borne detritus contacts seawater.
Partly because of the relatively high solubility of Si relative to Al, a Si-rich type of
physil should be formed. More recently, Mackin and Allen (1984) and Mackin
(1986) determined the Si and A1 content of pore water in shallow marine muds and
in the overlying sea water. In the pore waters, Si increased and A1 decreased with
depth. The water overlying core samples was sampled over a 12-hour period to
estimate fluxes of solutes across the sediment-water interface. Si increased and A1
decreased with time. In order to maintain a low concentration of A1 and a high
concentration of Si, a solid must form which is relatively aluminous in comparison
to starting materials.
They concluded the Al-Si-H relation was a thermodynamic equilibrium relation
+

which was very rapidly attained. Calculations indicated the relation was not due to
equilibrium with respect to the detrital material, but was due to equilibrium with
respect to an authigenic surface phase. They suggested the authigenic mineral was a
dioctahedral chlorite (Mg,,,,Al 5,0Si2,70,0(OH)8), the only reasonable mineral with
a Si/Al < 1. However, all calculations of this type indicate that the amount of
authigenic physil formed would be so small as to be virtually undetectable.
On the other hand, a study in the Tamer Estuary, southwest England (Morris et
al., 1986) found that dissolved A1 is generated in the water column through
desorption and/or dissolution from tidally resuspended sediment particles in the
high salinity portion of the estuary and is removed from the water by sorption onto
resuspended sedimentary particles in the very low salinity region.
Delange (1986) and Delange and Rispens (1986) found that in deep sea turbiditic
sediments the concentrations/depth profiles of dissolved iron ( Fe z + )and silica were
similar and that both were more abundant in intervals with a relatively high organic
content. The increase in soluble Fe2+ is due to the reduction of Fe3+ during the
decomposition of organic matter. The removal of iron oxide and/or organic matter
from the surface of amorphous silica particles increases the surface area and
presumably the solubility.
Fe and Si, which diffuses upward into oxidizing sediments, precipitates as Fe
hydroxide and possibly quartz. (Why not glaucony or betluerine?) The Fe and Si
that diffuses downward into suboxic amorphous sediments is believed to precipitate
as an amorphous Fe-Si material similar to nontronite. Calculations suggest the
amount of physil-like material that could form, in these particular samples, would
be < 0.1%. The covariation of Fe, Si and organic material is of particular interest in
regard to the origin of glaucony (p. 386).
351

FORMATION O F PHYSILS FROM MARINE VOLCANICS

Introduction

Volcanic material generated at spreading centers and submarine volcanoes reacts


with seawater to produce a variety of physils, primarily smectites, "celadonites" and
zeolites (and a variety of less abundant minerals). Physils are formed at tempera-
-
tures ranging from 0" to 500°C. There are three proposed modes of low-tempera-
ture formation: alteration of cooled volcanic rocks at low temperatures, which is
extremely slow; low-temperature reaction of biogenic silica with Fe-oxyhydroxides
released from hot basalt when it encounters seawater; and precipitation and
alteration by hydrothermal fluids.
Molten basalts have a temperature of approximately 1300°C. During slow
cooling of oceanic basalts the formation of physils begins at -
500°C is very rapid
between 400" and 200°C and slows down considerably at the final stage of cooling
(Kurnosov et al., 1981). Apparently, along much of the mid-ocean ridge system
hydrothermal fluids have altered basalt to zeolite and greenschist facies at tempera-

Formation of clay minerals


for thick basalt f l m
and SINS only

400 -

300 -

200 -

100 -

01 v
Fig. 6-2. Diagram showing the formation of clay minerals as a function of temperature in oceanic
environment from basalts; based on natural data. From Kurnoson et al., 1981. Copyright 1971 Elsevier.
Pub. Co.
352

tures of 200" to 300°C (Deffeyes, 1970; Miyashiro et al., 1971; Tomasson and
Kristmansdottir, 1972; Williams el al., 1974).
Experiments on the hydrothermal alteration of basalt by seawater (Mottl and
Holland, 1978) indicated that smectite was the only physil formed in the tempera-
ture range of 200" to 400°C. At 500°C smectite and talc formed. Fig. 6-2 shows the
sequence of formation of physils during the slow cooling of thick basalt flows and
the warming of basaltic material by hydrothermal fluids. Trioctahedral Fe Mg
smectites form at relatively low temperatures in reducing alkali conditions (top of
the basalt). It is the most widespread physil in basalts and occurs in the ground-
mass, veins and vesicles. Under oxidizing conditions, in the upper part of the basalt,
celadonite, dioctahedral Fe montmorillonite, nontronite and Fe hydroxides form at
temperatures near 20°C. During warming, the penetration of sea water, under
oxidizing conditions, into previously formed basaltic complexes leads to the forma-
tion of dioctahedral smectites and micas (from alkaline basalt). When seawater is
replaced by juvenile solutions the Mg concentration (released from basalt at high
temperatures) can be sufficiently high that chlorite and Ch/S (corrensite) form
(Kurnosov et al., 1981). Juvenile solutions and/or seawater presumably account for
the formation of saponite, palygorslute and sepiolite which occur in many ridge
areas (Skornyakova et al., 1979; Bowles et al., 1971).
Many of these neoformed physils have exotic shapes. Kurnosov et al. (1981)
describes morphologies as rosette-like, lacy, leaf-like, tubes, flower-like and hair-like.

High-Temperature Reactions

The crust beneath the normal oceanic basin is about 6 km thick. Assuming a
geothermal gradient of 20 or 30"C/km, the temperature at its base is about 120" to
18OOC. Though the crust is thinner in the crest of the major ridges the heat flow and
geothermal gradients are much greater so that extensive metamorphic recrystalliza-
tion can occur. Temperatures at the base of the crust beneath the ridge crest should
range from 150" to 450°C (Miyashiro et al., 1971), or about 300°C (Melson et al.,
1968). Thus, basalts and gabbros in the ridge areas are subjected to burial metamor-
phism, with the metamorphic grade increasing from the top to the base of the crust.
Due to faulting and the formation of rift valleys it is possible to obtain samples of a
relatively complete section of the crust.
Most of the metabasalts are in the zeolite and greenschist facies. The base of the
basaltic section commonly consists of amphibolite facies rocks containing chlorite.
The greenschist rocks are composed largely of chlorite and quartz. The greenschist
have the same Fe content as the unmetamorphosed basalt. Upward (cooler) the
chlorite gives way to Ch/S and vermiculite. Weathered basalts contain mixed-layer
smectite/vermiculite. Smectite, in varying amounts, is commonly present in all
facies but is most abundant in the weathered facies (Miyashiro et al., 1971). In
many instances the smectite is saponite which was formed by hydrothermal fluids.
The saponite replaces the massive basalt as well as occurring in veins and replacing
olivine and chlorite. The saponites commonly contain minor amounts of celadonite,
353

chlorite, talc, Ch/S and I/S. Note that though the high temperature smectite is
commonly saponite ( - 20% MgO), the smectite that forms at near ambient ocean
water temperatures is commonly either Fe-rich montmorillonite or nontronite.
The greenschist facies chlorites are considered to be Fe-rich. They contain almost
equal amounts of FeO and MgO ( - 20%)(Melson and van Andel, 1966). Siever and
Kastner (1967) reported the presence of Mg-rich chlorite in ridge detritus and
suggested that Fe-rich chlorite altered to Mg-rich chlorite in the ocean. It is also
possible the Mg-rich chlorite is hydrothermal.
In the hydrothermally altered basalts in Iceland, on the Mid-Atlantic Ridge,
Fe-rich chlorite of variable composition forms at temperatures higher than 230" to
250°C. Ch/S is the dominant physil between 200' and 240°C. At temperatures
lower than 200°C only Fe-rich saponite is present. Montmorillonite occurs in the
near-surface geothermal acid leached zone (Kristmannsdottir, 1978). A variety of
zeolites forms throughout the same temperature range. The geothermal fluids, which
were originally seawater, are depleted in Mg and enriched in K, Ca and Si.
On the basis of mineralogic analyses of detritus from ponded basins on the flanks
of the Mid-Atlantic Ridge, Siever and Kastner (1967) concluded that the bulk of the
sediment was derived from the continents. From 1 to lo%, primarily chlorite, was
believed to be ridge material. Analyses of the REE distribution of the chlorites
(coarse) and fine-grained montmorillonite indicated the former, a minor component,
was derived from the ridge and the latter had a continental origin (Copeland et al.,
1971). Thus, the contribution of ridge detritus to ocean sediments is probably
insignificant on a global scale.
Mottl (1983) has proposed a model to explain why some of the hydrothermal
ridge physils are Fe-rich and others Mg-rich. Both field data and laboratory
experiments indicate that during basalt-seawater interactions, over a temperature
range of at least 70 to 50O0C, Mg2+ is rapidly removed from seawater into
secondary minerals such as smectite, chlorite, tremolite-actinolite or talc. The rate of
removal increases with increasing temperature. The Mg2+ uptake is balanced
primarily by the loss of Ca2+ (to form anhydrite). The Mg2+ is released as
Mg(OH),, which results in the generation of H + and a drop in pH. Above 15OoC,
K + is leached from the basalt and no K-rich phases form.
Seawater contains 0.13% Mg (0.22%MgO). Up to a seawater/rock ratio of 50,
basalt can remove nearly all the Mg2+ from seawater. Thus, the Mg/Fe ratio of the
secondary physils is largely dependent on the seawater/rock ratio. At - 300°C and
a seawater/rock ratio > 50, the stable assemblage should be Mg-rich chlorite and
quartz.
It is important to note that basaltic magma is undersaturated with respect to
water and contributes little or no juvenile water to the hydrothermal system.
Modified seawater is the main heat (and ion) transfer medium in geothermal
systems (Spooner and Fyfe, 1973).
The seawater/rock ratio in hydrothermal areas near the axis of mid-ocean ridges
is about one (i.e., Galapagos), whereas, areas farther from the axis have a much
higher ratio. In these latter areas, which have a steep temperature gradient, down-
welling seawater supplies the Mg to produce a relatively Mg-rich chlorite and quartz
354

suite. A convection system is created and the hydrothermal waters ascend near the
ridge axis. The only source of heat for the warm-spring solutions (Galapagos) are
these ascending solutions whch are cooled when they mix with cold seawater. In the
latter area temperatures are lower and little Mg is available (seawater). The resulting
physils, nontronite-type, are relatively Fe-rich. I have taken some liberties with this
model.

Low-Temperature ( < 100°C) Reactions

There have been many studies of the surface weathering of basement basalts (see
Initial Reports of the Deep Sea Drilling Project). I will review only a portion of the
data. The most abundant physil is commonly smectite, but significant amounts of
celadonite, Ce/S, Ch/S, talc and chlorite are present.
Young basalts show only the early stages of weathering, i.e., thm (up to 1 mm)
surface coatings, infilling of some vesicles and veins, very occasionally phenocryst
alteration, and minor patchy replacement of the groundmass. Generally only a few
percent of the rock is altered. The weathered product is commonly smectite and
associated calcite. The calcite is formed from Ca released during the weathering of
the basalt. The smectites range in color from bright green to orange-brown.
Humphries et al. (1980) found that in weathered basalts (0.5 to 3.4 m.y.) from the
East Pacific Rise, both high Fe, K, and high Mg smectites were present. Most
samples are rich in Fe and K, with iron (reported as FeO) ranging from 26 to 30%
and K,O values mostly in the range of 4 to 6%. The iron is presumably in the ferric
form and the physil is a mixed-layer Ce/N. Smectites replacing plagioclase pheno-
crysts contain 22 to 24% MgO (low K) and are presumably saponites. For both
types of samples, the A1,0, values are mostly in the range of 0.5 to 5%. The
temperature of formation was low but how low is not known. The Fe- and
K-enriched smectites apparently represent the initial response of the cooling basalt
to reaction with seawater and are likely to have been formed at a slightly elevated
temperature.
A study of weathered basalts (3.5 to 13 m.y.) from the west flank of the
Mid-Atlantic Ridge illustrates the effects of more intense weathering (Robinson et
al., 1977). Smectites fill fractures and vesicles and occur in alteration zones 10 to 30
times the width of the associated fracture. In general H,O, CO,, K,O and
Fe,O,/FeO ratios increase with increasing alteration. The smectites in veins and
vesicles often exhibit a distinct color and compositional zonation. Most commonly
the outer layer (nearest the basalt) is brown or green. This grades inward through
shades of yellow to bright golden orange or red layers. The red layer is often
followed by a light brown to colorless material in the core area. Table 6-4 gives the
compositional variations in one vesicle. Most of the iron is presumably in the ferric
form.
The smectites are the first mineral to form, followed by phdlipsite, carbonate or
pyrite. The first smectitic material to form, “1” in Table 6-4, is a typical Fe, K-rich
material, presumably the Ce/N observed in other areas. This material was deposited
355

Table 6-4
Compositional Variations of Smectites in a Zoned Vesicle (332B-44-2, 26-28 cm). (After Robinson et at.,
1977.)
1 2 3
SiO, 44.59 27.68 41.52
TiO, - tr -
A1203 4.40 3.09 12.66
FeO 27.65 45.98 5.33
MnO 0.01 0.13 0.13
MgO 8.03 5.83 25.96
CaO 1.20 1.18 1 .oo
Na20 0.09 0.06 0.09
K2 0 3.92 0.19 0.02
Total 89.89 84.14 86.71
Note: FeO = total iron; tr = trace; - = not detected; 1= greenish-yellow smectite lining vesicle wall;
2 = red smectite rimming layer of yellow smectite; 3 = light brown smectite filling center of vesicle.

early under nonoxidative conditions and perhaps when the temperature was elevated.
When seawater entered the fractures oxidative conditions prevailed and the smec-
tite, apparently nontronite, crystallized in the red layer. The final smectite to form
was a saponite with a relatively high A1 content. Some of the Mg could have been
obtained from seawater, but the high A1 content suggests much of the A1 and Mg
was derived from the basalt. Though one basalt in this study had an age of 3.5 m.y.
and the other -
13 m.y., no variations in alteration with age were observed.
In the basalts penetrated in the Nazca plate, alteration products occur most
commonly in veins and fractures (1-10 mm) throughout the 11 m of basement
penetrated (Seyfried et al., 1976). The veins are composed primarily of green to
blue-green smectite in association with minor pyrite, calcite and celadonite-
glauconite. The smectite has a relatively high content of both Fe and Mg and is
probably a mixture of nontronite and saponite. A1 and K are both relatively low.

Of more interest is the presence of well-developed celadonite-glauconite (Fig.


6-3). Chemical data is lacking but the x-ray patterns indicate the physil is an Fe-rich
10 A phase with only a few expanded layers. The x-ray pattern resembles that of a
well-crystallized glauconite. Oxygen isotope analyses indicate the physils formed
under low temperature conditions.
Two adjacent basaltic cores from the southern edge of the Bermuda Rise have
been studied in considerable detail to determine the effects of age on alteration. The
basalts are 110 m.y. old and are primarily pillow basalts. The physils are present in
veins, fractures and vesicles, as in the younger basalts. The basalts of Hole 417D
have been slightly altered, whereas those in Hole 417A, located on a basement high,
are strongly altered. The latter basalt was exposed to seawater for a longer period of
time than the former. Rb-Sr dates of the vein physils are the same as the age of the
basalts within the analytical uncertainties(Hart and Staudigel, 1979). The physils
356

I I 1
30 28 26 24 22 20 18 16 14 12 10 8 6 4 2
Degrees 2 8 CuKa radiation
Fig. 6-3. X-ray diffraction pattern of marine hydrothermal celadonite-glauconite. CuKa radiation.
Samples were oriented on microporous silver filters. From Seyfried et al., 1976.

formed relatively rapidly (few million years) following deposition of the basalt, and
little alteration has occurred since. Isotopic studies of calcite veins indicate a
temperature of formation of 14” to 4loC, which is appreciably higher than the
normal seawater temperature (Lawrence, 1979). Also, as the calcites form later than
the physil, these temperature values can be considered minimum values for the
physils.
The pillow margins exhibit a concentric zoning, with a hyaloclastic zone, a glassy
zone, a variolitic zone, a spherolitic zone, and the pillow core (Juteau et ul., 1979).
In the least weathered basalts, Hole 417D, fresh black glass is abundant; i t is partly
altered in the hyaloclastic zone, where it contains rims of brown palagonite, and in
the variolitic zone, where it contains fibropalagonite.
Palagonitization occurs when the hot basalt comes in contact with cooler seawater.
The glass is strongly hydrated ( - 15% H,O). Si, Al, Mg, Ca and Na are leached
from the glass and K, Fe and Ti are concentrated or enriched in the residue. This
reaction is somewhat variable.
X-ray patterns of the palagonite zones indicate the major physil is a dioctahedral
smectite (montmorillonite). K and Fe are enriched in the outer portion of the zones,
suggesting the presence of a “ protoceladonite” (mixed-layer?). The palagonite zones
in deeper samples contain saponite. Vesicles and veins are filled by dioctahedral
357

Fig. 6-4. Synthetic section of a pillow margin, DSDP Hole 417A, with typical concentric structure.
HZ = hyaloclastic zone; A = elongated palagonitic green fragments, coming from a poorly fractured
glasssy zone, in a calcitic matrix; B = angular green and brown palagonitic fragments, coming from a
highly fractured margin, in a complex fine-grained matrix (clays, iron oxides, zeolites, calcite). G Z = glassy
zone; A+poorly fractured, with palagonitic rims developing around the phenocryts; B = Mghly frac-
tured, with palagonitic concentric layers developing parallel to the fractures. VZ = variolitic zone.
SZ = spherulitic zone. PC = pillow core. Phenocrysts: PLA = analcitized plagioclase; OL idd. =
iddingsitized olivine; Cpx = fresh clinopyroxene. Veinlets, fractures, and segregation vesicles (SV) are
filled with green (GP) and brown smectites (BC), and calcite (Cc). From Juteau et al., 1979.
358

green smectite (inner position), trioctahedral brown smectite and calcite (center). On
the other hand, Scheidegger and Stakes (1979) report that the physils are primarily
celadonite and Fe-rich saponite. Their x-ray patterns confirm that some 10 A -
celadonite is present but some of the material appears to be a mixed-layer phase,
apparently some variety of celadonite/smectite. K,O values are as high as 6.4%.
In the more highly weathered basalt of Hole 417A (Fig. 6-4) the glass is
completely palagonitized. Olivine phenocrysts are replaced by iddingsite, calcite and
green (di) and brown (tri) smectite. Plagioclase is strongly altered but not to physils.
Vesicles and veinlets are filled with the same minerals as in Hole 417D, di- and
tri-smectite and calcite. Scheidegger and Stakes (1979) report that the physils are
more Al-rich than in Hole 417D. AI,O, values are on the order of 15 to 23%.
compared to 5 to 10% for Hole 417D. They identify the physils as montmorillonite,
" proto-celadonite" and saponite.

Rusinov et al. (1979), in a study of Hole 417A, found that the phases identified
optically as palagonite contain trioctahedral smectites (060 1.53 -
and varyingA)
amounts of amorphous material. Light colored smectites formed metasomatically

1.530

1.514

1.534

5 n 1.514

Fig. 6-5. Portions of diffractograms of smectite showing the reflection of (060). 1 and 2 = palagonite: 3.
4, and 6 = green smectite from altered glass; 5 = white smectite from the vesicle. From Rusinav et al..
1979.
359

1 b-1

2’
eg

h 4.51 / I .

& J -7J eg
.-. 12.1

2
3.33 3.80 4.234,51
16-17

4 eq

I , , I , , l , l I I I I I I I I 1 I I l l 1 1 I I I 1
30 25 20 15 10 5’28

Fig. 6-6. X-ray diffractograms of montmorillonites (and 1,’s) from basalt Sample 417A-30-4, 122 cm.
1 = white smectite in a large cavity; 2 = white smectite in a calcite vesicle; 3 = green smectite replacing
glass; 4 = white smectite in an amygdule; 5 = brown smectite with iron hydroxides, e.g., = ethylene
glycol. From Rusinav et al.. 1979.
360

Table 6-5
Microprobe Analyses of Smectite and Celadonite from DSDP Hole 417A (From Rusinov et at., 1979.)
1 2 3 4 5 6 I 8
SiO, 53.28 51.63 52.51 43.49 39.02 53.3 49.06 50.37
TiO, 0.0 0.06 0.0 0.88 0.0 0.0 0.04 0.02
A1 2 0 3 26.30 4.27 0.03 4.93 5.57 2.6 3.59 5.18
FeO 3.02 - - - -
Fe203 - 13.06 10.18 27.62 28.70 23.34 25.69 24.74
MnO 0.06 0.04 0.01 0.01 0.0 0.0 0.01 0.01
MgO 5.38 20.6 24.64 10.64 10.32 5.1 5.06 5.59
CaO 0.48 0.82 0.55 1.80 1.24 0.45 1.22 1.62
Na ,O 0.08 0.10 0.06 0.0 0.05 0.0 0.0 0.05
K2O 0.59 0.14 0.12 2.36 1.60 7.4 5.82 4.09
Total 89.19 90.2 87.7 90.75 85.35 91.25 89.46 90.71

Tetrahedral
Si 3.56 3.71 3.75 3.45 3.11 3.90 3.69 3.68
Al 0.44 0.29 - 0.46 0.52 0.10 0.31 0.32
Fe - - 0.25 0.09 0.37 - - -

Octahedral
Al 1.63 0.07 - - - 0.12 0.01 0.13
Fe 0.17 0.75 0.30 1.67 1.35 1.29 1.46 1.36
Mg 0.54 2.20 2.62 1.26 1.23 0.56 0.57 0.61
Interlayer
Ca 0.04 0.06 0.04 0.15 0.11 0.03 0.09 0.1 3
Na 0.01 0.02 0.01 - 0.01 - - 0.01
K 0.05 0.01 0.01 0.23 0.17 0.69 0.56 0.38
Note: 1 = pale green repacing pagioclase; 2 = dark brown, from altered glass in basalt groundmass;
3 = brown, from glass in basalt; 4 = yellow, from a vesicle associated with celadonite and Fe hydroxide;
5 = green; 6 = green, altered glass in basalt groundmass; 7 = yellowish green, margin of vesicle; 8 =
reddish brown, altered glass at pillow margin.

after plagioclase and filling vesicles and cracks are close to dioctahedral (060 =
1.514 A). Mixtures of the two occur in some samples. Fig. 6-5 shows the nature of
the 060 reflections. Fig. 6-6 shows typical x-ray patterns obtained from a relatively
small area of one core section. Table 6-5 lists chemical analyses (microprobe) of the
various physils. Area one in Fig.6-6 contains a relatively pure smectite. The other
three areas contain mixed-layered 10 A/smectite material containing 50 to 60%
10 A layers. Rusinov et al. (1979) described three types of smectite-mixed-layer
physils. The colorless to light green variety (Table 6-5, Column 1) is an A1 smectite
in the range of montmorillonite to beidellite. The high number of octahedral cations
suggest some of the Mg is present in the interlayer position. These physils can have
as much as 4% K,O and are presumably I/S physils. Al-rich (12 to 15% AI,O,, 2 to
3% KzO) mixed-layer physils are also abundant in basalts from St. Paul's Rocks
(0'56" latitude, 29'22'W longitude) (Melson and Thompson, 1973). This is of
interest as it indicates that under some conditions illite-like layers can form rather
361

than celadonite layers. Green and light brown smectites (Table 6-5, Columns 2 and
3) commonly consist of Mg-rich trioctahedral smectite. Many of these physils can be
classed as Fe-rich saponites. The brown variety have a high Fe content (Table 6-5,
Columns 4 and 5) and should be classed as nontronites; however, they contain
abundant Mg and are apparently trioctahedral and possibly should be classed as
Fe-rich saponites. It is quite likely that many of these analyses are of two or more
physils. The seventeen analyses indicate there is a wide range in the Fe/Mg ratio
and the ratio has relatively little relation to color.
Celadonite or “protoceladonite” (Table 6-5, Columns 6, 7, 8), present in a variety
of environments, is less abundant than smectite and decreases with depth. Some are
composed predominantly of 10 A layers but many contain an appreciable amount
of expanded layers. Because of the low content of octahedral Al, these physils are
presumably mixed-layer celadonite/nontronite.
It is apparent from this discussion that a variety of poorly defined physils are
present in these altered basalts and that associations are complex and interpretation
somewhat speculative. The most distinctive feature of the secondary physils in the
basalts in Hole 417A is the presence of A1 smectites (montmorillonite-beidellite).
The A1 smectites commonly, but not always, partially replace plagioclase, which is
presumably the source of the Al. The problem is whether the A1 smectites were
formed early, at a slightly elevated temperature (Rusinov et al., 1979), or formed
late after prolonged exposure to seawater (Scheidegger and Stakes, 1979). Based on
a review of other sites the latter authors concluded that physils in young crust ( < 15
m.y.) are Mg-rich; those in crust of intermediate age ( - 15 to 50 m.y.) are
dominantly enriched in Fe and Mg; and those in older crust have high contents of
A1 and K, but the variation is due to the length of time they were exposed to
seawater rather than due to the age of the basalt.
An answer to the origin of the A1 smectite might be obtained if it were
determined how much K is tied up in the A1 smectites (I/S). Analyses of whole
samples show a systematic increase in K 2 0 from 0.1% in the basal basalt of Hole
417A to approximately 6% at the top of the basalt (Donnelly et al., 1979). In
contrast to other areas, Mg as well as Ca is depleted upward. Al, Si, and Fe values
remain essentially constant, though the Fe3+/Fe2+ratio increases upward. The K is
believed to have been extracted from seawater. As the isotopic data indicate a
slightly warmer temperature than normal seawater, it was concluded the alteration
process was caused by upward-moving seawater, which passed downward into the
crust over an unknown lateral extent. Calculations, based on the K content of
seawater, indicate the amount of water (30°C) that moved through the basalt was
700 to 1600 volumes of the original void space. It seems likely that intensive
leaching could eventually make sufficient A1 available to allow A1 smectites and I/S
to form.
Not only does the composition of the physils vary locally within vesicles, but
some of the regional samples have physil suites that differ in their Fe, Mg, A1 and K
content. As the composition of the basalts is relatively uniform, the differences in
physil chemistry apparently depend primarily on the nature of the fluid phase and
the volume available (largely time-dependent). Isotopic measurements of low-tem-
362

perature weathering products indicate that most of the minerals formed at tempera-
tures < 50°C but hgher than the normal ocean water.
The secondary physils apparently obtained most of their Mg and K from
seawater and Fe and A1 from the basalt. Si was obtained from the basalt or biota.
Seawater is by far the major source of water. The hot basalt may come in direct
contact with cold seawater or seawater may descend into the basalt, be heated to
several hundred degrees (be depleted in Mg and sulfate and enriched in K and H),
and move laterally for a considerable distance before returning to the ocean.
Superimposed on all this is the oxidative state of the environment. Thus, it is easy to
see why there is so much variation in composition. Even though environmental
conditions vary widely, some variety of smectite is the earliest formed phase.
Apparently with time some of this is converted to celadonite and I/S.
In a study of submarine weathered basalts (Leg 34, eastern South Pacific) Bass
(1976) identified four physil suites that he believed were formed by four sets of
processes:
(1) late magmatic-deuteric alteration: biotite, chlorite (paper thin veins), iddingsite,
minor talc and possibly smectite.
(2) seawater alteration (preburial): palagonite, green to blue-green smectite.
(3) nonoxidative diagenesis (postburial, limited oxidation): well crystallized smec-
tite, Fe-rich saponite, talc, chlorite, celadonite and Ch/S (altered deuteric
chlorite). Saponite = X0.33(Feo3.:*Fe~.:,Mg,.os )(Si,,oAl0.s,Fe03.0+3)0,o(OH)2.
(4) oxidative diagenesis (postburial, intense oxidation): celadonite and
celadonite/smectite formed; smectite and chlorite oxidized and eventually de-
stroyed.
There has always been some dispute as to where to place the compositional
boundaries defining celadonite and glauconite. Celadonite is considered to be the
tetrasilicic end member 2:l physil. Though some dioctahedral Fe micas exist with
no tetrahedral Al, they are rare; the Clay Minerals Society Nomenclature Commit-
tee (Bailey et al., 1979) has placed an upper limit of 0.2 Al per four tetrahedral
positions. Mg provides an additional criterion. Most celadonites (from altered
basalt) have more than 0.5 of the octahedral positions filled with Mg. Glauconites
(round green pellets) typically contain less Mg than this (Weaver and Pollard, 1973).
An added criterion is that the (060) value is < 1.510 A for celadonite and > 1.510 A
for glauconite. These values basically reflect chemical differences. An increase in
(060) is usually related to an increase in Fe3+; however, it is unlikely that Fe3+
alone controls the (060) dimension. Most of the layer charge in celadonite originates
in the octahedral layer. This charge is carried by the apex oxygens and they
mutually repel each other. This leads to an increase in the thickness of the
octahedral layer (2.48 A vs. 2.12 A for muscovite) (Zvyagin, 1957) and, presumably,
a thinning in the b axis direction, thus a higher octahedral charge should produce a
smaller (060).
Few of the “celadonites” described in the various Deep Sea Drilling Reports
(DSDP) meet all the criteria for celadonite; however, Buckley et al. (1978) described
four samples from ocean-floor basalts which are excellent celadonites. We can
conclude, therefore, that celadonites do form by the marine weathering of basalts.
363

Many of the “celadonites” described in the DSDR have more than 0.2 tetrahedral
Al, commonly 0.3 to 0.5, and are, at least, glauconitic. Most of these samples have
less than 6 to 7% K,O and are mixed-layer physils. The tetrahedral A1 is possibly
present in the expandable layers, but if all, or most, of the tetrahedral A1 were
assigned to the expanded layers (10 to 30%) the values would be much too high.
Either some glauconitic material is present in the basalts or the samples contain
more than one phase.
There is no obvious reason both celadonite and glauconite cannot form from
marine basalts. Glauconites formed at normal seawater temperatures, whereas most
continental celadonites formed hydrothermally. It is reasonable to expect that
marine celadonites formed at slightly elevated temperatures. It is possible the two
minerals, or two compositions, could be used as temperature indicators.
A final point: It is worth noting the relative abundance of K in the secondary
smectitic physils formed from the ocean-floor basalts. The great majority of these
physils are mixed-layered with a micaceous component. Though micaceous layers
are common, illite apparently does not form. Many of these 10 A layers are Fe-Mg
micas, but the high A1 content of some of these mixed-layer phases indicates some
illite layers are present. The problem is whether the illite layers formed hydrother-
mally or at the near zero bottom water temperatures.

High to Low Temperature Transition

DSDP Site 504, south of the Costa Rica Rift in the eastern Pacific, penetrated
274.5 m of sediment and 1075.5 m of basalt (Fig. 6-7). This is the first hole to
recover in situ basalts containing greenschist facies rocks and to show the transition
from low-temperature to hgh-temperature physils. The physils occur both filling
cracks and in bulk rocks, where they fill void space and replace igneous minerals
(Alt et al., 1983).
In the upper alteration zone (274.5-584.5 m), which was relatively open to
seawater, the primary physils are celadonite-nontronite and saponite, accompanied
by Fe-hydroxides. Chemical changes involved oxidation, hydration, gain of K, and
loss of Si. Anoxic conditions existed in the remaining portion of the basalt. Saponite
and smectite with a few chlorite layers are the primary physils between 584.5 and
836 pm. Olivine and glassy pillow rims are completely replaced by physils; plagioc-
lase is partially replaced; clay veins are present. The altered basalts have lost Ca and
gained Mg. Temperatures of alteration throughout the pillow section were probably
less than about 100°C. The physils are similar in the 836-898 m interval, but
temperatures were probably higher than 100°C.
Greenschist facies minerals (chlorite, epidote, actinolite) appear at 898 m and
persist to total depth (1350 m). Minor talc is present, replacing olivine, but chloritic
material is by far the predominant physil. Some discrete chlorite is present, but most
of the material is Ch/S, Ch/V and Ch/S/V. The authors empirically divided the
chlorite/ expandable layer physils on the basis of the relative proportion of chlorite
layers present (5 = fewest chlorite layers and 1 = nearly pure chlorite). Regular
364

274 5
3w
I I
I

I I
400

I
I
500
Gv
GI
AA-me

600 ! !
Gv
I I AA
I
I
I
7w
AA

'I 1'1
RML
Gn. Cp, SI

I
RML

I
I
I I I Il 'l
I
CP 51
I
I RML
I RML
RML

I 1 I
I

4
Fig. 6-7. Distribution of secondary minerals with depth in Hole 504B.+ includes analcite, stilhite,
thompsonite, and natrolite. + + Gy = gyrolite, AA = aegerine augite. Me = meanlite, RML = regular-
mixed-layer chlorite-smectite, G n = galena, Cp = chalcopyrite, SI = sphalerite. * mixtures range from
chlorite-rich (Type 1) to expandable layer-rich (Type 5 ) ; see text. * * mixtures range from pure smectite
(a) to pure vermiculite (e); see text. From Alt et al., 1985.
365

mixed-layer (RML) Ch/S or corrensite is present throughout the interval. Based on


isotopic data and mineral associations the chloritic material is believed to have
formed over a temperature range of 200-250°C (minimum) but possibly as high as
380°C.
For most of the chloritic material Mg is slightly more abundant than Fe. The
+
Fe/Fe Mg ratios range from 0.25 to 0.53. The mo1% of MgO decreases with
depth. There is a great deal of heterogeneity in the composition of the chloritic
material.
It is somewhat surprising that the chloritic materials contain so many expandable
layers in the “greenschist” facies (low-grade metamorphism). This is presumably, in
some way, related to the hydrothermal origin and continued availability of water.
Basalts with a low water content, when subjected to regional metamorphism,
generally produce chlorite with few, if any, expanded layers. However, vermiculitic
material has been reported from low-grade metamorphic rocks (see Chapter VII).

Discharge Deposits

In many parts of the ridge system geothermal brines are discharged into the
cooler seawater, and physils and other minerals are precipitated from solution. The
Red Sea and the Galapagos are two areas that have been studied in detail.
In the Red Sea hot brines are discharged into localized depressions along the
central rift zone. The brines contain approximately 60 ppm dissolved SiO, and 80
ppm Fe”. The incoming temperature is as high as 250°C. After mixing with
seawater the brine pool has a temperature of about 60°C. When the temperature
cools, dissolved silica is supersaturated, a portion of the Fe2+ is oxidized and
smectite precipitates.
Material precipitated from the brine forms beds 20 m thick. The upper 5 m.
forming at the present time, contains a dark brown “soupy” mud (pH = 6.0) in
which 75% of the solids are smectite; the rest are Fe hydroxide and detrital pelagic
carbonate (Bischoff, 1972). Oxygen isotope studies indicate the smectite formed
over a temperature range of 80 to 140°C (Cole and Shaw, 1983). A typical structural
formula for the smectite is:
( ~ ~ ~ . ~ ~ ~ ~ ~ . ~ ~ ~ g ~ . ~ ~ ~ ~ ~ , ~ ~ ~ ~ o . ~ o C u ~ . ~ 6 ) ( S i ~ . ~

o10 (0H)2Ca0.045Na0.50K0.04

The octahedral cations total 2.34. Bishoff (1972) has included Mn, Zn and Cu in
the octahedral layer. If these metals are excluded, the total octahedral occupancy is
1.99. The same situation applies to the other three analyses reported by Bishoff.
This suggests that these metals may not be present in the octahedral sheet. The
calculated layer charge for the Bishoff formula is 0.53, which compares fairly well
with the total Ca, Na, K charge of 0.63. If Mn, Zn and Cu are assigned to the
interlayer position the calculated layer charge is 1.23 and the charge of the sum of
interlayer cations is 1.33. When the ferrous Fe is oxidized the layer charge will
366

decrease to 0.87, a more reasonable value. The smectite only expands to 15.5 A
when treated with ethylene glycol, which would indicate it has a high charge or
contains some interlayer material that prevents full expansion. The relatively h g h
060 value of 1.53 A reflects the high Fe content of the tetrahedral layer. The physil
belongs to the nontronite family even though it differs somewhat from those formed
on the continents (Weaver and Pollard, 1973). The Red Sea nontronites commonly
have a fibrous morphology.
In the high temperature zone near the vent (160" to 200°C) the smectite is a
montmorillonite-beidellite type (low Fe content) with some interlayer hydroxide
material (Cole and Shaw, 1983). This indicates that A1 as well as Si and Fe was
present in the discharging brine.
Samples underlying ( - 400 to 500 cm) the surface gel-like layer are more dense
and have more micaceous tendencies (Butozova et al., 1979). The low b-value of
9.06 A suggests there is no Fe in the tetrahedral layer. The K,O content is 2.31%
and approximately 30% of the layers are contracted to 10 A. When the material was
treated with K,CO, additional layers contracted, producing an I/S with 80 to 90%
contracted layers. The authors suggest that the nontronite and amorphous material
in the surficial layer is unstable. With time the shallow burial conditions become
more reducing and the pH values are lowered to 5.5, resulting in the formation of a
more-stable glauconitic or celadonitic phase. The physil is presumably a mixed-layer
glauconite/ nontronite or celadonite/nontronite.
The Galapagos hydrothermal mounds are located 18 to 32 km south of the
Galapagos rift axis and are aligned in rows above near-vertical basement faults
(Lonsdale, 1977). The mounds consist of various mixtures of pelagic oozes and
hydrothermal sediments. The smectitic physils occur as green to greenish-black
semiconsolidated angular aggregates ( < 1 to 20 mm) (Fig. 6-8) and replacing
calcareous and siliceous microfossils.
The smectitic material has a range of compositions and has been referred to as
nontronite and Fe-rich montmorillonite (McMurtry et al., 1983). Fig. 6-9 illustrates
the compositional range of a number of marine hydrothermal smectites and con-
tinental (detrital) smectites. There is essentially a continuous range of Fe/Al values.
In most cases it has not been established whether the Fe-montmorillonites are
mixtures of authigenic nontronite (Fe) and detrital montmorillonite (Al) or whether
Al-rich volcanic detritus has been a source of A1 and allowed a single phase Al-rich
nontronite to form. Table 6-6 contains typical chemical analyses of authigenic
marine nontronites.
There are two thngs to note in particular. The calculated layer charge, for the
Galapagos samples, ranges from 0.43 to 0.77 per O,,(OH),. This compares to a
value of 0.30 to 0.40 for most montmorillonite. The K,O content ranges from 1.8 to
3.2%. As suggested by Rateev et al. (1980) and Donnelly (1980), many of these
physils are mixed-layer physils. Though some of the x-ray patterns indicate smectite,
consisting largely of expandable layers, is present, others suggest some of the
material contains an appreciable number of 10 A layers. The high low-angle
background in the x-ray patterns in Fig. 6-10 indicates the sample contains
approximately 50% 10 A layers. The upper sample contains 3.9% K,O. The rela-
367

Fig. 6-8. Morphology of a brown smectitic granule from Galapagos spreading center. (1) General aspect.
(2) Detail of the constituent grains (up to 10 pm). (3-4) Detail of constituent grains; free-growing
smectites. TEM pictures indicate smectite crystals have a lath shape. From Hoffer et al., 1980.

tively low Mg content indicates the 10 A layers are more likely to be glauconitic
than celadonitic. Thus, the physils are apparently mixed-layer glauconite/ non-
tronite. In other analyses the Mg is high enough that the 10 A phase could be
classified as a celadonite.
Oxygen isotopic analyses (McMurtry et al., 1983) of Galapagos nontronite and
Fe-montmorillonite (Gl/N) suggest both minerals formed in the range of 25" to
47°C.The present temperature in the mounds is less than 15°C.The smectites
apparently formed earlier, when temperatures were higher, or formed at spreading
centers and were transported to the mounds by bottom currents.
368

Fe203

A‘2°3 L
Fig. 6-9. A1 ,O,-Fe,O,-MgO variation diagram for deep-sea smectites. Symbol identification is as follows:
A = Mounds nontronites, < 0.2 pm CED. A = mounds nontronites. ave. of 20 bulk analyses (Schrader et
al., 1980.) v = Red Sea nontronite ave. (Bischoff, 1972). T = Loihi Seamount nontronite (Malahoff et al..
1982. A = Famous nontronite ave. (Hoffert et al., 1978). A = Bauer Basin nontronite (Dymond and
Eklund. 1978). 0 = East Pacific Rise and Bauer Basin Fe-mont. (McMurtry and Yeh, 1981.) 0 = Mounds
Fe-mont. < 0.2pmCED. 0 = OCP Ridge Fe-mont. (Rateev et al., 1980). 0 = NE Pacific Fe-mont (Acki
et al., 1974). 8 = Domes Fe-mont. (Hein et al., 1979). 0 = Detrital Al-mont./beidellite (Weaver and
Pollard, 1973). = Beidellite from Panama soil (Weaver and Pollard, 1973). From McMurtry et al.. 1983.
Reprinted with permission from Geochim. Cosmochim. Acta. Copyright 1983 Pergamon Journals. Ltd.

The nontronitic Fe-Si phase, which forms from hydrothermal fluids issuing from
the vents, is apparently metastable in cool seawater. With time K and Mg are
extracted from the seawater and nontronite is diagenetically altered to a mixed-layer
Gl/N or Ce/N, perhaps eventually to glauconite or celadonite. As most glauconites
contain expanded layers, much of the material in its present form could be classed
as a glauconite or glaucony. In areas where there is abundant downwelling of
seawater the glauconitic phase could form by direct precipitation.
As was discussed in the Pacific Ocean section, it has been suggested that the Fe
hydroxides formed at vents along the ridges can be transported considerable
distances by marine currents and react with biogenic Si to form Fe-rich montmoril-
lonite (Heath and Dymond, 1977; Aoki et a/., 1974). Oxygen isotopic data o f
Table 6-6
Comparative Chemistry of Authigenic Marine Nontronites. (After McMurtry er al., 1983.)
Galapagos Mounds Red Sea Bauer Famous Loihi
Dredge Leg54 Leg54 Leg54 Leg54 EDTA Basin Seamount
bulk ' bulk bulk < 1 p m 4 < 0.2 p m and
CBD 5 CBD
smectite
grains
<44pm
CBD
-
Na,O 1.54 2.05 2.40 0.15 0.16 2.71 0.60 1.82 0.2
K2O 1.78 3.24 3.00 2.70 2.43 0.78 2.04 3.03 0.6
CaO 0.74 0.30 0.80 0.08 0.78 0.43 0.47 1.68 0.7
MgO 2.44 3.70 4.40 2.95 3.13 1.26 6.10 2.79 2.3
A1203 0.18 0.35 0.34 1.57 1.06 2.45 2.44 0.24 2.4
SiO, 47.06 50.78 52.20 48.09 52.80 36.60 52.25 39.68 50.2
Fe203 36.36 29.52 27.50 30.26 27.36 31.62 29.70 35.90 31.9
MnO 0.31 0.11 0.10 - 0.32 0.45 0.64 3.74 0.02
TiO, - 0.03 0.02 0.20 0.10 0.03 0.02 0.3
Cr203 0.00 0.00 0.00 - 0.03 - - 0.00 0.03
L.O.I. 9.59 * 9.94 8.10 14.07 11.83 * 19.30 5.73 * 7.95 11.35 *
Total 100.00 100.02 98.86 100.07 100.00 95.60 100.00 96.85 100.00
Tetrahedral
si4+ 7.33 7.68 7.77 7.57 7.96 6.76 7.45 6.56 7.58
~ 1 3 + 0.03 0.06 0.06 0.29 0.04 0.53 0.41 0.05 0.42
Fe3+ 0.64 0.26 0.17 0.14 0.71 0.14 1.39
8.00 8.00 8.00 8.00 8.00 8.00 8.00 8.00 8.00
Octahedral
Fe'+ 3.57 3.10 2.91 3.44 3.10 3.69 3.04 3.07 3.62
AP+ 0.15
Ti4+ 0.02 0.01 0.03
Mg2+ 0.43 0.83 0.98 0.54 0.70 0.31 0.96 0.69 0.35
4.00 3.93 3.89 4.00 3.96 4.00 4.00 3.76 4.00
In terlayer
Ca2+ 0.12 0.05 0.13 0.01 0.13 0.09 0.07 0.30 0.11
Mg2+ 0.13 0.15 0.04 0.34 0.17
Mn2+ 0.04 0.01 0.01 0.04 0.07 0.08 0.52
Na+ 0.14 0.60 0.69 0.05 0.05 0.97 0.17 0.58 0.04
K+ 0.35 0.62 0.57 0.54 0.47 0.18 0.37 0.64 0.12
0.78 1.28 1.40 0.75 0.69 1.35 1.03 2.04 0.44
Layer Charge 1.10 1.36 1.54 0.95 0.85 1.55 1.51 2.85 0.74
Interlayer 1.07 1.34 1.54 0.91 0.86 1.55 1.52 2.86 0.72
Notes:
' Average of N-1, N-2 and N-3 (Corliss er al.. 1978).
Average of 5 samples of green clay-rich material (Hekinian ef ul.. 1978)
Average of 20 samples of green clays (Schrader er al.. 1980).
Average of 3 samples from < 1 um fraction (Rateev er ul., 1980).
' Average of 6 samples from McMurtry et al., 1983.
' Average of 4 samples chemically treated to remove carbonate and iron oxide (Bischoff, 1972).
' Average of 10 smectite grains from coarse fraction ( > 62 p m ) (Dymond and Eklund. 1978).
' Average of 5 dark green and yellowish green samples of CYP74-26-15 and 16 (Hoffert et al., 1978).
' From Malahoff et al. (1982).
* Calculated by difference from ideal total of 100.00%.These values are for comparative purposes only.
370

L
'20A

1 S m i l e 424.2-3. 67-69 cm

I
10 10 14 10 17A
HEATED UNTREATED GLYCOL-TREATED

Fig. 6-10. X-ray diffraction charts of the clay fraction of Galagapos hydrothermal deposits. (a)
Non-saturated samples; (b) Mg-saturated sample. High background above 17 A suggest physil is I/S
with approximately 50% smectite. From Hoffer et al., 1980.

Fe-rich smectite samples from the East Pacific Rise and the adjacent Bauer Basin
indicate they were formed by hydrothermal processes at temperatures of 30" to
5OOC. Thus, though these smectites may occur a long way from their source, it is
371

apparent that it was the smectite and not the Fe hydroxide precursor that was
transported (McMurtry and Yeh, 1981).

Hydrothermally Altered Sediments


In addition to the basalts being altered and metamorphosed, one might expect
that under some conditions the heat from the basalt would affect the adjacent
sediment. This has been observed in only one area: the Guaymas Basin Rift, Gulf of
California (Kelts, 1982; Kastner, 1982). The sediments in this area are diatomaceous
silty clays; the detrital physils consist of I/S, illite, chlorite and kaolinite. Dolerite
sills (20 to 30 m thick) have intruded these sediments at depths ranging from 33 to
350 m below the sediment-water contact. The contact metamorphic zone ranges
from 2 to 50 m thick. In these zones opal-A is converted to quartz, well crystallized
smectite is formed, and the detrital physils are largely destroyed (also calcite
recrystallizes and dolomite forms). Illite and, to a lesser extent, chlorite are uncom-
mon but are present adjacent to some of the sills. Oxygen isotopic data (calcite)
indicates temperatures in the range of 140" to 170°C. It is not clear why an I/S
phase did not form.
In addition to the sills, there is apparently a magma source a few hundred meters
below the bottom (300 to 400 m) of the drill holes. Chlorite increases with depth,
and near the base of the hole the rocks have a characteristic greenschist-facies
assemblage of quartz-albite, chlorite and epidote. The alteration temperature was
300 k 50°C. The Mg/Fe ratios of the chlorite range from 1.0 to 1.9.
The heat from the sills caused pore water and dissolved material to be expulsed
from the sediment. This material moved to the surface along faults where it
precipitated (hydrothermal) large amounts of talc and minor amounts of Fe and Mn
compounds (Einsele el al., 1980). Such a mechanism may account for the scattered
deposits of talc and sepiolite in the ridge areas. The formation of talc rather than Fe
smectite is presumably because the hydrothermal fluids had relatively little contact
with the hot basalt and derive most of their ions from seawater and alteration of the
sediments.
Geochemists have long been concerned about the global budget of Mg. Calcula-
tions indicate considerably more dissolved Mg is being transported to the ocean by
rivers than is being removed from the ocean. So, the search for Mg sinks (reviewed
by Honnorez, 198l)! The Mg deposited in pelagic clays per year is only one fifth of
the dissolved Mg flux from rivers. About half the river flux (1.3 X l O I 4 g) is
incorporated in the seafloor sediments (ion exchange, Mg replacing Fe in physils,
formation of authigenic minerals--glauconite, sepiolite, palygorskite, dolomite). Only
about 0.1% of the upper few meters of new igneous oceanic crust undergoes a
low-temperature ( < 100°C) reaction over a period of tens of millions of years and is
a minor sink. Actually some Mg may be added to the seawater (Thompson, 1983).
However, calculations for the flux in the upper 500 m of oceanic crust, where
temperatures are higher (up to amphibolite facies), show Mg uptake that amounts to
approximately half the river flux. It seems likely that this is the long-sought Mg sink
(Honnorez, 1981).
312

Ambient Marine Smectite

The preceding material indicates that most of the physils formed from ridge
volcanics formed at slightly elevated temperatures (hydrothermal) and crystallized
shortly after the hot volcanics solidified. We must now consider the alteration of
cool volcanic material by normal seawater.
Bonatti (1967) divided submarine lavas into two distinct types. In one type the
lava flows quietly on the sea floor. Interaction with seawater causes the instanta-
neous formation of a thin insulating crust of glass at the surface of the flow. This
allows the lava to cool without contact with the seawater. Any alteration is
extremely slow. The second type (hyaloclastic) interacts extensively, both physically
and chemically, with the seawater and is shattered and pulverized. The resulting
lava debris (hyaloclastites) is rapidly hydrated at high temperatures, forming mainly
palagonite. Both smectite and zeolite can form at elevated temperatures (hydrother-
mal) during the palagonitization process. Presumably palagonite continues to alter
to smectite and zeolite after cooling. The two different responses of the lava to
water are due to differences in viscosity (high viscosity produces hyaloclastites)
whch in turn is determined by the temperature of the lava when it is ejected on the
ocean floor. Composition does not appear to be a factor.
One of the most distinctive features of palagonite is its high water content, whch
can be more than 20%.Deep sea palagonite forms mainly at high temperatures.
(Lava comes in contact with seawater at temperatures near 1000°C.) The rate of
hydration decreases exponentially with decreasing temperature and little or none
appears to form at normal seawater temperature ( - 0°C) (Bonatti, 1967). Fine
volcanic fragments with a large surface area may be completely altered to palago-
nite, containing varying amounts of smectite and zeolite, mainly phillipsite.
Heulandite and clinoptilolite are apparently the second most abundant zeolites
associated with altered volcanics; they are particularly abundant in older deposits.
These two zeolites form under conditions where excess silica in solution is present
and phllipsite forms in a Si-deficient environment (Elderfield, 1976).
Bonatti (1967) believes that alteration of fresh volcanic material under normal
-
marine conditions (0" and pH 7) is extremely slow and most of the physils and
zeolites form from palagonite developed at high temperatures. In a study of a young
volcanic island, Surtsey, off the coast of Ireland, Jakobsson and Moore (1986) found
that palagonization of the volcanic glass (sideromelane) is strongly temperature
dependent, the rate doubling for every 12°C increase. Less than 5% of the glass is
altered to palagonite at 40°C and all of it is altered to palagonite at 120°C. The
plagonite is in part composed of smectite, probably nontronite.
On the other hand, Hekinian and Hoffert (1975) believe that the degree and type
of palagonitization is a function of age. They calculated the rate of alteration based
on the thickness of the Mn crust, which occurs on top of the palagonite crust. The
Mn accumulation rate was assumed to be 3 p/103 years. The palagonite crust on
basalts from the inner floor of the k f t Valley, Atlantic Ocean, are 6 to 350 pm
thick and have an age of less than 12,000 years. Four to six kilometers away from
the rift axis the crusts are commonly 1000 to 3000 pm thick and have a calculated
373

age of 134,000 to 482,000 years. The palagonite crusts contain a minor amount of
poorly crystallized smectite (and chlorite, according to the authors). If these ages are
real they indicate the smectites form at an extremely slow rate.
Earlier, Moore (1966) showed young ( < 200 yrs) pillow basalt contained palago-
nite rims 2 to 15 pm thick, whereas older (lo4 to lo6 yrs) basalts (based on the
thickness of Mn rinds) had rims 250 to 700 pm thick. He also concluded that
palagonitization of submarine basaltic glass proceeds faster than that in fresh water,
and much faster than hydration of obsidian in a subaerial environment.
The alteration product of basaltic glass is commonly called palagonite. Regard-
less of its origin, it is apparently the precursor of most of the smectite formed from
marine volcanics. The interface between fresh glass and palagonite is usually distinct
and riddled with microcracks entering the glass from the boundary. In thin section
palagonite appears as a slightly yellow to deep yellow-brown, isotropic to slightly
anisotropic amorphous material; dark yellow to brown birefringent materials are
called fibropalagonite.
Palagonitization is nearly isovolumetric. The amount of glass converted to
palagonite equals the amount of newly-formed authigenic minerals. Staudigel and
Hart (1983) evaluated the chemical analyses of palagonite and reached a number of
conclusions. Of the external ions in the altered rim, H shows the deepest penetra-
tion. This indicates H,O is a necessary reagent which initializes the alteration of
obsidian. The most obvious chemical changes are an extreme increase in K and a
drastic loss in Ca and Mn. Though the composition of palagonites are quite
variable, they were able to establish common trends.
Microprobe analyses indicate all palagonites are enriched in FeO (total Fe) and
TiO, compared to the adjacent glass. Authigenic minerals are Ti free and there

Table 6-7
Average Gains or Losses of a “Normal” Mid Ocean Basalt Glass (Melson et al., 1979) During Alteration
to Palagonite. (After Staudigel and Hart, 1983.)
Normal MORB Average% g/100 g- Calculated Palagonite A
gain or loss remaining dry nor- “wet” mole%
wtcg during Pala- wt% mole%
gonitization wt%
SiO, 50.53 52.52 - 50 25.27 51.71 41.88 30.41 - 22.11
Ti02 1.56 1.22 + 104.5% 1.56 3.19 2.59 1.41 +0.19
A1203 15.27 9.35 - 55 6.87 14.06 11.39 4.87 - 4.48
FeO 10.46 9.09 - 12 9.20 18.83 15.25 9.26 +0.17
MnO 0.17 0.15 - 93 0.01 0.02 0.02 0.01 -0.14
MgO 7.47 11.58 - 67 2.47 5.05 4.09 4.43 - 7.15
CaO 11.49 12.80 - 88 1.38 2.82 2.29 1.78 - 11.02
Na ,O 2.62 2.64 - 81 0.50 1.02 0.83 0.58 - 2.06
K2 0 0.16 0.11 + 900 1.60 3.22 2.65 1.23 + 1.12
p2°5 0.13 0.06 - 90 0.01 0.02 0.02 0.01 - 0.05

H2 0 0.14 0.49 +13,470% - - 19.00 46.01 +45.52


ax 100.00 100.00 48.87 100.00 100.00 100.00 + 0.29
314

modal abundance is proportional to the amount of palagonite formed from glass.


TiO, is considered to be conserved (residue) and can be used to calculate the
amount of loss of material during the alteration. Thus, if a palagonite shows a TiO,
increase of loo%, only 50% is left of the original material; if the glass contains 50%
SiO, and the palagonite 458, there would have been a net loss of 27.5% SO,,
largely replaced by water. Using this approach, Staudigel and Hart were able to
calculate the gain or loss of various oxides (Table 6-7).
The Ti content of palagonites is quite variable, from 0 to 220% increase over
glass values, and is a measure of maturity. The network forming cations in basaltic
glass, Si and Al, remain essentially constant. The absolute values of SiO, and Al ,03
are approximately 20% less than in the glass due to the dilution by addition of
water. The behavior of FeO is similar to TiO,, passive accumulation during
alteration. Approximately 80% of the Fe2+in fresh glass is oxidized to Fe3+ in the
palagonite. CaO and MnO are largely removed from the glass during alteration,
primarily during the early stages. A moderate amount of MgO is lost; the loss
appears to be continuous with increasing maturity. Na,O losses are moderate and
are particularly high during the early stages of alteration. K ,O increases drastically,
particularly in the very early stages of glass alteration. The rate of increase in K,O
decreases with maturity (increase in Ti). Staudigel and Hart suggest that the extent
of glass alteration can be measured by the K concentration, as well as the Ti
concentration. However, the various data do not demonstrate that K continues to be
adsorbed once the glass-palagonite has cooled to the temperature of seawater. If this
were the situation, the physils in altered older volcanics would be largely celadonites
or illites rather than smectites. It seems more likely that the physils forming from
palagonites might incorporate the K present in the palagonite but probably not
adsorb additional K from seawater.
The high adsorption of K explains why the “smectites” formed from palagonite
are commonly mixed-layer mica/smectite physils. The composition of many of the
smectites formed from palagonite indicates that, at a later stage, Mg is also
adsorbed from seawater. The major question is why the A1,0, content of the
“smectites” formed from palagonite is appreciably less than that in the palagonite.
Staudigel and Hart concluded that the alkali budget of the total oceanic crust is
dominated by the formation of palagonite and smectite in the upper 600 m of the
oceanic crust and that the crust is a sink, rather than a source, for seawater alkalis.
The oceanic crust at DSDP Site 418A (typical?) consists of 16% palagonite, 10%
smectite, 7% carbonates and 4% unaltered glass. The remaining rock consists of
pillows, breccia and massive flows. Most of the elements provided by ocean crust
glass alteration are redeposited in veins and vug fillings as smectites, carbonates and
zeolites.
There is a considerable difference of opinion as to whether smectites (also I/S
and celadonite) form on the sea floor at the temperature of ocean bottom waters
and, if it does form, does it form only from palagonite developed at elevated
temperatures or can it also form from fresh volcanic glass? Smectites are abundant
in the Cenozoic and Mesozoic sediments overlying the basal basalt layer (Layer 11).
Much of this material has apparently altered at low temperatures from volcanic
375

debris; however, most of this material ranges in age from 10 m.y. to 150 m.y. and
most of the alteration has occurred after burial and presumably over a long period
of time. I will present enough of the data to illustrate the problem.
Both fresh and altered volcanic ash and pumice occur in sediments of the ocean
floor. Bonatti (1965) suggested that the altered ash (palagonite) formed by interac-
tion of lava with seawater (at elevated temperatures). Such glass will alter relatively
rapidly to smectites and zeolites, among other things. The fresh glass formed during
subaerial eruptions hydrates and crystallizes at very low rates (laboratory experi-
ments). He also notes that acidic glass alters at a slower rate (20%)than basic glass;
however, the difference in alteration is minor when compared to the difference in
alteration rate between a palagonitic and a normal glass of the same composition.
Volcanic ash layers are abundant in the marine sediments off the west coast of
South America and Central America. The ash, andesitic to rhyolitic, was derived
from subaerial eruptions. In ash beds up to 300,000 years old and as much as 16 m
below the surface there is no evidence that the glass is altered (Bowles er al., 1973;
Ledbetter, 1985). Chamley and d’Argoud (1978) examined the data on volcanogenic
deposits in Mediterranean cores and found no evidence that volcanic glass was
altering to smectite. Fresh ash deposits of Eocene age have been reported in Site 336
(Talwani et al., 1979). At Site 285, 400 m of Middle Miocene sediments are
composed of 70 to 90% glass shards. The remaining material is primarily manop-
lankton. Site 286 contains 200 m of Middle Eocene sediments composed of - 80%
volcanic glass (silt and sand). The porosity of these sediments is approximately 60%
(Andrews et al., 1975). Donnelly and Melson (1973) reported volcanic glass, with no
trace of alteration, is widespread in the Early Miocene, Oligocene and Paleocene of
the central Caribbean. Ash beds of Late Cretaceous age have glass shards com-
pletely altered to smectite. Does it take 100 m.y. for volcanic glass fragments to alter
to smectite?
A study of a series (20 to 50) of ash beds from wells in the Bering Sea provides
some insight into the rate of alteration (Hein and School, 1978). They describe as
present: bentonite beds (> 75% smectite in < 2 pm fraction), smectite-rich beds
(significant volcanic input, 50 to 75% smectite) and unaltered ash beds. Unaltered
ash beds are the most common beds in the upper 300 m of sediment, Pleistocene to
Lower Pliocene (5 m.y.), but occur as deep 850 m, in Upper Miocene sediments.
The unaltered glass (andesitic or more silicic) consists predominantly of glass
shards. With increasing depth of burial, compaction causes fragmentation of glass
shards and the progressive replacement of glass by smectite (Fig. 6-11). Bentonite
layers occur in sediments as young as Late Pliocene ( - 2 m.y.), but most occur in
the Lower Pliocene and Upper Miocene deposits. The smectites are dioctahedral,
Al-rich and apparently montmorillonite. The formation of montmorillonite rather
than Fe-Mg smectites is presumably due to the relative acidic composition of the
glass.
Present-day temperatures at the bottom (600 to 900 m) of most holes is 40°C or
less; however, oxygen isotope measurements on a carbonate bed at 871 m (Site 189)
suggest a temperature of about 70°C. (This would be equivalent to a burial depth of
- 3000 m in the Gulf Coast.) This carbonate bed is underlain by an unaltered ash
376

. ,
I 15 14 12 Ib 9 7
Fig. 6-11. X-ray diffractograms of Mg-saturated and glycolated smectite. Samples d, e, and f show
various stages in early formation of smectite within ash beds. Sample a is typical detrital clay mineral site
with smectite (60% expandable layers), chlorite and illite; b represents nearly pure smectite from
bentonite bed (90% S); C is smectite from smectite-rich bed with - 75% smectite layers. S = smectite,
C = chlorite, K = kaolinite, I = illite, CI = clinoptilolite. From Hein and School, 1978, Geol. SOC.Amer.
Bull., 89,197-210.

bed. Though there is a general trend of increased alteration with depth, there are
numerous exceptions, and unaltered ash beds and bentonite beds occur within a few
feet of each other.
Hein and Scholl (1978) also describe the presence of a basal clay, or basal
bentonite, overlying an extensive basalt. Much of the basalt (Upper Cretaceous) and
overlying clay has been altered by low temperature hydrothermal fluids to relatively
pure smectite and calcite. The smectites are primarily saponite and nontronite, in
contrast to the overlying montmorillonite beds derived from more silicic ash. Similar
basal smectitic clay beds have been described from other localities. For example,
Kastner (1976) describes such a clay bed in the southern South Pacific overlying
altered Upper Cretaceous basalt. The clay is composed of authigenic smectite,
clinoptilolite and quartz and contains relatively unaltered volcanic glass.
377

t
15 20 25 30

Sulfate
Calcium and Magnesium

01.. rnM rnM


-4 -3 - 2 -1 0 1 0 2 4 6 8 1 0 0 0.1 0.2 0.3 0.4 0.5
0 " ' I P
n

'"9 Bw
rn 300

5001 "- Potassium Strontium

rnM

:i
0.0 0.2 0.4 0.6

rn

6 00
200

TH Ti/AI

I-
Magnesium

r
0 0 400 800

Si/Al
I-
Site 336
Fig. 6-12. Interstitial water and sediment chemistry of sediments overlying basalt (IV). DSDP site 336.
Elemental ratios are atomic ratios. From Gieskes and Lawrence, 1981. Reprinted with permission from
Geochim. Cosmochim Acta. Copyright 1981 Pergamon Journals, Ltd.
378

Why is there so much fresh glass in marine sediments? Composition is a factor.


Generally rhyolitic glass altered more slowly than basaltic glasses. Peterson and
Griffin (1964) showed that in the South Pacific the smectite/mica (detrital) ratio
was larger in areas where basaltic glass was present than in areas where the glass
was acidic and believed this was an indication of the difference in the rate of
alteration. It is also likely that submarine ashes are more apt to form palagonite
during the quenching stage than are ashes of subaerial origin. The latter are also
likely to be more acidic. Thus, much of the fine, fresh glass in marine sediments may
have a subaerial origin, whereas those altered to smectite are more likely to be from
submarine explosions.
This late-stage alteration of volcanic tuff deposits is of interest from the stand-
point of the origin of ancient bentonites and K-bentonites. We tend to envisage
volcanic ash beds altering to montmorillonite shortly after deposition and while they
retain some contact with the overlying water. It now appears that these ash beds,
particularly the more acid types, were probably buried to a considerable depth.
They altered after they no longer had direct contact with circulating ocean waters,
but presumably were exposed to circulating and expressed water. Most marine
bentonites are presumably diagenetic (or epigenic) in origin rather than authigenic
(or syngenetic).
The review of the marine alteration of basalt brings to light another problem.
Chlorite and Ch/S are common alteration products in ancient volcanic deposits, yet
these physils are not forming today except under relatively high temperature

14
14
175

145

I62
164
29
17

33

23

1 1 I L 1 I I I I

I I I , t I I l 1 5" 10"
2 e 5" lo"
Fig. 6-13. X-ray diffraction patterns of soil smectites form Japan after removal of free oxides. Compare
with marine smectites in Fig. 6-11. From Masui and Shoji, 1969. Copyright 1969 Israel Prog. Sci.
Translation.
379

conditions in the ridge areas. Does the presence of these physils indicate an elevated
temperature? In many cases this appears to be true.
The interstitial waters have been analyzed from a large number of deep sea cores
(DSDP). With some variation they show essentially the same trend (Fig. 6-12). Mg
decreases with depth, with values near zero immediately above the top of the basal
basalt, and Ca increases. K and S S ” 0 values show a systematic decrease with
depth. Ti/Al and Mg/Al ratios of the bulk solids tend to be relatively constant. For
review see Gieskes and Lawrence (1981) and McDuff (1981).
The gradients in Ca and Mg can be explained in terms of mass transfer, through
the sediments, between the underlying basalts and the overlying ocean. Alteration
reactions in the sediment column are not great enough to explain the Sl8O depletion;
therefore, extensive alterations must occur in the underlying basalt. It is well
established that Ca readily dissolves from basalt. Sinks are requried for Mg and K.
The main sink would appear to be smectitic physils with (10 A layers) formed by
alteration of the basal basalt and volcanic glass fragments in the lower part of the
sedimentary section.
The oxygen isotope mass balance calculations indicate that 10 to 20% of the
upper 1000 m of basalt must be altered to account for the H’!O depletion. The
fluxes of Ca, Mg and K are reltively small, and the changes in the bulk composition
of the sediments would be hard to detect. McDuff (1981) interprets the data to
indicate that, when the sediment cover is thin, the mode of tranportation of solutes
is convection circulation. As the cover thickens, transport is by diffusion through a n
unreactive sedimentary layer. Thus, most of the alteration is presumably due to
reactions taking place near the spreading axis while convective exchange of seawater
is still active.
A number of authors have attempted to distinguish between marine authigenic-
hydrothermal smectities and detrital smectites on the basis of their x-ray character-
istics. In a local area, this may work but, in general, the range of “crystallinity” and
the range of 10 interlayers are similar for marine and continental-formed smec-
titic material. Fig. 6-13 contains x-ray patterns of smectitic material from soils
formed on volcanic ash deposits of Japan (Masui and Shoji, 1969). Compare Fig.
6-13 with Fig. 6-11. When the amorphous Si and A1 were leached from the soil
samples, relatively sharp, low-background smectitie peaks were obtained.
There is another suggested method of forming smectite. Von Bennekon and Van
Der Gaast (1976) found small (0.02 pm) srnectite particles on the surface of living
diatom frustules. The frustules contained 1.5% A1,0,. They suggested, but could not
prove, that the smectite nucleated on the frustules.

Palygorskite and Sepiolite

Palygorskite, with minor sepiolite, is a common constituent in DSDP cores of


deep ocean sediments. It has been well documented that the chain physils form in
soils, alkaline lakes and perimarine environments, under arid to semi-arid condi-
tions. There is a considerable difference of opinion as to whether it forms in a
380

“normal marine” environment. Couture (1977), Kastner (1981) (and references


therein) believe it forms authigenically or diagenetically in deep sea sediments.
Weaver and Beck (1977), Chamley (1979), Singer (1979) and Callen (1984) believe
most of the deep sea chain physils are detrital. Both factions agree some deposits are
hydrothermal, but the relative volume of these deposits is debatable.
Hydrothermal occurrences will be discussed first. Relatively pure samples of
sepiolite and/or palygorskite have been collected from distinct tectonic features on
the ocean floor (Hathaway and Sachs, 1965; Bonatti and Joensuu, 1968; Bowles et
al., 1971). It is believed that these deposits are hydrothermal; Mg was obtained from
the seawater or hydrothermal fluids; Si was obtained from volcanics, smectites or
hydrothermal fluids. On the other hand, Fleischer (1972) found veins of pure
sepiolite filling fractures in Middle Miocene diatomite beds (opal-cristoballite)
immediately off the coast of southern California (Santa Cruz Basin). The sample
was from a scarp wall that had presumably been exposed to seawater. However,
Fleischer suggested the Mg was supplied by Mg-rich ground water. In any event
there is no suggestion of an elevated temperature.
In core samples from basement basalt (Mariana Fore-Arc Region, Sites 458, 459),
Natland and Mahoney (1981) found palygorskite filling veins and associated with
smectite, celadonite and Fe hydroxides. They concluded it was hydrothermal and
formed at moderately elevated temperatures. The atypical presence of abundant
palygorskite with the more common smectite and celadonite led them to suggest
both the pH and Eh values were higher than normal. Hydrothermal fluids from the
East Pacific Rise are normally quite acidic and have relatively low Eh values.
The palygorskite from Sites 458, 459 is not typical. It has a basal (110) reflection
of 11.0 A rather than the common value of 10.5-10.6 A. It has less A1 and more Fe
and Ca than brackish water palygorskites.
(Si 7.53A10.47 ) ( A ~ o Fe:.~,Mg2,,)02,(OH)2(OH2
.~I )4 4 H 20(Nao.51
Ko.12Ca,.ol
)

Sepiolite, Mg,(H,SiO,),(OH),, commonly has a composition similar to the


theoretical formula and is a Mg-hydroxysilicate. Natural sepiolites contain 22 to
25% MgO and have a SiO,/MgO ratio around 2.3. Palygorskite,
,
Mg, (A1,Fe),Si 8020(OH) . 8H ,O, though commonly referred to as a Mg-rich fibrous
physil similar to sepiolite, is, in fact, quite different. Well-crystallized ancient
palygorskites contain approximately 10% MgO and a similar ( - 9%) amount of
Al,O,. Deep sea palygorskites contain from 3 to 8.5% MgO, generally < 5%, and
have Al,O,/MgO ratios ranging from 1.8 to 6, with most values > 3 (Bowles et al.,
1971). These latter samples contain more Fe,O, than MgO. The MgO content is less
than that of many montmorillonites, particularly the Cheto-type (4.4 to 7.4 MgO)
(Weaver and Pollard, 1973). One thing these two minerals have in common is a high
silica content, and thermodynamic calculations indicate they should form in en-
vironments where amorphous silica is relatively abundant. Deposits commonly
contain opal-cristoballite.
“Sepiolite” is relatively easy to synthesize in the laboratory, but palygorskite has
not been formed experimentally. Wollast et al. (1968) added Na metasilicate to
seawater and precipitated what was apparently a poorly crystallized sepiolite
381

-
(pH 8). They stated that sepiolite is the only non-aluminous, cation-bearing
silicate that can be precipitated directly from seawater. In view of the availability of
amorphous silica (biogenic and volcanics), sepiolite should be abundant on the sea
floor.
A more recent study by Kent and Kastner (1985, and references therein) showed
that Mg2+ can be removed from suspensions containing amorphous silica at low
temperatures by adsorption and precipitation of a Mg-hydroxysilicate resembling
sepiolite. In a series of experiments Kent and Kastner (1985) showed that the extent
of Mg2+ adsorption onto amorphous SiO, decreased with increasing NaCl con-
centration due to displacement of Mg2+by Na+. They concluded that adsorption of
Mg2+ onto amorphous SiO, is an insignificant process in seawater and that the
principal reaction between Mg2+ and amorphous SiO, in marine sediments is
-
sepiolite precipitation, at pH 8. (It should be pointed out that they were unable to

1
18
8 Brucite kepiolite j
16

14

12

...........
...............
10 l ......J
lot

Interstitial sea water


6- volcanic sea water
(Deception Island)
Average river
and lake water
4-
Ground waters draining
ultramatic rocks
Saline lake water
2-
Volcanic sublimates
I -
-~
-6 -5 -4 -3 -2
log %iOp (a q )

Fig. 6-14. Activity diagram for the system MgO-SiO,. Modified from Wollast et al., 1968. From
Elderfield, 1976, Chem. Oceanog. 5. Copyright 1976 Academic Press.
382

identify sepiolite in their experiments.) The rate of sepiolite formation increases


with increasing Mg2+concentration, dissolved silica concentration, and pH. How-
ever, the solubility of sepiolite increases with decreasing temperature and increasing
pressure. Thus, on the sea bottom where temperatures are < 5"C, sepiolite is
unlikely to form. Kent and Kastner believe the precipitation of sepiolite is only
likely to occur in the sediment column where temperatures are greater than 10°C -
and it is most likely to occur in carbonate-containing siliceous sediments where
CaCO, dissolution helps to maintain pH values near 8.0. The solubility-temperature
relation explains the origin of hydrothermal sepiolite and the lack of sepiolite in
recent marine sediments. They note that there are only a few cases of "unequivocal
evidence" [my quotations] of authigenic marine sepiolite. The activity diagram for
the system MgO-SO2-H20 (Fig. 6-14) indicates various marine waters have suffi-
cient silica for the formation of sepiolite. The main requirement is the addition of
Mg2+, assuming the pH is near 8.
Also, note that talc, Mg3Si,0,,(OH)2, is the thermodynamically stable Mg-sili-
cate in water with silica concentrations in excess of 100 p m at temperatures of 25°C
and below (Drever, 1974), but "sepiolite" is the physil that precipitates under these
conditions.
It is evident from the above discussion that there is a problem with the authigenic
formation of sepiolite in the deep sea environment. The problem with palygorshte is
much more difficult and has not really been explored. In general its origin is lumped

PALYGORSKITES 0 m.y. LATE PLIOCENE - HOLOCENE

Fig. 6-15. Late Pliocene-Holocene palygorskite-sepiolite occurrences. Dots are generalized DSDP and
oceanic occurrences, diagonal shading is continental data. Prefix D indicates soil or calcrete. Cross-hatched
areas are soils superposed on sedimentary basins with palygorskite. Letters and numbers refer t o
references in Callen, 1984. From Calen, 1984. Copyright 1984 Elsevier Pub. Co.
383

PALYGORSKITES 60 m.y. MAASTRICHTIAN - EARLY EOCENE

Fig. 6-16. Late Cretaceous-Early Eocene palygorskite-sepiolite occurrences. Symbols as in Fig. 6-15.
From Callen, 1984.Copyright 1984 Elsevier Pub. Co.

with that of sepiolite. It seems reasonable that sepiolite can precipitate from the
proper Mg-Si solution but palygorskite, with 10 to 18% A1,0,, is a horse of a
different color.
Weaver and Beck (1977) suggested that palygorskite formed in brackish water
environments and the palygorskite in marine sediments is either detrital or formed
by the post-depositional circulation of brackish water. We still have not seen any
" unequivocal evidence" for the authigenic formation of palygorskite in a normal

marine environment. We may be wrong. A few examples will illuminate the


problem.
Fig. 6-15 shows the distribution of palygorskite-sepiolite occurrences for the Late
Pliocene-Holocene interval. Most of the continental deposits are in calcareous soils
and calcretes. The latitudinal association of the continental and deep sea deposits
suggests the latter were derived from windblown dust (Callen, 1984). Note that the
deposits are largely restricted to latitudes 20" to 40"N and 10" to 35"s where
climates are dry to arid.
The latitudinal pattern is present throughout the Cenozoic and Cretaceous (Fig.
6-16) largely between 30" to 40" North and South (Callen, 1984; contains maps of
other time intervals). However, the ocean values are suspect as the distribution of
DSDP cores is not random; no distinction was made between hydrothermal
deposits and those that might be authigenic, and there was no way to allow for
misidentification which is known to have occurred.
Palygorskite is abundant in the recent marine sediments in the Arabian Sea (Fig.
5-37). It is detrital and was transported primarily by winds from the desert area to
384

the north and west. In this same area palygorskite is present in marine sediments
extending from the Recent to the Late Cretaceous. The distribution of palygorskite
in the Miocene sediments is similar to that in the recent sediments (Fig. 5-37),
increasing systematically from approximately 10%in the west-central Indian Ocean
to 40 to 60% near the southern coast of Arabia (Matti et al., 1974a,b). The
palygorskite distribution in the Pleistocene and Pliocene sediments has a similar
pattern. The pattern strongly suggests a detrital origin, particularly when many of
the sediments with a high content of palygorskite are described as brecciated and
turbidite deposits. At least some of the Upper Cretaceous sediments, containing
palygorskite, are described as turbidites.
Palygorskite and, to a lesser extent, sepiolite are relatively abundant (70 to 90%,
< 2 pm, in some samples) in DSDP wells off the northwest coast of Africa. It is
most abundant in Upper Cretaceous, Paleocene and Eocene sediments. The equiv-
alent age sedimentary rocks along the west coast of the African continent, from
Morocco to Angola, contain appreciable deposits of palygorskite and sepiolite (up
to 500 m thick) (Millot, 1970), which was apparently formed in coastal brackish
water environments (Weaver and Beck, 1977). The marine palygorskite and sepiolite
are believed by Peterson et al. (1970), Berger and von Rad (1972) and others to be
authigenic or diagenetic. These physils occur in volcanogenic brown clays, pelagic,
partly zeolitic clays, hemipelagic muds, marl oozes, dolomitic siliceous muds and
cherts. It is apparently not selective as to where it grows (or is deposited). Peterson
er al. (1970) suggested the palygorskite was formed from volcanic ash that reacted
with Mg-rich brines that formed in near-shore or lagoonal environments and flowed
down dip, through the sediments, to the base of the continental slope. Why not
grow the palygorskite in the Mg-rich coastal region and then transport it seaward?
Palygorskite and minor sepiolite are present in sediments ranging in age from
Upper Cretaceous to Pleistocene (9 wells, DSDP 2, 14, 47). It is consistently most
abundant in Paleocene-Eocene sediments, as it is on land. Samples commonly
containing 70 to 90% ( < 2 pm) palygorskite.. It is erratically present in other age
sediments, commonly comprising 20 to 40%(a few higher values) of the physil suite,
including the Pleistocene. Kaolinite (20 to 40%), illite (10 to 45%), and quartz are
present in most samples containing palygorskite. The physil suite, the age distribu-
tion, and nearly continuous deposition for the past 100 m.y., strongly suggest most
of the palygorskite is detrital. At various times it was probably transported by both
water and air. Einsele and von Rad (1979) suggested that during the Cretaceous a
large (600 km wide) coalescing delta system existed along the northwestern coast of
Africa, which presumably was the source of much of the detrital material.
A more recent series of wells immediately off the northwest coast of Africa
(DSDP, Leg 41) has essentially the same distribution of palygorskite-sepiolite.
Timofeev er al. (1977) found palygorskite and/or sepiolite in organic clays,
carbonate, terrigenous and chemogenic sediments. Most of the minerals are detrital
and occur in turbidites and deltaic deposits interbedded with pelagic beds. They
concluded that most of the palygorskite-sepiolite was transported from the African
continent by fluvial and eolian means and some retransported in the marine
environment. They also suggest that pelagic, dolomitic palygorskite-sepiolite clay,
385

without admixtures of terrigenous and volcanogenic material, formed authigenically


or diagenetically in pelagic environments. The high Mg and Si concentrations
necessary for their formation are due to the “intense inflow” of Mg and Si released
during lateritic weathering. The intense inflow of fresh water could produce brac-
kish water conditions in the semi-enclosed marine depression and produce the
conditions necessary for the formation of palygorskite (Weaver and Beck, 1977).
The only evidence that the Paleocene-Eocene palygorskite is authigenic or
diagenetic is its relative abundance. The section, composed primarily of clays,
thickens as the rate of sedimentation increases, and the quartz content increases in a
shoreward direction. In general the concentration and volume of palygorskite plus
sepiolite increases shoreward (Berger and von Rad, 1972; Rex, 1970). These trends
suggest a detrital origin.
Palygorskite is present in DSDP cores from the southern portion of the Gulf of
Mexico and the Bahama area in sediments ranging in age from Upper Jurassic to
Pleistocene. Concentrations in the < 2 pm fraction are mostly less than 30% except
for a few Eocene samples, which contain 60 to 75%. Some of the Lower Cretaceous
deposits are shallow water and subaerial deposits which, during the Upper Creta-
ceous, were faulted far below sea level (Enos and Freeman2 1979). Other deposits
containing palygorskite are commonly described as containing shallow-water clasts
and pebbles (pebbly mudstone), shallow-water debris, turbidites and perireef de-
posits. It is likely the palygorskite was derived from back-reef areas on the Yucatan
and Bahama Platform.
There are additional complications. Some of the palygorskite may not be paly-
gorskite. At Site 198, Matti et al. (1973) reported the presence of from 10 to 60%
palygorskite in the Upper Cretaceous sediments. Okada and Tomita (1973) reported
no palygorskite but, instead, called the material hydrous mica. On the other hand,
both papers reported palygorskite in the Cretaceous of Site 196. Unless samples are
glycolated and/or heated, it is difficult to distinguish between palygorskite (and
sepiolite) and I/S with a high illite content. It has been demonstrated (Nesteroff,
1973) that palygorskite is misidentified in some of the routine, automated x-ray
analyses reported in the DSDP reports.
Kastner (1981) determined the frequency of the lithologic units in the DSDP
cores from all oceans containing smectite + palygorskite + clinoptilolite (Fig. 6-17),
which she considered to be an authigenic assemblage. There is a distinct maximum
(40%) in the Late Cretaceous and much lower values in the Early Cretaceous and
Cenozoic. Thus, if this mineral suite is indicative of the submarine alteration of
volcanics, it suggests the favorable conditions for this type of alteration was
restricted in time.
As shown in Fig. 6-17, the “authigenic” assemblage, clinoptilolite + palygorskite
+ smectite, occurs in 40% of the Late Cretaceous rock units. Another graph by
Kastner (1981) indicates that 75% of all Late Cretaceous rock units contain
clinoptilolite; probably a similar or larger number contain montmonllonite. Thus,
only about half the clinoptilolite-bearing rocks contain palygorskite. Forty-two
percent of the Eocene rock units contain clinoptilolite and only 74% contain the
“authigenic” assemblage. It is easy to conceive that the palygorskite could be
386

0- - -0 Clinoptilolite + Palygorskite + Smectite assemblage


Clinoptilolite + Palygorskite + K-Feldspar assemblage
40

Plcist.- Mioc. Olig. Eoc. Paleoc. L.Cret. E.Cret.


Plio.
Geologic age
Fig. 6-17. Changes in frequencies of the mineral assemblages: (0)clinoptilolite + palygorskite + smectite;
(X) clinoptilolite + palygorskite + K-feldspar, with age of deep-sea sediments. Compiled from 37 DSDP
volumes. From Kastner, 1981. Reprinted with permission from the Sea, 7. Copyright 1981 John Wiley
and Sons.

detrital. In a study of the Miocene palygorslute-montmorillonite deposits of the


southeastern United States (Weaver and Beck, 1977) and the continental shelf
(Weaver, 1968), clinoptilolite occurs exclusively with montmorillonite and never
with palygorskite. Off the northwest coast of Africa the rocks with the lughest
content of palygorskite-sepiolite (60 to 95%), and most likely to have formed in
place, do not contain clinoptilolite. Palygorskite may form in a deep sea environ-
ment, but its association with montmorillonite and clinoptilolite is no proof that it
did.

Miscellaneous “Authigenic” Physils

Rex (1967) found a montmorillonite that had been exposed to seawater for 60
m.y. contained some interlayer ferric hydroxide which he suggested had precipitated
from seawater. Swindale and Fan (1967) reported chlorite forming from gibbsite in
the Wainea Bay, Hawaiian Islands, but the data are not conclusive. Bonatti and
Arrhenius (1965) reported chlorite forming from Fe oxide coatings (desert varnish)
on quartz grains in Baja California.
Studies of pore waters from muds in the East China Sea (Mackin and Aller,
1984) indicate there is a systematic decrease in dissolved A1 with depth. The data
suggest a dissolution-reprecipitation mechanism, whereby a small amount of A1 is
rapidly dissolved from the detrital physils, complexes with dissolved Si to form
dioctahedral chlorite. The theoretical calculated formula is:
X0.9~M~0.77A15.0Si2.7010(0H)8~
387

They concluded that the amount of authigenic chlorite formed would be very low
and difficult to detect.
Rex and Martin (1966) described a sample of weathered granodiorite, primarily
K-feldspar, from the Carmel and Monterey submarine canyons, California, which
contained kaolinite, mica, montmorillonite, halloysite and talc. They believed this to
be an example of submarine weathering. Keene and Kastner (1974) found what
appeared to be authgenic mica (no Mg and trace of Fe) in a chert layer from the
Pacific Ocean. They suggested it formed from montmorillonite, releasing Si to form
the chert. No other occurrence has been reported.
There is some possibility that talc forms authigenically under marine conditions
but likely in waters that are locally hypersaline (see Chapter V). Glauconite and
chamosite are discussed in the following section.

PHYSILS THAT GROW IN SHALLOW MARINE, BRACKISH, A N D


02/09/2020 (1)
EVAPORATIC ENVIRONMENTS

In this section we will consider the authigenic physils that form in shallow marine
(glaucony and berthierine) brackish or schizohaline (palygorskite and sepiolite) and
evaporatic (Mg-silicates) environments. The volume of physils formed in these
environments is small but the physils are of particular interest because they are
excellent environmental indicators, even though we are not exactly certain of the
specific environmental conditions in which they grow or crystallize.

02/09/2020 (1) Glaucony and Berthierine

02/09/2020 (1) Origin


Rounded, greenish colored grains are relatively common in Holocene continental
shelf sediments. These grains are either classified as glauconite or berthierine
(formerly chamosite). The term glauconite has been used to refer to green grains and
as a name of a mineral species. Odin and Matter (1981) suggested the term glaucony
(pl. glauconies) be used to designate green grains composed primarily of 2:l physils.
The term glauconite would be used only as a mineral name. Glauconite is basically
an Fe-rich, dioctahedral illite. In a broader sense, the term glauconite is used to
designate a family of Fe-rich 2:l physils with varying ratios of expanded (smectite)
and non-expanded layers (p. 86). Berthierine is an Fe-rich, trioctahedral 1:1 physil.
Both of the Fe-rich physils are authigenic but commonly modified during burial
diagenesis.
Glaucony is abundant in sedimentary rocks, primarily sandstones and limestones,
ranging in age from Cambrian to Recent. It is particularly abundant in rocks of
middle Cretaceous, early Cenozoic, and Quaternary to Recent ages (Odin and
Matter, 1981). Glaucony can occur as a few scattered grains or be abundant enough
so that the sand is called greensand.
In modern sediments, glaucony is abundant, 10 to 80%, in sands and muddy
sands on the outermost shelf and upper slope between 200 and 300 m water depth;
388

Fig. 6-18. Distribution of glaucony on the present sea floor. Solid round symbol: glaucony identified by
X-ray diffraction; hatched areas are occurrences of unidentified green grains. From Odin and Matter,
1981. Copyright 1981 Elsevier Pub. Co.

however, it is commonly present at depths ranging from 100 to 500 m (Odin and
Matter, 1981). Glaucony is also present in deep sea sediments, commonly at depths
between 1000 and 2000 m. These glauconies occur mostly on rises, ridges, and
seamounts in tectonically active areas (offshore Japan, west of Mexico and Cali-
fornia, and southeast of southern Argentina) (Odin and Stephan, 1981). It is not
known whether these glauconies formed at depth, on “highs”, where deposition was
slow; or if the present “hghs” were originally much more elevated and subsidence
occurred since glaucony formed at shallow water depths. Fig. 6-18 shows the
distribution of glaucony on the present seafloor. Much of the glaucony in modern
sediments is presumably authigenic; but in some areas, such as off the southwest
coast of Africa, it is reworked from older sedimentary rocks.
Berthierine occurs in shallower water than glauconite and commonly in areas
adjacent to river deltas. It is restricted to the equatorial region, suggesting warm
temperatures are required for its formation. Berthlerine is believed to form by the
alteration of kaolinite, and perhaps other physils, in moderately reducing environ-
ments where Fe is mobilized during the early stages of diagenesis. Like glauconite,
berthierine usually occurs in the form of pellets, probably fecal, or ooids with a
peloidal nucleus (Bhattacharyya, 1983).
Glaucony occurs as granular material, mostly in the 100 to 1000 pm size range,
and as surface films partially replacing various substrates such as calcareous fossils,
feldspar, quartz, etc. Films usually indicate an early stage of glauconitization. As
replacement continues the entire fragment can be glauconized, producing glaucony
grains. The morphology of granular glaucony has been described in detail by
Triplehorn (1966), Boyer et al. (1977), Odin and Matter (1981), and others. The
latter authors observed that four kinds of substrate (replaced material) are common:
389

Fig. 6-19. Main types of glauconitized substrates. (A) Dark green casts of foraminifera1 tests (continental
shelf of N.E. Spain). (B) Light to dark green coprolites reflecting different stages of glauconitization
(subsurface Quaternary off Senegal). (C) Green shell debris with original shell texture (zebra texture)
preserved on lower surface of particle (center) and a cracked and bulbous upper surfaces (Lutetian, Paris
Basin). (D) Accordion habit developed by glauconitization of mica grains (shelf off N.W. Spain). From
Odin and Matter, 1981. Copyright 1971 Elsevier Pub. Co. courtesy G.S. Odin.

(1) internal molds or casts of carbonate microfossils, commonly Foraminifera; (2)


argillaceous or limy fecal pellets, usually produced by filter feeding organisms; (3)
biogenic carbonate debris; and (4) mineral grains and rock fragments including
quartz, feldspar, biotite, muscovite, calcite, dolomite, phosphates, volcanic glass
shards, volcanic and plutonic rock fragments, and chert grains. The various grains
are described as ovoidal, spheroidal, tabular, discoidal, lobate, vermicular botryoidal,
etc. For pictures of various morphologies see Triplehorn (1966). Fig. 6-19 shows a
few of the wide variety of morphological types. Vermicular or vermiform glaucony
pellets are usually assumed to have formed from mica books, but many probably
formed from decapod fecal pellets. These are rod-shaped pellets with transverse
bands consisting of separate packs of fecas (Pryor, 1975).
The genesis of glaucony has long been, and still is, a problem. One would expect
that with the abundance of authigenic glaucony in modern sediments it would be
easy to determine the mechanism of its formation. For whatever reason, it has not
been easy. In the 1930’s it was suggested that it formed from gels (Twenhofel, 1936),
recrystallization of fecal pellets (Takahashi, 1939), and alteration of biotite grains
390

IQ

K,O %

1.47

2.30

3.0

4.25

4.19

6.60

Fig. 6-20. X-ray diffractograms on powder mounts of sediments from the Gulf of Guinea. Top, mud fine
fraction; others, glaucony (grains). The samples are arranged from top to bottom in order of increasing
potassium content. The evolution is shown: (1) in the change of position of (001) diffraction peak from
14A towards 10A; (2) in the disappearance of the initial substrate of glauconitization: kaolinite (k) with a
little quartz (q) and traces of calcite (c). The most evolved grains (bottom) show no definite traces of the
initial substrate. After Odin and Dodson, 1982. Reprinted by permission from Numerical Dating in
Stratigraphy. Copyright 1982 John Wiley and Sons.

(Galliher, 1935). Burst (1958), and later Hower (1961), proposed that glaucony
formed from degraded layer lattice silicates by the simultaneous adsorption of Fe
and K and an increase in the proportion of 10 A layers. Millot (1970), Odin and
Matter (1981), and others have proposed that glauconitic minerals form by direct
precipitation or replacement of other minerals.
Fig. 6-20 contains x-ray patterns of the fine fraction (top curve) of bottom mud
and glaucony grains from the Gulf of Guinea. The physil in the Recent muds is
kaolinite, possibly with some K containing amorphous material (1.47% K 2 0 ) . The
earliest formed, light-colored grains are composed of glauconitic smectite with 2 to
3% K,O. As the grains evolve and increase the K 2 0 content (up to 6.6%), the
391

proportion of glauconite 10 A layers increases (top to bottom). Some kaolinite


persists in the initial stage of glauconite formation but is eventually destroyed, as is
quartz.
The occurrence and nature of glaucony affords a wealth of information concern-
ing its origin. Glaucony forms primarily on continental shelves between 100 and
500 m. In this depth range deposition is negligible, reworking occurs, and the
substrates are exposed for sufficiently long times to be glauconized. Glaucony is
commonly associated with unconformities. During marine transgressions the zone of
formation progressively moves landward and glaucony is formed or spread by
reworking over large areas. It occurs from 50' south to 65" north, suggesting climate
is not a controlling factor. It forms in the pH range of 7 to 8 in a semi-confined
microenvironment where there is sufficient organic material to maintain a reducing
environment. The microenvironment (shell tests, fecal pellets, etc.) occurs in an
oxygenated environment. The composition of the substrate is not an important
factor and need not contain any Si, Al, or Fe. The presence of glauconitic material
in foraminifera1 tests, replacing calcite, phosphate, etc., indicates it can form directly
from solution. The obvious conversion of physil-rich fecal pellets to glaucony
indicates it can form by transformation of pre-existing physils, which presumably
are the source of much of the Si and Al. The origmal fecal pellets commonly contain
abundant kaolinite (Odin and Matter, 1981); Ehlmann et al., 1963; Burst, 1958)
rather than 2:l physils. This suggests the physil structure is drastically altered, if not
completely destroyed or converted to an amorphous phase.
The initially formed phase is an Fe-rich ( > 19% Fe203) mixed-layer glauconite-
nontronite (G/N) containing around 2 to 3% K,O and 30 to 50% 10 A glauconite
layers. It has been well established that the Fe is incorporated in the glaucony at an
early stage and much of the K is added later (Ehlmann et al., 1963; Birch et al.,
1976; Odin and Matter, 1981). With continued exposure to seawater, the K 2 0
content and percent of glauconite layers systematically increase to about 9% and
958,respectively. Holocene glauconies contain an appreciable amount of smectite
layers. The only well-ordered glauconite ( < 10%smectite) on the present shelves are
relict, pre-Pleistocene grains. Odin and Matter (1981) estimate that the time
required for the formation of the initial glauconitic smectite is lo3 to lo4 years;
about lo5 to lo6 years are required for a highly-evolved glauconitic mica to form.
Bornhold and Giresse (1985)reported radiocarbon ages of 3,200 to 4,000 years for
Foraminifera containing glauconitic fillings. The continued evolution requires that
the glaucony is exposed to seawater. If buried by an influx of detritus evolution may
stop or be inhibited. However, as most older glauconies have few expanded layers, it
is likely that diagenesis can continue after burial.
Odin and Matter (1981)state that glauconites contain > 15% Fe203 and < 11%
A1 ,03and illites contain < 10% Fe203 (and > 20% A1 ,03). They concluded there
was a lack of samples containing between 10 and 15% Fe,03, suggesting there was
no continuous series between illite and glauconite, and therefore it was unlikely illite
was transformed to glauconite by the substitution of Fe for Al. In a more recent
study Berg-Madsen (1983) described Cambrian Al-rich glauconites from Bal-
toscandia, which contain 8 to 22% A1203 and 5 to 18% Fe203 (microprobe
392

Fig. 6-21. SEM photomicrographs showing the main types of manostructure of glaucony. (A) Ill-defined
globules and caterpillar structures typical of nascent glaucony (4% K20), (B) Boxwork and rosette
structures of evolved glaucony (6-5% K20), (C) Lamellar structure of highly-evolved glaucony (8%K20).
From Odin and Matter, 1981. Copyright 1981 Elsevier Pub. Co. Courtesy G.S. Odin.

analyses). These analyses suggest there is a continuous series between illite and
glauconite but does not necessarily indicate that one transforms to the other. It is
not known whether the A1 glauconites formed on the sea floor from an Al-rich
substrate or if A1 replaced Fe during burial diagenesis. The authors suggest it
formed in a cool shallow marine environment.
SEM pictures (Fig. 6-21A) indicate that in the initial stage (nasent) the glaucony
often contains small, ill-defined globules less than 0.5 p m in diameter. These
coalesce to form caterpillar-like structures. At the stage where the initial substrate is
completely destroyed, bladed 4-5 p m aggregates resembling rosettes are present
(Fig. 6-21B). Well-ordered micaceous glauconite has lamellae up to 5-10 p m long
arranged in a sub-parallel alignment (Fig. 6-21C). In general the early formed
glaucony grains are white, yellowish, or light green in color. With increasing K,O
content and percent of mica layers the glaucony becomes greener. In general, dark
green grains are the most “mature”. As the expanded layers convert to glauconite,
water is lost and the specific gravity increases from about 2.2 to 3.0.
Many glaucony grains have a -
10 p m rim that contains 5 to 15% less Fe oxide
than the main body of the grain (Velde and Odin, 1975; Odom, 1976), which
suggests either a later concretionary growth or crystallization outward from the
interior.
Odin and Matter (1981) found that the granular substrates on which glaucony
form are always highly porous either because of primary intraparticle porosity
(boring or solution of biogenic particles) or the presence of fractures and fissures. In
detrital micas the glauconite does not replace the mica initially, but grows in the
open pore space between mica sheets. They believe that initially glauconitic crystals
grow in these pores. After the pores are filled the substrate disappears, by dissolu-
tion, as glauconitic material eventually “replaces” the entire grain. At the same
time, the earlier formed smectitic phase recrystallizes, incorporating K, to form a
393

more highly evolved micaceous phase (precipitation-dissolution-recrystallization


theory). Why Fe-K silicates should grow under these conditions is not clear.
Presumably the microenvironment in the void areas allows the Fe, Si, and eventually
K to concentrate, perhaps added by microbial activity.
Glauconitic smectite forms at or near the sediment/water interface in the
transition zone between oxidizing conditions above and reducing conditions below.
Ireland et al. (1983) suggest Fe2+ is mobilized at depth and migrates upward,
causing the pore waters in the transition zone near the sediment/water interface to
be supersaturated in Fe2+. Nontronite or glauconitic smectite is precipitated,
depending on the concentration of Fe2+. This theory may explain some of the
occurrences of glaucony, but the association of glaucony with biotite and areas
where abundant Fe-organic complexes are delivered to the oceans suggests there are
other sources of Fe.
An excellent example of early glauconitization is the study of recent glaucony on
the continental shelf off Vancouver Island, S.W. Canada, by Bornhold and Giresse
(1985). Glauconitic smectite is forming today, and for the past 5,000 years (radio-
carbon ages). The principal substrates are eroded semi-indurated Pleistocene
glauciomarine muds and sandy muds which underlie the outer shelf and slope.
Other substrata include Pleistocene foraminifera and mica grains.
Blades of authigenic glauconitic smectite (G,”) containing up to 30% Fe20,
precipitate along fissures, in pores, and on the inside walls of foraminiferal tests.
The occurrence is similar to that of secondary physils in subsurface sandstones.
Clusters of blades range from 2 to 9 pm in diameter (Fig. 6-22). X-ray patters of
glaucony grains indicate both the glauconitic smectite and the substrate physils -
illite, chlorite, and smectite - are poorly crystallized. This suggests there has been
some breakdown of the original physils. K 2 0 values of the early precipitated
glauconitic smectite range from 2.4 to 3.8%K 2 0 (microprobe). Where development
of the glauconitic material is more advanced, dark green grains, the K 2 0 values
range from 4.3 to 7.4%.The total grains contain only 0.8 to 2% K 2 0 . It appears that
the K begins to enter the glauconitic smectite structure after the grains achieved an
Fe203 content of 27 to 30% Fe203. They also found that this second stage or
advanced stage of authigenesis occurred before much ( < 20%) of the substrate was
replaced. The Fe is believed to have been derived from detrital Fe-rich silicate
minerals.
Though fecal pellets, and probably foraminiferal test fillings, have a relatively
high organic content, which is believed to play a role in the mobilization of Fe, by
the time of glauconitization much of the organic material has been destroyed by
microbial activity (Cahet and Giresse, 1983). Further, though various organisms
modify minerals as they are passed through the organism (Pryor, 1975), there is little
evidence to indicate that fecal pellets are significantly different from the intake
material.
With all the investigation of glaucony and glauconite, the processes by which Fe
is concentrated in certain microenvironments, the original substrate silicates are
destroyed, and K is progressively incorporated in the 2:l silicate are not really
understood. The formation of abundant low temperature hydrothermal
394

Fig. 6-22. (A) Glauconitic smectite (right) within mica (left). (B,C,D) Highly evolved glauconitic smectite
in mudstone fragments. (E,F) Rosette clusters of glauconitic smectite in interstices of rnudstone
fragments. Single bar = 5 pm. double bar =10 pm. From Bornhold and Giresse, 1985. Copyright 1985
Soc. Econ. Paleo. Miner. Courtesy B.D. Bornhold.

celadonite/nontronite and celadonite-glauconite in the ocean ridge areas indicates


that K-Fe-layer silicate is probably a stable phase in the marine environment.
In the marine environment Fe is considerably more mobile than Al, and if an
authigenic physil is going to precipitate in any volume it should be an Fe silicate.
However, there is the constraint that glauconitic physils contain on the order of 5 to
395

10%Al,O,. In the ridge areas Fe is obtained from basalt. In the coastal areas Fe is
available in Fe-containing minerals, Fe organic complexes, Fe hydroxide particles,
and Fe coatings on quartz and detrital physil particles. Under reducing conditions
some of this Fe is mobilized and can diffuse towards voids where oxidizing
conditions exist. It is necessary that S’- not be present; otherwise, the Fe will be
precipitated as pyrite (Berner, 1984). Silica can be supplied by the dissolution of
biogenic silica, but as A1 and Mg are also required it is likely that in most instances
most of the Si, Al, and Mg is obtained from the solution and/or microbial
alteration of detrital physils. Some of the Mg may come from seawater as it does in
the alteration of deep-sea basalts. The K concentration of seawater is not high
enough for glauconite to precipitate directly. This causes the formation of glauconite
to be a two-step process, with a glauconitic smectite precipitating initially. When the
K activity reaches a critical level, apparently periodically, portions of the glauconitic
smectite recrystallize to glauconite. Though much of the K presumably comes from
seawater, some may come from K silicates, such as micas and K-feldspars. Organic
activity may play a role in concentrating K. Some algae and bacteria have the ability
to concentrate K in the cell sap 1000-fold over that in the nutrient solution. The ash
of marine algae contain 20 to 40% K,O. When these organisms die the immediate
environment should have a high K content (Weaver, 1967). Birch et al. (1976) found
a good linear relation between the K,O and MgO content of glauconites with a wide
range of smectite layers. This suggests that Mg as well as K is obtained from
solution; the increase in Mg could cause an increase in layers charge and promote
the fixation of K.
Harder (1980) was able to precipitate glauconitic smectite at 20°C from a
solution with a pH of 8.5 and containing 1 ppm Fe, 0.15 ppm Al, 13 ppm Si, 1000
ppm KCl, and 1000 ppm dithionite. When the relative amount of Si was signifi-
cantly reduced, berthierine was precipitated. Thus, berthierine is more likely to form
in shallow water in muddy sediments where soluble silicious organisms (i.e., diatoms
and radiolarias) are less abundant.
It is of interest to note that Fe smectite can convert to Fe illite (glauconite) on
the seafloor and without any increase in temperature, whereas A1 smectite com-
monly requires relatively deep burial and temperatures of approximately 200 to
250°C to alter to illite with 5 to 10% smectite layers. There is some evidence to
suggest that A1 smectite can alter to illite at shallow depths and before burial. As in
the case of glauconite, the transformation is presumably forced by changes in
solution chemistry rather than temperature. In both instances a high concentration
of K is required.
The distribution and significance of glauconitic minerals, berthierine (7 A), and
chamosite (14 A) throughout the geologic column has been reviewed by Van Houten
and Purucker (1984). Berthierine (7 A) is probably formed first and with age, or
temperature, alters diagenetically to 14 A chamosite; however, as both can occur
together the explanation is not that simple. Van Houton and Purucker refer to both
minerals as chamositic minerals. They concluded that, in general, ancient glauconite
and chamositic deposits have an environmental distribution similar to Recent
deposits: chamositic peloids on the inner shelf and glauconitic peloids on the middle
396

MA

100

200

300

400

500

600
Fig. 6-23. Comparative Phanerozoic record of glauconitic peloids and chamositic ooids. Tally of oolitic
ironstones is based on actual number reported (mainly from Kimberley. 1978, and Zitzmann. 1977,
1978). Plot of glauconitic deposits is an estimate of relative abundance based on incomplete data. Sea
level curve and time scale mainly after Vail et al., 1977. From VanHouten and Purucker, 1984. Copyright
1984 Elsevier Pub. Co.

and outer shelves. However, there are numerous exceptions. Suggested environments
range from intertidal to shelf, deep marine, and delta front. Some of these occur-
rences are probably due to reworking.
Though Recent berthierine occurs as structureless peloids, ancient chamositic
minerals are in multi-coated ooids, presumably formed by rolling on the sea floor.
These latter grains are commonly concentrated in chamositic oolitic ironstone beds.
Fig. 6-23 shows the distribution of glauconitic peloids and chamositic ooids
throughout Phanerozoic time. Their temporal distributions are somewhat similar.
Both attained a maximum when cratonic blocks were widely dispersed and sea level
was high. The development of chamositic ooids commonly coincided with repeated
regional transgressions.
On the basis of a large number of chemical analyses of glauconies from Russia,
Shutov et al. (1972) suggested that the initial glauconitic minerals were Fe-rich
nontronites or Fe-Al-montmorillonites when mafic minerals were supplied to the
basin and montmorillonites when the source material was weathered and/or acid
pyroclastic material. They also observed that with increasing age the " illitization" of
glauconite occurred by the diagenetic replacement of octahedral Fe by Al.
391

On the other hand, Sorokin et al. (1980) found that the glaucony in a 40 m thick
Lower Paleogene coastal sandstone had a systematic increase in A1 and a decrease
in Fe and K with decreasing depth in the section. The amount of glaucony and the
proportion of 10 A layers increased with depth. The grain size increased upward
and the authors believed the trend reflected a gradual shift from an open sea facies
(bottom) to a nearshore freshwater facies (top). The difference in chemistry is
believed to reflect differences in water salinity.
Odin (1984, 1985) has observed that the Fe-rich green grains, composed of a 7 A
physil, that occur in the tropics off the west coast of Africa, are not berthierine as
has been assumed. The octahedral sheet of berthierine is filled primarily by Fe2+
and Mg2+ions, whereas, Fe3+is the dominant cation in the 7 A coastal marine clay.
The detailed mineralogy of the material, provisionally called phyllite V *, has not
been determined. It occurs in shallow marine environments with a fluvial influence,
which presumably lowers the pH. This physil, phyllite V, has not been found in
sediments older than Late Quaternary and is presumably converted to another
phase during diagenesis.

Isotopic Age Dating


Because glauconite apparently acquires most of its K over a time period of lo3 to
lo6 years and, as growth of glauconite normally occurs in shelf-slope areas where
deposition is slow, the K-Ar ages of glauconite should be reasonably close to the
depositional age. This is only sometimes true. Though there have been many K-Ar
measurements of ancient glauconites, relatively few measurements have been made
of modern glauconite. If glauconite is truly generated from K obtained from
solution, modem glauconites should have a very young age, presumably less than a
few hundred thousand years.
The K-Ar data on glaucony in modern sediments has been reviewed and
contributed to by Odin and Dodson (1982). They presented data on glauconized
coprolites from the Gulf of Guinea. The marine clay in this area is primarily
kaolinite; however, it contains about 1.5% K,O and has an apparent K-Ar age
around 500 m.y. The authors do not know the site of the K and the radiogenic
argon, but it may be in weathered amorphous material.
The glaucony pellets are estimated to be less than 25,000 years old. The apparent
K-Ar age of the glaucony pellets ranged from 149 to 12 m.y. and decreased as the
K 2 0 content increased from 3.00 to 6.6%. Thus, even the g!auconitic samples with
6.6% K 2 0 and no evidence of the substrate contain some inherited Ar. Analyses of
samples from other areas produced only one with a zero age. Plots of apparent K-Ar
age vs. K,O content indicate that all inherited Ar is not lost until the K 2 0 content
is equal to or greater than 7%. Odin and Dodson found that buried Recent and
Cenozoic glauconitic grains that have less than 7% K,O have inherited Ar. They
concluded that isotopic equilibrium (zero age) and nearly full mineralogical evolu-
tion to glauconite with 7 to 8% K 2 0 occurs mostly or entirely while the grain is in
contact with seawater. Once a grain is buried, geochemical evolution in glaucony

* This material has been given the name Odinite (Bailey, S.W., 1988, Clay Minerals, 23: 237-247).
398

essentially stops. This conclusion is apparently based on the study of shallow buried
sediments. Deep burial may cause diagenetic changes whch would increase the K
content and decrease the inherited radiogenic Ar.
Krylov and Logvinenko (1979) reported apparent K-Ar ages for approximately
20 glauconites in Recent ocean sediments that ranged from 2 to 70 m.y. These
glauconites are presumably detrital or Recent, containing inherited Ar.
Baadsgaard and Dodson (1964), Hurley (1966), Odin et af. (1978), and Odin
(1982) have published extensive reviews on the K-Ar dating of ancient glauconites.
Comparison of K-Ar ages versus stratigraphic age (Hurley, 1966) indicates a close
correspondence for some Cenozoic glauconites and a relatively systematic diver-
gence with increasing age. For older glauconites, K-Ar ages average 20% lower than
the stratigraphic age.
It has been demonstrated that if glauconite grains in young sediments are
carefully chosen, mainly having a high K ( > 7%) content, the K-Ar ages will be
reasonable (Odin et af., 1978; Fitch et al., 1978; Odin, 1982). Three explanations
have been given for the low K-Ar ages. Ar may be lost due to increase in
temperature (degassing); K may be gained during diagenetic reactions which are
also temperature dependent; or the expanded layers may not be effective in
retaining Ar (Odin et af., 1978).
When glauconite is exposed to a burial temperature in excess of - 200°C there is
a loss of Ar and a decrease in apparent age. The loss is related to recrystallization
and is gradual. Some radiogenic Ar is present even when the glauconite has been
exposed to temperatures higher than 400°C (Odin, 1982). Studies of glauconite-
bearing formations that have been subjected to low-grade metamorphsm (Frey and
Hunziker, 1973; Clauer and Kroner, 1978) showed that glauconite was converted to
stilpnomelane. Illite crystallinity data indicated that the conversion occurred at the
anchizone level of metamorphism (200"-300"C). The apparent K-Ar ages of the
stilpnomelane containing rocks are similar to the age of metamorphism, indicating
the K-Ar clock had been reset. The apparent K-Ar ages obtained by Clauer and
Kroner were slightly higher than the Rb-Sr ages. They interpreted this to indicate
that during the conversion of glauconite to stilpnomelane K was lost more readily
than 40Ar.
It has not been established, to my knowledge, that glauconites increase their K
content during burial diagenesis as does montmorillonite.
Potassium-argon has been the usual method of dating glauconites, but there has
been some interest in dating them by the rubidium-strontium method, partly as a
check on K-Ar ages (Brookins, 1976), and partly because the Rb-Sr system is
thought by some to be "less susceptible to mild disturbance than is the K-Ar
system" (Morton and Long, 1980). Herzog and coworkers (1958) suggested that the
Rb/Sr ratio may be unaffected by weathering, but this has not been corroborated.
Hurley et al. (1960) thought that with deep burial, although there is a loss of Ar*,
there probably would not be of Sr*; however, Morton and Long (1980) think that
deep burial does cause the ages to be too young. This effect is not as pronounced,
and may not be present, if the matrix rocks are relatively impermeable (Morton and
Long, 1980). Daughter Sr* does not seem to escape as readily as does daughter Ar*,
399

and ways have been developed to obtain dates which are in close agreement with
those expected (Thompson and Hower, 1973; Morton and Long, 1980).
There is a tendency for Rb-Sr, as well as K-Ar, ages to be too young and for the
discrepancy between the expected and the experimentally determined ages to be
greater the older the sediment (Hurley et al., 1960; Herzog et al., 1958; Bentor and
Kastner, 1965; Morton and Long, 1980). This has not been found in all cases:
Chaudhuri and Brookins (1969), for example, found close correlation with what they
expected for their Middle Cambrian glauconite. The low ages can result from either
the addition of 87Rb(similar to the fixing of K in the conversion of montmorillonite
to illite) or the loss of 87Sr*(probably from the expanded layers), the more likely
prospect (Morton and Long, 1980). Rubidium does not seem to move as easily as
strontium. Morton and Long (1980) in a series of experiments utilizing ammonium
acetate (NH,OAc) as a cation exchanger and HCl, Na-EDTA, and buffered acetic
acid as leaching agents on Paleozoic and Cretaceous glauconites, found that
common strontium is exchangeable, whereas most of the radioactive rubidium is
not. They suggest that the 10 to 15%too-young ages often seen in older glauconites
may represent the maximum Sr* available to cation exchange. Even lower ages
could be the result of oxidation of Fe2+ (which frees Sr by decreasing the lattice
charge) during weathering, or of open system behavior such as pressure solutioning.
Morton and Long (1980), in their dating of Paleozoic and Cretaceous glauconites,
found that most samples gave reasonable ages after leaching with dilute acids or
cation-exchanging with the NH,OAc. A few yielded reasonable ages without
chemical treatment. Apparent ages from the other samples were 6% to 20% too low.
After treatment, ages from most of these were what they were expected to be.
Apparent ages from some, but not all, of their Upper Cambrian glauconites, which
persisted in being too low, were interpreted as reflecting deeper burial or metamor-
phism. The degree of ordering and the percentage of expandable layers did not seem
to play a role in the reliability of the data, although samples with a very high
percentage of expandable layers were adversely affected by both acid and NH,OAc.
Harris and Bottino (1974) and Harris (1976) found good agreement between the
Rb-Sr age and the stratigraphic age of Upper Cretaceous samples from North
Carolina. The only acid leaching they did was to remove calcite impurities in one
sample. Harris and Bottino’s dates were calculated using the model age method.
Harris (1976) recalculated the age using the isochron method. The isochron method
alleviates the problem of having to rely on estimates of the initial 87Sr/86Sr ratio
and the Sr of the sample, formed in its microenvironment, was not in equilibrium
with the Sr of the rest of the ocean. For their Upper Cretaceous samples the dates
calculated by the two methods agree very well. One problem of using the model age
method is that the initial ratio must be assumed. Even in more recent time, for
which the ratio is fairly well known, there is a problem with samples whch formed
in microenvironments not in equilibrium with the open ocean. However, Chaudhuri
and Brookins (1969) did find that the date calculated for untreated (with acid)
glauconite using the model age method agreed reasonably well with that calculated
for a whole rock shale sample using the isochron method and the stratigraphic age.
Numerous studies have shown that reasonable Rb-Sr apparent ages can be obtained
400

from Cenozoic glauconites (Harris and Baun, 1977; Keppens and Pasteels, 1982;
Berggren and Aubry, 1984).
All authors found it necessary to take considerable care in selecting material to
analyze. For example, if other phases are present as intergrowths-such as calcite or
phosphate, which contain a high percentage of common strontium--intense leaching,
enough to remove a “substantial part” of the radiogenic strontium, is necessary
(Morton and Long, 1980).
Rb-Sr ages of Paleozoic glauconites from the Llano Uplift, Texas, are low even
after NH,OAc treatment. Morton and Long (1984) concluded that the ages were the
time of diagenetic recrystallization of the glauconite. The recrystallization was
believed to have occurred during times of regional emergence above sea level and
the influx of meteoric water.
There is an obvious need for continued research in both Rb-Sr and K-Ar dating
of glauconite.

Palygorskites and Sepiolite

Glaucony and berthierine crystallize on the continental shelf where the rate of
sedimentation is relatively slow. Landward, in the beach and deltaic areas, deposi-
tion is so rapid that there is little opportunity for authigenic physils to grow.
However, in some intertidal, supratidal, lagoonal, and similar areas with restricted
circulation, a number of physils, primarily Mg-rich, can grow. The waters in these
environments is commonly different from normal seawater. Two primary processes
affect the chemistry of the waters in many of these peri-marine environments:
mixing of seawater with continental fresh waters and concentration by evaporation.
Talc is present in some tidal sediments, both sands and carbonates, and it is
likely that it crystallizes during the evaporation of trapped sea water. Coastal,
peri-marine deposits of palygorskite and sepiolite are relatively abundant in Ceno-
zoic and Cretaceous sediments: it is questionable whether similar environments exist
at the present time. The formation of chain structure physils in alkaline lacustrine
and possibly in deep marine environments has been discussed elsewhere. A small
amount has formed in desert soils.
The distribution and origin of palygorslute-sepiolite has been discussed by
Weaver and Beck (1977), Lomova (1979), Singer (1979), and Callen (1984). Margi-
nal marine deposits formed along the European, Middle East, and North African
shores of the Tethys Sea and along both flanks of the South Atlantic (mostly in
Africa), and the southern portion of the North Atlantic. Most of these deposits
occur between 30 and 40’ north and south. They were formed in flat-lying coastal
areas in shallow lagoons or bay-like bodies of water that had significant influxes of
fresh water. Climatic conditions appear to have ranged from semi-arid to Mediter-
ranean. Fresh water, relatively rich in Mg and Si, was supplied to the coastal
depressions during the wet season; palygorskite and sepiolite formed during the dry
season when the shallow bodies of brackish water were evaporated to dryness or
near dryness (mud cracks are common) (Fig. 6-24). Another possibility is that heavy
401

Fig. 6-24. Mud cracks in palygorskite clay bed, Miocene, Georgia. Dehydration cracks are filled with
detrital smectite transported by influx of water to dry lagoon.

rain occurred in the interior highlands and the ions were transported to an arid
coastal region. Calcite, dolomite, and opal-cristobalite are commonly precipitated
along with the palygorskite and sepiolite.
The Upper Oligocene and Lower Miocene palygorskite-sepiolite deposits of the
southeastern United States are fairly typical of marginal marine deposits. They have
been described in detail by Weaver and Beck (1977). Two commercial palygorskite
402
palygorsktte Montrnorillonite sepiolite
Meters
0 50 100 50 1 3 a/.
Marine clay
Burrowed

Lagoonal I, 8 : Burrows
M = Matrix
PzPebbln

I
Tidal

\
(I

Marine sand

b
\
\
SOll

Supratidal

Lagoonal

Tidal
Marine sand
i'
Fig. 6-25. Lithology and mineralogy of MC-1 core from La Camelia Mine, Florida. Two cycles of
regression and transgression are evident. Blank areas are commercial clay beds. From Weaver. 1984.
Copyright 1984 Elsevier Pub. Co.

clay beds (less than 10% smectite) 0.9 to 4.5 m thick occur in southwestern Georgia
and northwestern peninsular Florida (Fig. 6-25). The lower clay beds contain
sepiolite; the upper does not. They are separated by a paleosoil which has well-de-
veloped peds and argillans and concentrations of secondary sepiolite.
The clay beds commonly contain clay peds and clasts. The texture and sedimen-
tary structures closely reflect variations in the depositional environment.
Montmorillonite occurs in the marine and continental facies. Palygorskite, sepiolite
403

(with minor stevensite), and dolomite were formed in lagoonal and tidal flat
environments under schizohaline conditions. Where reworking has occurred, both
by currents and by burrowing animals, various mixtures of montmorillonite and
chain physils occur.
Lateral and vertical changes indicate the lower clay bed was deposited during a
regressive phase that was climaxed by the deposition of a thin organic-rich flood-
plain deposit on top of the lagoon-tidal sediments. A soil developed on the flood
plain. This was followed by a transgression and depositon of a shallow brackish
water sand. A second regression occurred during which the upper clay bed was
deposited. A ped and burrowed structure was developed on top of this clay bed.
During the final transgression montmorillonitic brackish water clays, sands, and
coquina were deposited. The sea shortly withdrew from the area and a thin layer of
continental sands was deposited.
TEM and SEM pictures show a number of interesting features. Short, 1 p m fibers
comprise the bulk of the palygorskite-sepiolite clay, but long (greater than 10 pm)
fibers are locally abundant (Fig. 6-26). Long fibers occur in small areas with
desiccation features, indicating they grew from residual fluids when dehydration
was nearly complete. These occur in a matrix of short fibers. Long fibers occur in
the soil samples where they form mats and are also aligned perpendicular to vein
walls.
Short fibers were observed forming from montmorillonite, replacing quartz and
calcite fossils, and by the coalescing of small opaline spheres. Much of the clay
occurs as thin, parallel laminae, suggesting a periodic supply of detritus (montmoril-
lonite) to the lagoon. Many quartz and feldspar grains in the soil have been etched
and many contain a clay skin (palygorskin).
Coarse, rice calcite occurs in vertical fractures and in horizontal bands in the
clay. In both instances it was precipitated in desiccation voids. When the desiccated
and fractured clay surface at the bottom of the lagoon was invaded by water and
detritus, some Mg was adsorbed by the physils. As evaporation commenced fine
calcite precipitated in the voids, eventually recrystallizing into spar calcite. The
Ca/Mg ratio was lowered to the extent that palygorskite (from montmorillonite)
and/or dolomite (from solution) formed. Shells in marine sands underlying some of
the lagoonal clay have been replaced by palygorskite and dolomite. Seeping Mg-rich
waters from the lagoon established a dolomitization gradient.
In some areas it appears that montmorillonite reacts with dolomite or Mg calcite
to form palygorskite and calcite (fine disks).
Opal-cristobalite formed from dissolved diatoms and sponge spicules is relatively
abundant. It is commonly massive but occurs as bladed spherules and well-rounded
opaline spheres.
Chemical analyses indicate the montmorillonites are the Wyoming type. Ap-
proximately half the octahedral positions of the palygorskite are occupied by Al.
Calculations suggest the smectite in the palygorskite-rich clay beds is stevensite.
Chemical calculations suggest that most of the short-fiber palygorskite formed by
the direct alteration of montmorillonite. The A1 and Fe remained constant and
additional Si, Mg, and H were obtained from solution. When the Si and Mg content
Fig. 6-26. SEM picture af long and short palygarskite fibers, Miwene, Ga. White bar = 1 pm.
405

'16 7 1

/
,
. .. .. . .
. . . . ........
. . . .Mont.

12t \
-6 -5 -4 -3
log b4sioi]

Fig. 6-27. Stability relations among simplified (Fe-free) palygorskite and montmorillonite and various
corrensites at 25'C. Continuous lines (- ), "ideal" corrensite; dashed lines (- - -), Bradley and
Weaver (1956) corrensite: dotted lines (. ....), hypothetical corrensite. Stippled area shows range of sea
water composition from Elderfield (1976). From weaver and Beck, 1977. Copyright 1977 Elsevier Pub.
co.

of the solution is sufficiently high and the pH is in the range of 8 to 9, montmoril-


lonite will convert to palygorskite. When montmorillonite, or a source of Al, is not
present, sepiolite, rather than palygorskite, tends to precipitate. Dolomite is com-
monly formed contemporaneously with both sepiolite and palygorskite. Calcite is
commonly deposited out of phase with the Mg minerals. Much of the palygorskite
in the limestones is detrital.
Thermodynamic calculations (Fig. 6-27) indicate that in all cases the chain
silicates are favored by an increase in one or more of [Mg2'], pH, and [H,SiO",].
Palygorskite also requires an appropriate input of A1 (and Fe), either inherited
directly from the precursor clay or taken from solution. Sepiolite requires
log[H,SiO",] = -4.25 (around 3.0 ppm SiO, in seawater, assuming y[H,SiO",] =
1.13) for stability with respect to aqueous solution. Palygorskite should form from
montmorillonite at log [H,SiOO,] > -4.29 (around 2.7 ppm SiO,). Thus, from the
point of view of thermodynamic calculations, only slight modifications of normal
seawater conditions are required to form sepiolite and palygorskite. However, if this
were true these minerals should be more common.
Field observations indicate they are also favored by less than normal salinity and
by high temperature.
Calculated stability field boundaries between palygorskite, montmorillonite, and
two different corrensites (Fig. 6-27) show that regardless of the choice of corrensite
composition, it is favored over montmorillonite by higher [Mg"] and pH. The
406

[ H4SiOo4]effect is minor. For the corrensite-palygorskite reaction the importance


of [Mg2'] and pH is variable depending on the choice of corrensite composition.
However, in all cases high [H4SiOo4]favors palygorskite. These calculations tend to
confirm the idea that corrensite is more abundant in the Paleozoic and early
Mesozoic because evaporitic environments with high p H and high Mg were abun-
dant.
The range of seawater compositions stradles the montmorillonite-hypothetical
corrensite boundary. Hypothetical corrensite is assumed to be a montmorillonite in
which half the interlayer space is filled with hydroxy Mg, A1 sheets. This is
reasonable. Though corrensite, as such, may not form in a normal marine environ-
ment, it is likely that hydroxy sheets are present in some marine montmorillonites.
In the Georgia-Florida area palygorskite and limpid dolomite developed in
shallow, coastal brackish to schizohaline waters. Warm temperature caused a high
pH. Both increased the solubility of silica (largely from diatoms). Cooler conditions
during the Middle Miocene made conditions unfavorable for the development of
palygorskite. Magnesium was probably obtained from both seawater and fresh
water.
During the early Cenozoic the westward-flowing Tethys currents supplied warm
waters to the Caribbean region. The convergence of the African and Eurasian plates
in the late Oligocene and early Miocene allowed these currents to swing to the north
and increase temperatures in the coastal waters of the southeastern United States,
allowing palygorskite and phosphate to form. The collision of Europe and North
Africa at Gibraltar at the beginning of the Middle Miocene modified the Atlantic
circulation pattern, allowing cold Arctic waters to enter the western North Atlantic.
With the decrease in temperature the growth of palygorskite ceased.
In lacustrine and soil environments the Si and Mg needed for the growth of
palygorskite is supplied by weathering; however, the common association of phos-
phate and palygorskite in the peri-marine deposits suggests the primary source of
ions is the sea. For whatever reason, dilution of sea water with fresh water (or
hydrothermal water) appears to be necessary for the formation of palygorskite. This
is apparently the same type of mixing that favors the formation of dolomite, but
with a difference, as palygorskite is apparently not present in most mixing-zone
dolomites.
When seawater is evaporated the initial minerals precipitated are, in order,
calcite-dolomite, gypsum, and halite. Though authigenic Mg-rich physils are also
present the physils are usually 2:l or 2:2 physils rather than the chain structure
type. A Mg/Si ratio other than that found in normal seawater is required for the
chain physils to crystallize. The thermodynamic calculations suggest relatively more
Si is required than is found in marine evaporite waters.
Palygorskite has been reported forming in at least two Recent coastal environ-
ments that may be similar to those in which the older deposits formed. Along the
southeast coast of South Australia (Hodge, et a/., 1984) and the northwest coast of
Egypt (Hassouba and Shaw, 1980) beach ridges and dunes are aligned parallel to the
shoreline. The interridge areas contain marls (and gypsum in Egypt) and palygors-
kite that were deposited in swamp (Australia) or lagoonal deposits under evapora-
407

tive conditions. In Australia the ground water in the swamp area is seawater but the
swamps are seasonally flooded by fresh water. The character of the water in the
Egyptian area is not known but as it is adjacent to the Mediterranean and partially
fed by waters from Lake Maryut in the Nile Delta, it is likely that both seawater
and fresh water enter the lagoon. The situation in these coastal areas must be similar
to those in which many of the peri-marine palygorskite deposits were formed.

Evaporite Physils

Evaporite minerals, in any volume, are deposited in climates where evaporation


exceeds precipitation, normally a hot, arid climate. At the present time evaporite
formation is restricted to the tropical and subtropical climatic zones. Paleozoic
evaporites were also deposited in a near-equatorial zone. However, in the Mesozoic
most evaporites were deposited in the northern paleolatitudes and in the Cenozoic
northern temperate paleolatitudes. Though evaporites always formed in a warm
climatic zone, the zone did not have a fixed position with respect to the paleoequa-
tor (Zharkov, 1981).
Evaporites usually contain a variety of Mg-rich physils. The method of their
formation is not well understood.
As seawater evaporates, CaCO, is the first precipitate to form, followed by
CaSO,, NaCl, and finally by the highly soluble Mg and K salts. In natural rock
sequences the initial sediments commonly consist of interbedded anhydrite and
dolomite. This material is overlain by NaCl, which may be followed by K and Mg
salts. Even though appreciable Mg is tied up in dolomite, Mg silicates are present in
all the rocks, indicating Mg and Si are present at all stages. For any appreciable
thickness of evaporites to develop there must be a periodic or continuous influx of
water to the depression where evaporation is occurring.
Salt sequences contain both detrital physils, transported by water and wind, and
authigenic physils. Both may be modified by burial diagenesis. Weaver (1958) and
Droste (1963) examined physils from a variety of North American salt (halite,
gypsum, and anhydrite) deposits and found that each of the major physil types,
including kaolinite, but not the chain physils, was present as a dominant mineral.
Thin clay beds in the salt are more likely to contain detrital physils than are the salt
beds. Droste (1963) noted the temporal distribution of physils in salts was similar to
that in shales (Weaver, 1959). Illite and chlorite (and Ch/S in some deposits) are the
dominant physils in the pre-Upper Mississippian salts; the younger deposits contain
a more complex physil suite. Corrensite, talc, serpentine, and a wide variety of
mixed-layer Ch/S phases, along with detrital smectite and illite, are commonly
present in younger evaporites.
Kaolinite, montmorillonite, illite, and Fe, Al-chlorite in evaporite rocks are
normally detrital; the authigenic physils are invariably sheet-structure Mg silicates
and Fe-rich illite. Among other things, it is not clear at what stage the various
physils form or are modified. They may form in the brine, at the boundary between
light and heavy brine, and at the bottom of the brine pool. Initial halite deposits
408

have a porosity of 40% which decreases to 5 to 10%by 6 to 12 m of burial (Holser,


1979); early diagenetic physils may grow in these pores. Much of the anhydrite in
evaporite rocks was originally deposited as gypsum (40% H 2 0 by volume). With
burial of 300 to 3000 m, gypsum converts to anhydrite. The released fresh water,
relatively warm, dissolves some salts and may cause the diagenetic formation or
alteration of physils. Other hydrous salts may lose water at depth, i.e., carnallite
(KMgCl, .6H,O) + sylvite + MgCl, + 6H20. With concentrated brines of various
compositions available at nearly all stages of deposition and burial, physils in
evaporites can have a complex history. Little or no information is available on the
growth history of evaporitic physils.
Chlorite is a major component of many Paleozoic salts. Microprobe analyses of
chlorite flakes from the Upper Silurian salt of New York indicate they are Mg-rich
and have a relatively uniform composition (Bodine and Standaert, 1977). The
average composition is:

The illite has a more variable composition and is presumably detrital. Lucas and
Ataman (1968) described the degradation of illite and its transformation (aggrada-
tion) to chlorite with increasing salinity (Triassic of France). Nelson (1973) sug-
gested illite was transformed to Mg chlorite in the Mississippian halite deposits of
Virginia. Chlorite is the predominant physil in some facies of the German Stassfurt
salts (Braitsch, 1971). In some instances the chlorite is believed to form by
diagenetic alteration of a primary 7 A phase.
“ Normal” illites are apparently degraded in hypersaline brines, and thermody-

namic calculations (Weaver and Beck, 1977) indicate it should transform to some
+
form of chloritic material in a Mg-rich brine: 2 illites 3.8Mg2++ 6 H 2 0 =
+ +
corrensite 1.6K+ 6 Hf. The illite-corrensite stability field boundary is described
by pH + 0.63 log[Mg2+]- 0.27 log[K+]= 8.17. The value of this expression in
seawater is about 7.5, and only a small increase in p H and/or Mg would be
required to favor corrensite over illite. However, various studies (summarized by
Kossovskaya and Drits, 1970) indicate that Fe-rich (6 to 16% Fe203)illites form in
mildly evaporitic environments (carbonate-gypsum rocks). The MgO content ranges
from 3 to 6%. These 1Md physils are believed to form both by chemical precipita-
tion and the transformation of montmorillonite, illite, I/S, and other Al-Si minerals.
It most commonly occurs in dolomites and is often associated with corrensite.
Well-crystallized illites, in contrast to degraded illite. is commonly in salt deposits
(Droste, 1963; Nelson, 1973; Bodine, 1975; Padan, 1984; Hall, 1985). In most
instances it is not known whether they are detrital or secondary, though K-Ar data
of illites from the Permian of west Texas (Wampler and Weaver, 1987) and
Pennsylvanian of Utah (Padan, 1984) indicate they are largely detrital.
The most abundant authigenic and/or diagenetic physils in post-Carboniferous
evaporites appears to be partially formed chlorites. The starting material may be
either detrital smectite or a vermiculite, authigenic saponite, degraded chlorite or
illite, or a concentrated brine. As the chloritic material contains 15-20% A1,0, it is
409

likely it had a detrital precusor. When the Mg concentration and pH values are
appropriate, Mg hydroxide is precipitated between the expanded layers. The se-
quence of development is not well established, in part because the layer charge of
the 2:l layers influences the development in the interlayer space.
Initially patches or islands of Mg hydroxide are precipitated in the interlayer
space. Complete sheets may not be formed, but the pillar effect prevents contraction
to 10 A when samples are heated at 100°C. However, they will collapse at a
relatively low temperature, - 300°C. As hydroxy growth continues, extensive
hydroxy sheets can develop between selected layers and a random Ch/S developed;
however, it is also possible for pillar development to continue without the formation
of discrete chlorite layers. In either event thermal stability increases. Under some
conditions a regularly interstratified Ch/S, corrensite, develops. In addition, de-
pending on the charge of the 2:l layer, Ch/V (high charge) or Ch/swelling chlorite
(low charge) can develop. With a further increase in soluble Mg and/or temperature
and/or pH, the remaining expanded layers are step-by-step converted to chlorite.
Corrensite and various forms of Ch/S are probably the most abundant physils in
evaporite rocks. It is quite common in dolomitic rock, where evaporation is
relatively minor. It most commonly occurs in shallow-water, tidal dolomites where
evaporation was intermittent and brine concentrations seldom went beyond the
dolomite or gypsum stage (Lippmann, 1956; Bradley and Weaver, 1956; Peterson,
1961; Weaver, 1961; Millot, 1970; Kubler, 1973; Rao and Bhattacharya, 1973).
However, it is also found in rock indicative of more advanced stages of alteration:
anhydrite, halite, and K-Mg salts (Devonian, western Canada (Droste, 1963);
Permian, Germany (Braitsch, 1971); Permian, New Mexico (Bodine, 1978); Per-
mian, Texas (Bassett and Palmer, 1981); Pennsylvanian, Utah (Padan, 1984)).
Talc is another physil that is a common constituent of evaporites. It is present in
the Silurian deposits of New York (Bodine and Standaert, 1977), Pennsylvanian of
Utah (Padan, 1984), Permian of England (Stewart, 1965), Germany (Braitsch, 1971),
New Mexico (Bodine, 1978), and others. In most deposits the talc is a relatively pure
Mg silicate but Al-rich talc with some interlayer hydroxy-Mg is present in the
Paradox Salt of Utah (Padan, 1984).
In many of the deposits that contain talc a 7 A physil is present that is commonly
called serpentine. Braitsch (1971) reported that during the metamorphism of carnal-
lite (KMgCl, . 6H,O) to sylvite (KCl), amesite (Mg-rich 1:l physil) was converted
to chlorite.
Saponite has been found in the Permian salt deposits of New Mexico (Bodine,
1978) and west Texas (Bassett and Palmer, 1981).
The few salt deposits that have been studied in some detail indicate that the
physil suites are commonly complex and are usually different in each rock facies. In
our study of the Paradox Salts of Utah (Padan et al., 1984) we found the physil
suites could differ at centimeter intervals. In fact, it was rare to get the same x-ray
pattern from duplicate samples.
In an early study Fuchtbauer and Goldschmidt (1959) found that in the Permian
Zechstein of Germany distinctive physil suites are associated with specific litholo-
410

gies: illite and chlorite in shales; illite in carbonates; talc, illite, chlorite, montmoril-
lonite, and random mixed-layer clays of various types in anhydrite rocks; and illite
and chlorite in chloride-salt-bearing rocks. Braitsch (1971) found that, in the same
deposits, talc is predominant in the anhydrite rocks and in the anhydrite-rich
horizons of the chloride-salt rocks. It is more abundant in the central part of the
basin. It is rare in potash seams but is present in the kieseritic sylvite-halite from the
Stassfurt seam, associated with some chlorite and with predominant ascharite. Two
varieties of chlorite were found: normal chlorite of the penninite-chlinochlore group
in the Stassfurt rock salt and in anhydritic sylvite-halite and anhydritic halite of the
Stassfurt seam, and chlorite of the amesite-berthierine group in the anhydrite-bearing
halite-carnallites and in sylvite-kieserite-halite rocks. Corrensite was found to be
particularly abundant in the Wera potash seam.
Bodine (1978) analyzed the clay assemblages in 47 insoluble residues of the
potash-bearing Permian Salado Formation of New Mexico. Well-crystallized illite
and chlorite (clinochlore) are the most abundant clay minerals. Corrensite is
particularly abundant in the salzton (clay-salt rock) samples. Random mixed-layer
chlorite/saponite is not present in salztons, but is present in other lithologies,
particularly in anhydrites from the Salado Formation. Talc and serpentine were
found mainly in the rock salt. Minor amounts of saponite and mixed-layer talc-
saponite were found in a few samples. Bodine concluded that the Mg-rich trioc-
tahedral physils formed as a result of extensive reaction of clay detritus with the
brines, followed by recrystallization during burial. He interpreted the characteristic
clay assemblages of the Ochoan salts (random mixed-layer Ch/S, saponite, or
mixed-layer talc/ saponite) as being less mature than the corresponding assemb-
lages (chlorite, corrensite, and illite) of Ochoan salztons. The relatively immature
clays have undergone extensive post-depositional (burial) reaction with salt-pore
fluids created by recrystallization of the salt minerals. On the other hand, the
salztons behaved as impermeable units after burial and remained isolated from the
pore fluids. Therefore the initial assemblages did not undergo the same changes as
the clays from the salts. The more mature clay assemblages in the salztons were
interpreted to evolve through a long time span in a static pore fluid environment.
Bassett and Palmer (1981) summarized the clay mineralogy of a core (No. 1 Rex
White) taken through evaporite sequences in Randall County in the Palo Duro
Basin of Texas, where halite is the end-phase evaporite. Discrete chlorite and illite
along with mixed-layer clays chlorite/swelling chlorite, Ch/V, and Ch/S were
identified in samples interpreted to be from various sabkha sub-environments. Both
mixed-layer Ch/S and illite are present in 50% of the samples.
Regular mixed-layer chlorite/swelling chlorite most commonly occurs in salt-pan
and mud-flat environments. Random to partially ordered mixed-layer Ch/V is
associated with fine-grained clastics in bedded halite of the upper sabkha, in chaotic
mudstone-halite, and in anhydrite of the marginal sabkha.
Mixed-layer Ch/S (expands on glycolation and contracts to 12.3 A at 400°C and
to 10 A at 600OC) apparently contains incomplete hydroxide interlayers that delay
layer collapse to a high temperature. The chlorite-rich mixed-layer clays collapse to
only 12-13 A at 600°C the partial collapse is due to more complete development of
411

hydroxide interlayers. Ch/S is primarily found in bedded halite from the salt pan
and in chaotic mudstone-halite from the saline mud flat.
Palmer (1981) suggested that detrital clays delivered to the sabhka (smectite,
kaolinite, illite, and mixed-layer species) were altered to more Mg-rich clays by
interaction with Mg-rich brines. Variations in the duration of contact between the
detrital clays and the brine resulted in the formation of a continuous series of
mixed-layer clays between saponite and chlorite end-members. In a sequence
beginning with Al-smectite, the detrital clay undergoes intervals of contact with a
Mg-rich brine and is converted to Mg-smectite, saponite. As saponite interacts with
the brine, mixed-layer chlorite-like clays form by the incomplete precipitation of
hydroxides in the interlayer space. Chlorite-type clay forms if there is uninterrupted
contact with the brine and completion of the interlayer hydroxides.
A typical evaporite cycle (80 m) in the Pennsylvanian age Paradox Formation,
Paradox Basin, Utah, was studied in detail by Padan (1984). The cycle starts with a
basal dolomitic zone which grades upward, successively, into anhydrite, halite,
carnallite-kieserite zones. Thin, 2-5 mm, “clay-rich” anhydrite bands occur at
regular intervals in the evaporites above the anhydrite zone. The anhydrite bands
presumably reflect annual influxes of fresher water. The halite and K-Mg salts
usually contain less than 1%water insoluble material of which only a portion is
physils. The anhydrite in the anhydrite bands, dissolved with EDTA, contains from
5 to 50% silicate minerals, mostly physils.
The vertical variability of the physil suite is impressive. The physil suite, talc,
Al-talc, Al-serpentine, illite, corrensite, Ch/S, and Ch/S with different degrees of
completeness of the brucite layers, not only change from zone to zone but on a
micron scale within zones. Both Al-bearing and Al-free physils are present; the
physil suites are related to the bulk composition of the rocks and presumably to
salinity.
The assemblages characteristic of the low-salinity intervals (dolomite and
anhydrite zones) consist of mixed-layer Ch/S and possibly authigenic illite. The
nature of the Ch/S changes from ordered (corrensite) to random and then to
partially ordered going upward. The amount of illite increases from a trace in the
dolomite zone to about 30% in the anhydrite zone.
A major change occurs in the physil assemblages going upward into the salt
zones. Illite disappears and the predominant 2:l physil is talc in halite bands and
Al-talc in anhydrite bands. The disappearance of authigenic illite and the presence
of authigenic talc, or Al-talc, is presumably due to the absence, or near absence, of
Al. The lack of A1 is due to the lack of detritus. It is interesting that despite a large
amount of dissolved potassium in the K-rich zones, illite was not formed. Authi-
genic Al-serpentine is present in the highest salinity intervals (high-K zones),
suggesting a high concentration of dissolved Mg and a relatively low concentration
of dissolved silica. The nature of the mixed-layer physils also changes. In the highest
salinity zones the mixed-layer Ch/S contains a large proportion of chlorite layers.
However, in the intermediate zones (e.g., the halite zone) the Ch/S contains
chloritic components showing swelling properties characteristic of the presence of
incomplete layers of brucite. The extent of the development of brucite layers in the
41 2

14.7
50 - 200 I 200- 300

1 300-400

490-550

Fig. 6-28. Series of X-ray patterns of mixed-layer Ch/S showing shift and disappearance of the main
X-ray peak as the sample is subjected to stepwise heating. Incomplete interlayer hydroxyl material is
progressively dehydrated. From Padan. 1984.

interlayer space may possibly be used to trace changes in salinity. The extent of the
development of hydroxy-Mg material in the interlayer space is indicated primarily
by the thermal stability of the Ch/S (Fig. 6-28) where development is only partial or
pillar-like collapse starts at -
30OOC. As the completeness of the layers increases,
the collapsing temperature increases. When fully formed in all layers there is no
appreciable collapse (chlorite). Under evaporitic conditions, the completeness and
stability of the brucite layers appears to increase with increasing salinity. This is
presumably because the concentration of Mg parallels the increase in salinity.
Relatively well-developed brucite layers can form in low to normal salinity environ-
ments (sandstone pores) where the concentration of Mg (and Fe) is relatively high.
Actually the OH/Mg ratio is probably more important than the Mg concentration.
In the Paradox Salt cycles, the physil suite not only reflects the broad-scale
changes in salinity but also reflects episodic small-scale salinity changes. Local
depositional and post-depositional changes in fluid composition produced a situa-
tion where, except for the pure halite intervals, the physil suites varied from
centimeter to centimeter (vertically). Therefore, the clay minerals in evaporites have
the potential to supply a great deal of detailed information on water chemistry.
It is not clear which, if any, physils are authigenic and which diagenetic. The talc
in the halite commonly occurs as delicate rosettes (Fig. 6-29a) and cornflake
aggregates (6-29b). It is unlikely the talc grew after burial and compaction. The
chloritic material has a more irregular shape (Fig. 6-29c) and may be diagenetic. In
413

Fig. 6-29. SEM pictures of authigenic physils from Paradox Salt, Utah. A. talc rosette in halite. B. Thin
plates of serpentine and cornflake-like aggregate of Al-talc in carnallite bed. C. Lettuce-head of Ch/S
containing small dolomite rhombs; in argillaceous dolomite. D. Thin sheets of Ch/S between plates of
anhydrite, protrudance is also Ch/S. Scale in pm.

other instances sheets of chloritic material alternate with thin plates of anhydrite
(Fig. 6-29d). When and how did this material form? A1 must have been mobilized.
Authigenic quartz crystals are present in this mixture. The presence of A1 in the
chloritic physils suggests they formed, in one way or another, from detrital physils,
probably illite, I/S, and smectite.
The K-Ar and paleogaphic data indicate that most of the water in the Paradox
Basin was derived from the southern ocean. However, at relatively regular intervals
fresh water from the north, Uncompahgre Uplift, flooded the basin. It was at this
time that the thin anhydrite bands were formed and it is in these bands that the
chloritic physils occur. The mixing of fresh water with concentrated brines ap-
parently favors the formation of 2:2 chloritic layer from 2:l layers (detrital) both by
transformation and solution and neoformation. It is not clear whether or not this is
the general situation. It is possible that the chloritic physils are formed at an early
stage, while the anhydrite or gypsum are growing. However, the situation need not
be static. Fine-grained magnesite and dolomite are intimately mixed or intergrown
414

with the chloritic flakes (Fig. 6-29d) and would serve as a Mg reservoir for further
chloritization during burial (Weaver and Padan, 1983).
The effect of burial diagenesis on the physils in salts is not well known, but a few
analyses of halite residues from the Weeks Island salt dome, Louisiana, provides a
clue. Talc, illite, and chlorite all occur as dominant physils in various samples. They
are all well crystallized (have strong, sharp x-ray reflections). The chlorite is
Mg-rich; the illite is the 1M polytype and does not have a high Fe content. The peak
width of these two physils is similar to those of anchimetamorphic physillites. The
Gulf Coast salts have been buried on the order of 10,000 m; however, it is likely the
salt started to move upward before the maximum amount of overburden was
deposited. Nevertheless, the crystallinity of the physils appears to have been
increased by burial diagenesis. Based on the temperature-illite crystallinity data for
shales (Weaver and Associates, 1984), the illites could have been exposed to a
temperature slightly in excess of 300°C, a reasonable value.
Much remains to be discovered about the significance of evaporitic physils.

Brief Summary
The physils we have discussed in this section make an interesting pattern (Fig.
6-30).
Fe3'-rich 2:l physils (glaucony) form on the outer shelf and upper slope.
Fe*+-rich 1:1 physils (berthierine) form in the intershelf area, presumably under

Arid

d, E
d
(olt wet and dry)
- b

lllite ('1

Palygomkite O'?;-"
Seplollte A - 4
CH /S
Talc sapmite Eerthterine

Fig. 6-30. Sketch illustrating the environmental situations in which many of the authigenic Mg-rich and
Fe-rich physils crystallize.
415

more anaerobic conditions. In the lagoonal-bay environment where fresh water


mixes with seawater a Mg-Al-chain silicate (palygorskite) forms. In arid areas
containing semi-closed depressions, fresh water is periodically mixed with hyper-
saline brines and Mg-rich 2:2 chloritic and/or 2:l talc form. Palygorskite and
sepiolite also form in alkaline lakes in arid environments. Climate has little control
on the formation of the Fe3+ silicates but possibly some effect on the Mg2+ silicate.
The chain silicates apparently form under relatively warm and moderately or
seasonably humid conditions. The chloritic material forms in a warm or hot, arid
climate. Fe physils preferentially form in normal seawater and Mg physils form in
both hyposaline and hypersaline waters.
This Page Intentionally Left Blank
417

Chapter VII

DIAGENESIS - METAMORPHISM

DIAGENESIS

Introduction

Once a physil is transported to a depositional basin it starts to settle and seldom


stops until the downward movement is reversed by tectonic activity. The initial
settling rate through the water column is relatively high and the physil is not
materially modified (unless it is eaten). Once the physil becomes part of the bottom
mud the settling rate slows considerably but does not stop. The physil may
gradually settle an additional 1000 to 10,000 m or more before the direction of
movement is reversed and the physil again “sees the light of day”. During this
period of slow settling, which can vary by a factor of 10 or more, the physil is
exposed to a constantly changing water chemistry, inorganic and organic, and to
increasing temperature and pressure; it strives to adjust.
Before we investigate the adjustment mechanisms we need to briefly consider
terminology. The term diagenesis is commonly considered to include all changes
which affect minerals and sediments from the time of deposition until the stage of
metamorphism. Metamorphism, or “true” metamorphism, is the stage where a rock
is completely recrystallized ( - 350°-375”C at 10 km) (Winkler, 1979; Turner,
1981). However, preceding this stage of metamorphism we have very-low-grade
metamorphism and incipient metamorphism. Diagenesis is divided into early,
middle, and late stages. The boundary between diagenesis and metamorphism is
based on both mineralogical and textural criteria which do not always coincide.
Boundaries based on the generation of new minerals are highly subjective and the
temperature can vary depending on which minerals are used as indicators. The
location of the diagenesis-metamorphic boundary is also influenced by the back-
ground of the geologists - whether they are metamorphic or sedimentary geologists.
Metamorphic petrologists tend to believe that in physilites no significant minera-
logic changes occur at temperatures less than - 200OC. On the other hand, clay
mineralogists-petrologists tend to be interested in what happens at temperatures
-
below 200°C (typical Paleozoic shale). However, this transition zone is now being
actively explored ( e g , Weber, 1976; Weaver and Associates, 1984; Frey, 1987). For
a detailed discussion of diagenesis-metamorphic terminology see Kisch (1983).
418

It is generally agreed that diagenesis refers to physical and chemical changes that
occur without the intervention of tectonic activity (i.e., slaty cleavage); thus, the
name “burial metamorphism” was introduced to refer to metamorphism which
occurred without benefit of lateral deformation (Coombs, 1961). The term “ burial
diagenesis” has been used but it would appear to be redundant.
I have chosen to use the terminology suggested by Kubler (1967) and used by
many of the European clay scientists: zone of diagenesis, anchizone (very-low-grade
metamorphism), and epizone (low-grade metamorphism, - greenschist facies). The
boundaries are based on the sharpness (Weaver, 1960) or width (Kubler, 1967) of
the 10 illite peak (see page 449). “Illite crystallinity” will be discussed in more
detail later in this section. The most prevalent diagenetic changes in muds-physilitc-s
involve the conversion of smectite to illite. I (Weaver and Associates, 1984)
suggested dividing diagenesis into three stages based on the I/S ratio of the < 2 pm
fraction:
Early Diagenesis: Ranges from the water-mud contact (not including the nepheloid
layer) to the stage where a regular mixed-layer I/S is, or should be, present
(approximately 60% illite layers; 90 to 140°C).
Middle Diagenesis: Ranges from the first development of a regular mixed-layer
phase ( - 14 to 13 Apeak when glycolated) to the disappearance of a discrete
glycolated peak for the mixed-layer phase ( - 90%illite layers; - 200°C).
Late Diagenesis: Stage where the glycolated mixed-layer phase appears as an
integral part of the 10 A peak ( < 10% smectite layers) to the beginning of the
anchizone (K.I. = 3.0; W.I. = 2.3; - 250” to 280°C).
The temperature at which these boundary changes occur varies with grain size. In
order to standardize the boundaries between diagenesis, very-low-grade and low-
grade metamorphism, it is best to use the < 2 p m fraction.
It is commonly stated that the boundary between diagenesis and metamorphism
occurs at - 200°C when recrystallization of physils occurs. Actually recrystalliza-
tion of physil clay starts at a temperature < 50°C and is not complete until
temperatures in the 350”-400°C range (Weaver and Associates, 1983).
Research in the area of diagenesis and early metamorphism (very-low-grade and
low-grade) of physilites and sandstones has exploded in recent years and there is not
time for an exhaustive coverage. For additional information see Weaver (1979),
Kisch (1983), and Singer and Muller (1983). Diagenesis in sandstones is discussed in
Chapter VIII.

Smectite -+ IIlite
All the physils undergo some modification upon burial. The most impressive
reactions are those involving smectite. The earliest discussions of the conversion of
smectite, largely montmorillonite, to I/S during burial appears to have been those
of Burst (1959), Powers (1959), and Weaver (1959), all from the Shell Research
Laboratory in Houston, Texas. They reported on the Cenozoic of the Gulf Coast
and Carboniferous of Oklahoma. These papers were followed by a number of
papers describing the diagenesis of montmorillonite in other areas: Cretaceous of
the Douala Basin, Cameroun, and Cenozoic and Cretaceous deposits of the
419

Camargue, southern France (Dunoyer de Segonzac, 1964, 1970); Jurassic of south-


eastern France (Artru and Gauthier, 1968); Cenozoic of Azerbaydzhan (Teodoro-
vich ef al., 1967); Jurassic of eastern Ciscaucasia (Teodorovich and Konyukhov,
1970); Mesozoic, New Guinea (Moort. 1971); Cretaceous, British Columbia (Fosco-
10s and Kodama, 1974; Powell et al., 1978); Cenozoic, Germany (Heling, 1974,
1978); and lower Cenozoic-Mesozoic, North Sea (Pearson et al., 1982).
More recent studies in the Gulf Coast, which is probably the most thoroughly
studied area, include Perry and Hower (1970), Weaver and Beck (1971), Hower et
al. (1976), Boyles and Franks (1979), Lee et al. (1985), Ahn and Peacor (1986), and
Bell (1986).
Before we proceed with the discussion, a brief mention should be made of the
basic reaction for the formation of diagenetic illite. It is generally agreed that in
mud-shales, but not necessarily in sandstones, diagenetic I/S and illite formed from
a smectite, primarily montmorillonite, precursor. In the process of smectite convert-
ing to illite the Al and K content is increased and the Fe and Mg content may
decrease. Possible reaction mechanisms of illitization can be classed as either
solid-state, where the 2:l layers remain basically intact (Weaver and Beck, 1971;
Hower et al., 1976; Bethke and Altanen, 1986), or dissolution-precipitation (Boles
and Franks, 1979).
Weaver and Beck (1971) proposed a solid state reaction whereby Al from
K-feldspar and/or kaolinite entered the tetrahedral sheet via the hexagonal holes in
the tetrahedral sheet and replaced Si by a rotation mechanism (Pollard, 1971):
smectite (Fe”) + K-feldspar/kaolinite + I/S/Ch(Fe*’) + Si+4
They proposed the mixed-layer phase contained dioctahedral chlorite layers as well
as illite and smectite layers. In this and all subsequent theories most of the K is
assumed to have been derived from K-feldspar and to a lesser extent mica..
Hower et al. (1976) proposed the following reaction:
+ +
smectite A13+ K + = illite + Si+4;
in addition, some Mg and Fe from the smectite contributed to the formation of
chlorite.
Bethke and Altaner (1986) tested a variety of solid-state transformation mecha-
nisms with a stochastic model, which accounts for interactions among 2:l layers.
Their modeling suggests that important features (x-ray patterns, I/S ratio change
with burial rate) of the illitization reaction can be explained by a layer-by-layer
model of a solid-state transformation. The computer randomly chose smectite layers
as candidates for “illitization”. The model then stochastically decided whether to
illitize the candidate layer, based on the candidate’s neighbors in its crystallite.
“In successful Monte Carlo runs, an illite neighbor acts to increase reactivity
of a smectite layer, whereas two illite neighbors sharply decrease smectite
reactivity, thereby suggesting that two or more interactions among layers com-
pete during illitization. For example, illite neighbors may act as “templates” that
lower energy barriers to reaction of a smectite layer. Two illite neighbors,
however, may polarize charge density in silicate sheets on both sides of smectite
interlayers (Sawhney, 1969), causing resistance to potassium fixation.”
420

The model also shows a moderate dependence of illitization profiles in subsiding


basins on burial rate (time).
Boles and Franks (1979) proposed the following reaction:
smectite + K+ illite + chlorite + quartz + H f
In this reaction there is no external source of Al. The A1 in the smectite layer is
retained (conserved) and the increased concentration of A1 is accomplished by
dissolving, at least partially, the smectite layers. The main dissolved product is Si.
This reaction results in a 24% reduction in the amount of 2:l physil in the shale
when the 1/S ratio increases from 1:4 to 4:l. Measured values are approximately
188, which suggests there is a significant dissolution.
Based primarily on TEM studies, Ahn and Peacor (1986) propose a reaction
mechanism which is intermediate between solid-state and dissolution precipitation.
This is discussed in detail further on in the chapter. Actually, all the proposed
reactions involve contributions from the two mechanisms.
From a study of smectitic shales in contact with a basalt dike, Lynch and
Reynolds (1984) found that the molar concentration of I/S increased progressively
as the dike was approached and concluded the I/S was neoformed. On the other
hand, Pollastro (1985) found that in the Upper Cretaceous of the Denver Basin,
Colorado, as the amount of illite in I/S increased with depth, the relative weight
percent of I/S decreased and discrete illite increased. Pollastro suggested illite and
I/S form diagenetically at the expense of I/S itself. Both illite and smectite layers in
the I/S are presumably cannibalized. Which of these reactions is likely to occur
probably depends on the availability of K.
In the Gulf Coast and most other areas the development of new illite layers in
smectite can be first observed (x-ray) at a temperature of approximately 50°C. The
detrital smectites in the Gulf Coast contain around 20% illite layers. The initial illite
layers are randomly interstratified with smectite layers (RO). With increasing depth
and temperature the proportion of illite layers increases and when approximately
50% of the layers are converted to illite, ordered I/S packets develop (RI); these are
regularly or randomly interstratified with illite layers. When the proportion of illite
layers increases to about 85% an IS11 ordering develops (R3). A sequence of typical
x-ray patterns (Fig. 2-20) illustrates the changes that occur during burial. In the
Gulf Coast ordering develops at 90" to 100°C. Most wells are not deep enough for
IS11 ordering to be observed. Once the I/S ratio reaches approximately 4:1,
commonly at temperatures of 120" to 140"C, the ratio remains relatively constant
until temperatures are in excess of 200°C. The final product is a 1M illite (Srodoh
and Eberl, 1984) that with increasing temperature is converted to the 2M polytype.
The final few expanded layers are not lost until about 350°C (Weaver and
Associates, 1984). Fig. 7-1 shows the percent illite layers in I/S as a function of
depth for a typical Upper Tertiary G d f Coast section (Hower et al., 1976). The
relatively smooth trend is due, in part, to the use of cutting samples. Analyses of
core samples generally show more scatter (Fig. 8-14). Note the difference in I/S
ratio for the two different size fractions in Fig. 7-1. Some of the variability reported
in the literature is due to the use of different size fractions. The variation in I/S
42 1

Percent Wite Layers


Fig. 7-1. Proportion of illite layers in illite/smectite as a function of depth. “Shale” samples from Gulf
Coast well (Miocene-Oligocene). From Hower e t a / . , 1976, Geol. Soc.Amer. Bull., 87, 725-737.

ratios for closely-spaced samples indicates factors other than temperature are
involved in the conversion of smectite to illite - presumably composition and/or
permeability. Analyses of Miocene shale core samples from three wells in an oil field
in Cameron Parish, Louisiana (Hinch, 1978), show that for a given depth the
proportion of illite layers in I/S commonly varies by 15 to 20%.
Weaver and Beck (1971) appear to have been the only ones to suggest the I/S
minerals in the Gulf Coast contain layers of chlorite (I/S/Ch), though Raman and
Jackson (1966) provide data that suggest chlorite layers are commonly present in
illite. Weaver and Beck’s suggestion was based primarily on the direction of shift of
x-ray peaks, peak-width when heated to 300°C, and the presence of approximately
7% easily soluble A1,03 and Fe,O,. TEM data tend to confirm that chloritic layers
are interstratified with Gulf Coast I/S and illites in general (Lee et al., 1984; Ahn
and Peacor, 1985; Lee et al., 1985). The chlorite layers are intergrown and
semicoherent with respect to the surrounding I/S. All layers constitute a continuous
structural network (Ahn and Peacor, 1985). It seems very likely that diagenetic I/S
contains some interstratified chlorite layers but probably not as much as suggested
by Weaver and Beck, and they are apparently trioctahedral rather than dioc-
tahedral, though both may occur.
422

- 7
m -
L
a,
c

m
E
9
- 4
c
c
a
a,
D

10 20
Percent potassium feldspar
Fig. 7-2. K-feldspar content of two size fractions of samples shown in 7-1. K-feldspar was not detected in
the < 2 p m fraction. From Hower et al., 1976, Geol. SOC.Amer. Bull.. 87, 725-737.

In the well in Fig. 7-1 and in other wells, the depth of maximum illitization
coincides with the disappearance of K-feldspar (Fig. 7-2), the major source of K. I t
is suggested (Srodoh and Eberl, 1984) that the failure of smectite to convert
completely to illite is due to insufficient K-feldspar. It is unlikely that this is the
complete answer. In the Conasauga shale-slate, illite and phengite crystallization
continued long after all K-feldspar was destroyed (Fig. 7-59).
Though I won’t discuss it here, it should be mentioned that many thick sections
where illitization of smectite has occurred, or is occurring, are undercompacted (or
overpressured). One cause of undercompaction, which is actually a high pore water
content at depth, is believed to be due to interlayer water that is released to the
pores as smectite is converted to illite (e.g., Bruce, 1984).
One of the major concerns is whether the conversion of smectite to illite is
primarily controlled by temperature, time (kinetic), composition, and to a lesser
extent, pressure. Table 7-1summarizes the Gulf Coast data. The wells are essentially
from two areas. The Louisiana wells contain sediments derived primarily from the
ancestral Mississippi River, and the Galveston, Texas area contains sediments
delivered by the ancestral Brazos-Colorado River system (Bruce, 1984). The west
Texas samples were presumably derived from a third source.
The table shows the temperature of the first detection of an increase in the
proportion of illite layers in the I/S, the temperature at which a stable ratio of
approximately 4:l was obtained, and the temperature interval over which the
423

Table 7-1
Depth-temperature data for the conversion of smectite to I/S in the Gulf Coast Region.
Location First Top I/S Tempera- Reference
and Age illite 4:l-7:3 ture
layers zone Interval
Pleistocene-Pliocene
Louisiana 73T 128OC 55 Perry, 1969
Pliocene
Louisiana 88°C Bruce. 1984
Pliocene-Miocene
Louisiana 55OC - 120°C + 70 Weaver, 1979
( - 70°C) (50)
Pliocene-Miocene
Louisiana 78°C 160°C 82 Perry and Hower, 1972
Miocene
Louisiana 80°C 155°C 75 Bruce, 1984
Miocene
Galveston, Texas 100°C 138OC 38 Bruce, 1984
Miocene-Oligocene
Galveston, Texas - 55OC - 100°C Hower ei a/..1976
( - 80°C) ’ -130°C ’
Oligocene
Galveston. Texas 100°C 135°C 35 Bruce, 1984
Oligocene
Galveston, Texas 85°C 125°C 40 Perry and Hower, 1972
Eocene
Galveston (west),
Texas 100°C 142°C 42 Bruce, 1984
Eocene
West Texas 6OoC 120°C 60 Boles and Franks, 1979
’ Smaller values are for a very minor increase in I/S ratio.
’ Coarser fraction (0.5-2 am).

diagenetic reaction occurred. Unfortunately, these temperature values can at best be


considered estimates and any detailed conclusions based on them, or values from
other areas, must be viewed with skepticism.
Temperature values are based on Geothermal Gradient maps and corrected and
uncorrected well temperatures. The error can be as large as the range of values listed
in Table 7-1 (e.g., see Figure 2, Boyles and Franks, 1979). Measured temperatures
are generally lower than corrected temperatures. The I/S ratio varies with the “size”
fraction x-rayed and with interpretive technique (k15-20%, Srodoh, 1980). For any
given depth there should be a difference in the I/S ratio depending on whether
cores or cuttings (average value) are used. Picking the depth of the beginning of the
diagenetic reaction can be fairly subjective and influenced by sample preparation,
sensitivity of the x-ray, and prediliction of the observer.
The temperatures of first occurrence of authigenic illite layers range from 55’ to
100°C. The values correlate more with the authors than with age or location. The
top of the 4:l zone is reported to range from about 100” to 16OOC and the interval
424

over which the reaction occurs (zone of illite production, ZIP) ranges from 35” to
82°C. There is no strong correlation with age, but the temperature interval over
which the reaction occurs appears to be larger for the Louisiana wells than for the
Texas wells. Bruce (1984) believes this is not due to differences in temperature
gradients but to a difference in the composition of the detritus supplied to the two
areas. He notes that in the Texas area K-feldspar disappears rapidly within the
temperature range of 127” to 14OoC, whereas in the Louisiana area the disap-
pearance is more gradual, with the final loss occurring at 150°C. Bruce also
determined that, for the wells he studied, the time from the beginning of diagenesis
to the production of an I/S 4:l phase ranged from 2 to 5 m.y. and was not related
to geologic age.
Analyses of shale samples from several Gulf Coast wells (Weaver, 1979; Freed,
1980; Hinch, 1978) indicate that at depths between 3,000 and 3,400 m there is
commonly a “reversal” in the I/S trend. Samples in this interval commonly have 20
to 40% more smectite layers in the I/S phase than adjacent samples. Unfortunately,
information about the wells is limited, but these “reversal” samples appear to occur
near the top of the high pressure zone, which suggests high pressures may inhibit the
conversion of smectite to illite.
In the Paleocene-Cretaceous of west Africa, ZIP occurs over an interval of
approximately 40°C (50” to 90°C) (Dunoyer de Segonzac, 1969). In the Cretaceous
of British Columbia (Foscolos and Kodama, 1974), ZIP occurs over a temperature
range of approximately 60°C (80” to 140°C). In the Upper Mississippian Spring
shales of Oklahoma (Weaver) ZIP occurs over a temperature range of approximately
60°C (60” to 120°C). In many older papers it is assumed that I/S does not start to
form until the 17 A, e g , smectite, peak is lost. However, calculations (MacEwan et
al., 1961; Reynolds, 1967) indicate that expandable clay can contain up to 60% 10 A
layers (40% 17 A layers) and still produce a 17 A peak. Thus, diagenetic effects start
at shallower depths and lower temperatures than the depths at which the 17 A x-ray
peak disappears. In the Gulf Coast, diagenetic effects can be detected at tempera-
tures approximately 20” to 30°C lower than the temperature at which the amount of
layer collapse is sufficient to cause the 17 A peak to disappear.
In the Rhine Graben area of Germany, where geothermal gradients are high,
diagenetic processes start at relatively shallow depths (few hundred meters). In
Oligocene strata the maximum temperature at which discrete smectite disappears is
about 80°C; whereas, in the younger Miocene strata this temperature is as low as
40°C. Temperatures where the 4:l I/S ratio is reached appear to be - 140” and
llO°C, respectively. The differences in the alteration temperatures are believed to be
due to variations in K-feldspar content (the proportion of illite layers in I/S
increases as K-feldspar increases) and differences in the composition of the detrital
smectite (Heling, 1974, 1978; Doebl et al., 1974). Similar rates of illitization are
found in both marine and continental sediments.
Teodorovich and Konyukhov (1970) observed mixed-layer clays forming at
relatively shallow depths in the Jurassic and Lower Carboniferous sediments of
Russia. They believed this was due to the ready availability of K from biotite.
The temperatures for the top and bottom of the ZIP in other areas are similar to
425

Expandability ( p e r c e n t )

"
-
20
?
40 -
./ \
60 -
80 -

;j 100-
0
Y

aJ
L
J 120-
e
%
140-
I-

160 -
180 - I /
I
I
200 - I

Fig. 7-3. Smectite to illite transformation in shales from different sedimentary basins. plotted as a
function of temperature. Central Poland from Srodoh (19x4): Douala Basin (Cretaceous) from Dunoyer
de Segonzac (1969) as interpreted by Srodon and Eberl (1984); North Sea (Lower Tertiary through
Cretaceous) from Pearson et at.. (1982); Gulf Coast from Hower et al., (1976), and from Boles and
Franks (1979). Dotted lines represent two extreme profiles for profiles from the North Sea. The three
Gulf Cost profiles are liocene-Pleistocene (lower line). Eocene-Oligocene (middle line), and Lower
Eocene (upper line). From Srodon and Eberl. 1984. Rev. Min. 13. Mica, 495-516. Copyright 19x4 Miner.
Soc. Amer. Central Poland curve should be shifted down (Srodob. personal communication, 1988).

those in the Gulf Coast, though the values from the Rhine Graben may extend the
temperature range. The reported temperatures for the beginning of the ZIP range
from 55" to 100°C and depths range from 1200 to 2300 m. The base of the ZIP is
encountered at temperatures ranging from 90" to 160°C and at depths ranging from
2500 to 6000 m. Variations in the geothermal gradient and permeability presumably
account for some of the variations in the depth values.
On the basis of the data in Fig. 7-3, Srodoh and Eberl (1984) concluded that
reaction time plays an important role, as does rock and fluid composition (K), in the
formation of the I/S sequence. The figure indicates that for a given temperature the
426

proportion of illite layers in the I/S increases with age. Actually, considering the
amount of variation due to different procedures the curves cluster fairly well except
for the data for central Poland (Srodoh, 1984). In this latter well, illite formation
starts at < 20°C and the 4:l ratio is reached at 72°C; this occurs at a depth of
< 2000 m. The I/S is a minor component and K-feldspar is more abundant than in
the Gulf Coast. Sufficient K is available for the reaction to approach completion,
and apparently at a lower temperature than other regions. The only trouble is, this
well penetrates a section extending from the Upper Jurassic to the Carboniferous, a
section covering a depositional period of at least 150 m.y. It is unlikely that the
present well temperatures are the maximum temperatures to which the sediments
have been exposed. Either the geothermal gradient could have been higher in the
past or the area may have been intermittantly uplifted and eroded. (Srodoh, 1988
personal communication, wrote me that they have recently fond that 1.7 km of
Cretaceous rocks have been eroded and the I/S curve falls on trend with the others.
This illustrates the type of pitfalls that exist in interpretative shale petrology.)
In the Upper Cretaceous/Cenozoic marine shales of the Campos Basin, Brazil,
the development of I/S with depth lagged behind that in the Gulf Coast area. This
is considered to be due to the original composition of the Brazilian smectite, an
Fe-rich beidellite or nontronite derived from soils (Couto Anjos, 1984).
A regional study of the Cretaceous, Mancos shale from the Colorado Plateau and
southern Rocky Mountain region (Nadeau and Reynolds, 1981) indicates that in
general the I/S ratio increases with increasing depth of burial; however, the I/S in
-
calcareous shales tends to have fewer ( 40%) illite layers than adjacent noncalcare-
ous shales. Bentonites enclosed in calcareous shales also have far fewer illite layers
than those in noncalcareous shales. In this example the permeability decrease
caused by the calcite may account for the low reactivity in the calcareous shales.
Examples of K-bentonites have been described in which the I/S ratio appears to
be too high or low for the burial temperature. Huff and Tiirkmenoglu (1981)
described an Ordovician K-bentonite from the Cincinnati ash which has a 4:l I/S
ratio but has not been exposed to temperatures greater than 80°C (as compared to
temperatures of 120" to 160°C for similar I/S in the Gulf Coast). The authors
suggest the I/S ratio was determined by the composition of the original srnectite's
relatively high octahedral charge due to relatively high Mg content ( O S ) , rather than
temperature and presumably, time.
Numerous examples have been given which suggest that the availability of K
limits the number of illite layers that can form in bentonite. Closely spaced
K-bentonite beds may have different I/S ratios; the thicker beds commonly have
more smectite layers (Bystrom, 1956; Altaner et a/., 1984). Thick K-bentonite beds
(several meters) are frequently zoned with the I/S ratio decreasing from both edges
towards the middle (Srodoh, 1976; Velde and Brusewitz, 1982; Altaner et a/., 1984).
The K distribution, along with other data, indicates the K was derived from the
surrounding shales and was transported into the bentonite by diffusion or
metasomatic processes. The increase in K correlates with an increase in Al,
suggesting the A1 may also have migrated into the bentonite. It is also possible that
the A1 distribution within the K-bentonite was rearranged.
421

The development of I/S in a complex volcanic geothermal area illustrates the


limited relation of temperature to the S -,I reaction. Studies of the New Zealand
geothermal area by Eslinger and Savin (1973) show that the alteration of montmoril-
lonite to illite is thermally complex. In borehole Br16, smectite persists to a
temperature of approximately 150°C where it gives way to an I/S containing about
10% smectite layers. This material persists with depth until a temperature of 18OOC
is reached. N o expanded layers are present at temperatures higher than 180°C. In
borehole Br7 at a depth of 162 m (112°C) the I/S contains 15 to 20% expanded
layers; at 433 m (209OC) it contains 90 to 95% expanded layers. Illite develops at
750 m (240°C).
Authigenic illite, chlorite, smectite, and I/S with a wide range of ratios are
present in the Triassic volcanogenic sediments from the Southland Syncline, New
Zealand. The distribution of the physils is not related to burial temperature and
both illite and chlorite formed from volcanic glass through a solution/crystalliza-
tion mechanism and not from a smectite precursor (Ahn and Peacor, 1985).
Whether the conversion of smectite to illite is an equilibrium or kinetic process, i t
has been assumed that the conversion was gradual, occurring as the sediment pile
settled through the appropriate temperature window (commonly 50 to 150OC). K-Ar
analyses of Gulf Coast sediments by Weaver and Wampler (1970), Perry (1974), and
Aronson and Hower (1976) showed that with depth the K-Ar apparent age of the
bulk sample and the fine fractions decreased with depth as new illite layers were
formed in the I/S phase (Fig. 7-4). More recently Morton (1985) has obtained
Rb-Sr ages from a Brazoria County, Texas, well (Miocene-Oligocene) adjacent to
the well studied by Aronson and Hower (1976) (Fig. 7-4). Both the K-Ar and Rb-Sr
ages increase proportionally to the increase in the I/S ratio. The age values for the
two sets of ~ 0 . pm 1 samples are quite similar. The Rb-Sr ages of the 0.05 p m
samples are approximately 10 m.y. less than those of the 0.1 pm samples. Probably
some detrital illite is present in the < 0.1 pm samples. The K content of the samples
parallels the I/S curve; however, for some reason, the Rb content increases
continuously from about 3.5 km until total depth.
The abrupt decrease in the K-Ar and Rb-Sr ages of the <0.1 p m fraction
coincides with the rapid increase in illite layers, which presumably caused the rapid
addition of K from K-feldspar (Weaver and Wampler, 1970; Hower et al., 1976) to
the layer silicate phase; Ar was lost in the process and a new K-Ar system, and
Rb-Sr system, created. The shift (3,150 to 3,250 m) in the Rb-Sr age of the 0.05 p m
occurs approximately 500 m shallower than the shift for the coarser material. This is
probably the general trend, for a given I/S ratio to occur at shallower depths (lower
temperature) for fine-grained fractions than for coarser fractions.
In general it has been assumed that the conversion of 1:4 I/S to 4:l I/S occurred
gradually, over a period of several million years, as a function of increasing depth
and temperature. Perry and Hower (1972) proposed that it was a two-step process
with the initial illite layers (random) developing more slowly than the later illite
layers (regular) (Fig. 7-1). Morton (1985), on the basis of his Rb-Sr and oxygen
isotope data, concluded that the 1:4 to 4:l ZIP (zone of illite production) occurred
2300 to 2700 m shallower than at present depth and that diagenesis occurred
428

DEPTH
km PERCENT I L L I T E LAYERS (< 0 . 1 m )

'
0 20
1 1
I 1 1 I I I 1
I'
L
0 K-Ar<O.Ipm
Rb-Src0.05pm
Y Rb-Sr<O.lpm
L

10 20 30 40 50 D m.y.

Fig. 7-4. Percent illite layers in I/S and K-Ar and Rb-Sr apparent age as a function of depth in adjacent
wells in south Texas. I/S and K-Ar data Aronson and Hower, 1976; and Rb-Sr data from Morton, 1985.

suddenly (episodic, punctuated) when there was a change in pore water chemistry.
The change in pore water chemistry is believed to have occurred 23.6 m.y. ago, the
Rb-Sr isochron age of the deeper 0.05 pm samples. This is the approximate age of
the Miocene-Oligocene boundary. The proposed shallower depth of formation
would indicate I/S diagenesis started after a few hundred feet of burial, which does
not seem likely. Morton suggested that early diagenesis may have been caused by a
period of rapid burial and the upward flushing of fluids derived from the compact-
ing muds or a major lowering of sea level and a downward flux of meteoric water.
More data are needed.
One other point should be made with respect to the age data in Fig. 7-4. Once the
4:l I/S forms, at around 3,700 m, the age should increase as it becomes more
deeply buried. Actually, all three sets of age data indicate the apparent ages
decrease with increasing depth of burial. This suggests there is some minor, but
continuous, mobilization of K and Rb and partial recrystallization of the I/S as the
layers become more ordered.
In contrast to the relatively systematic increase in the I/S ratio with temperature
(and depth) in the Gulf Coast area, the I/S ratio of Miocene arkosic samples from
the San Joaquin Basin, California, shows little relation to depth (Ramseyer and
Boles, 1986). Fig. 7-5 shows the relation of the I/S ratio of San Joaquin samples
from 15 wells to present burial temperature. The trend is more horizontal than
429

-""
0 20 40 60 #) 00
SMECTITE LAYERS IN I / S CLAY (XI
Fig. 7-5. Present burial temperature vs. percentage of smectite layers in < 2 p m illite/smectite (1,'s)
from the San Joaquin basin. Ordered interstratification is based on presence of 002 reflection between
5.2" and 8.9'20. Miocene Gulf Coast data from Bruce (1984). and Gulf Coast Plio-Miocene data from
Perry and Hower (1972). From Rayseyer and Boles, 1986. Copyright 1986 The Clay Min. Soc.

vertical and shows little relation to the Gulf Coast trend. There are a number of
factors that affect the I/S ratio. The composition of the smectites is relatively
uniform and is typical of montmorillonite:

( .50Fe,3.T8Mg0.3X)(Si3.X5A10.15 )OIO (oH)2Ca0.10 K0.0XNa0.05

Part of the variation is due to difference in the I/S ratio of sandstones and
shales, but this is minor, 5-108 more smectite layers in sandstones. A few
sandstones are cemented with calcite, which has retarded the illitization process.
Other than the few cemented sandstones, permeability does not appear to be a
factor. High concentrations of Ca and Mg in the pore waters tend to slow the
reaction rates of the formation of I/S (Roberson and Lahann, 1981). Though the Ca
and Mg concentrations are relatively high in the San Joaquin Basin, they have not
inhibited the illitization reaction where conditions were suitable. The authors believe
that most of the variation in I/S ratio is due to differences in residence time at the
appropriate temperatures.
One group of Miocene samples is overlain by Pleistocene sediments < 900 m
thick; another group is overlain by 1000 to 2500 m of Pleistocene sediments. Thus,
the burial rate and residence time at a given temperature varies for the two groups.
For example, areas containing I/S with > 95% smectite had a residence time
between 100' and 120°C of 2.0 to 2.1 m.y.; whereas areas containing I/S with
- 30% smectite had a residence time, at these temperatures, of 2.5 m.y. The near
horizontal distribution of various I/S ratios (Fig. 7-5) leads the authors to conclude
430

that the change from highly expandable I/S to a slightly expandable I/S occurred
over a narrow temperature interval (10" to 20").

Mechunism
I t is still controversial as to whether the smectite-to-illite reaction is due to
transformation or neoformation. Is it controlled by temperature, kinetics, or com-
position? If temperature is the control factor, the reaction should be in thermody-
namic equilibrium and energy (heat) determines the properties of the system. If
kinetics is involved, then time is a factor and the process is controlled more by the
relative rate of reaction than by thermodynamics. These two factors are important
only if the ions necessary for the reaction are present.
Eberl and Hower (1976) conducted a series of hydrothermal experiments to
determine the rate of formation of illite from smectite. Experiments were conducted
at 260" to 393"C, with run times ranging from 3 to 266 days. The initial glass had

tion went as follows: glass smectite + I/S


--$ -
the composition of a beidellite (Al, (Si3,66A10,34)0,,~(OH)2(K,Na),,,,4). The reac-
I/S + kaolinite (or pyrophyllite) +
quartz. Expandability of the I/S was inverse to run time. The first step, conversion
of glass to beidellite, went to completion in several days. The second step takes
place over months or years, depending on the temperature. They found that
K-saturated Wyoming montmorillonite formed illite layers much faster than their
synthetic smectite. This may be due to the presence of Mg in the former and not in
the latter. The montmorillonite probably formed a true illite, whereas the synthetic
beidellite (Si, Al) presumably formed a muscovite-like phase. They also found that
the presence of Na slows the reaction rate.
Using first-order kinetic plots and the Arrhenius equation, they calculated that
the activation energy for the conversion of beidellite to I/S (or M/S) was 19.6 f 3.5
kcal/mole. The size of the activation energy and the rate constants suggest the
alteration involves the breaking of chemical bonds in the 2:l layers (dissolution-re-
precipitation process). They concluded that either an equilibrium or a kinetic
interpretation for the range of mixed layering found in burial diagenetic sequences
is compatible with the kinetic data. According to their data (reaction time vs.
temperature), smectite should convert to I/S 4:l in less than 1 m.y. at a temperature
of 80°C. The reaction rate is apparently much slower in nature (Table 7-1). One
reason may be that Eberl and Hower used water/solid ratios much larger than
would exist in mud-shales. They suggested that the presence of abundant Na in the
natural system might slow the reaction.
Hydrothermal experiments (Eberl, 1978) with Wyoming montmorillonite
saturated with monovalent and divalent cations demonstrated that the rate of
formation of collapsed layers (10 A) was inversely related to the interlayer hydration
energy (AGO) of the monovalent cations. Dehydration presumably results because
of the attraction between negatively charged 2:l layers and the positive interlayer
cation. The greater the hydration energy of the interlayer region, the greater the
negative charge must be to cause dehydration, and, therefore, the slower the
reaction rate for the divalent cation clays was directly related to the interlayer
hydration energy. Smectites saturated with cations of hydration energy equal to or
43 1

Table 7-2
lnterlayer hydration energy of selected cations.
Cation AGO( Kcal/mole)
K+ - 2.62
Rb+ - 2.66
cs+ - 2.90
Na+ - 3.91
Li+ - 5.59
Ba++ - 5.96
Sr++ - 6.00
Mg' + - 6.60
Ca++ - 1.42

greater than Na+ did not form collapsed layers when heated to 300°C. Some
hydration energy values are listed in Table 7-2. Only K + , Rb+, and Cs+ produced
collapsed layers when heated to 300°C. Other cations produced collapsed layers
when the samples were heated at 400°C. Calculations (Eberl, 1980) indicate that a
layer charge of -0.77 (per O,,(OH),) is required to fix K + and a charge of -0.86
to fix Na'.
When smectites were treated hydrothermally in solutions containing various
mixtures of K, Na, Ca, and Mg it was found that the reaction rate and the rate of
ordering of I/S was inhibited by the addition of Na + , Ca++, and Mg"; the
inhibitory strength, on an equivalent basis, increased approximately in the ratio
1:10:30 (Roberson and Lahann, 1981). The authors suggest the presence of C a + +
and Mg++is the main cause of the slower kinetics in the natural system. Eberl et a / .
(1978) note that not only d o Na, Ca, and Mg block the formation of illite layers, but
so does increased water pressure.
The experiments (Lahann and Roberson, 1980) further showed that silica in the
reaction solutions frequently reached a maximum value in the first 24 hours. The
silica concentration appeared to vary independently of the formation of 10 A layers.
The activation energy for silica dissolution was on the order of 10 Kcal/mole,
whereas illite formation had an activation energy on the order of 30 Kcal/mole.
Apparently, in the first 24 hours of an experiment a silica-depleted montmorillonite
formed. With time, octahedral Al migrated into tetrahedral sites, creating a high
layer charge and fixing the K in the interlayer space (activation energy 30
Kcal/mole). The process is a solid-state reorganization as opposed to solution-re-
precipitation (Eberl and Hower, 1976). As shown by Weaver and Pollard (1973, Fig.
15), for a given amount of total Al illite contains less octahedral Al than montmoril-
lonite. The different outcome of the two experiments is in part due to the fact that
Roberson and Lahann's solutions were pH-buffered (pH 6 and 7), which minimized
the solubility of AI3+and prevented acid dissolution of the smectite. The final pH in
the Eberl and Hower experiments was in the range of 4 to 5 and some kaolinite was
produced. The former experiment is more likely to have simulated natural condi-
tions. Also, the presence of excess Na, Ca, or Mg in solution may have inhibited
A13+uptake from solution. In nature illite might form by either solid-state reorgani-
432

zation or dissolution-reprecipitation. The latter process is more likely to occur in


porous sandstones (Chapter VIII).
Lahann and Roberson (1980) also found that silica dissolution from smectite was
most rapid when K f exchange ions are present; increasing the concentration of K +
accelerated the silica removal, whereas increasing Na+, C a f t , and N a + + inhibited
silica removal. They suggested that in the natural system K + uptake by montmoril-
lonite may precede and accelerate illite formation. Early K + uptake was observed by
Ahn and Peacor (1986) in buried Gulf Coast smectites.
Also, as discussed in Chapter VIII, CO, is generated by the thermal maturation
of kerogen at temperatures in the range of 80" to 120"C, with peak generation
occurring at approximately 100°C. The acid conditions created by the CO, should
increase the solubility of K-feldspar and facilitate the growth of illite layers. At
these same temperatures organic acids are also generated which are capable of
complexing and transporting Al. Thus, variations in the organic content could have
an effect on the rate of conversion of smectite to illite.
It is of interest to note the results of hydrothermal treatment of a variety of
igneous rocks, e.g. dacite, rhyolite, basalt, volcanic ash, etc. (Hawkins and Roy,
1963). The various rocks and glasses were heated in saline, acid (HCl), basic, and
C0,-saturated waters. In the first three solutions montmorillonite and/or zeolite
were produced; the presence of Mg favored the formation of montmorillonite. In
the demineralized carbonic acid waters (pH 3.9) all the glasses (melted and quenched
rocks) formed illite. Presumably this is a result of high K concentrations in solution.
Hower and Altaner (1983) and Thompson and Jennings (1985) concluded that
the smectite to illite reaction is kinetically controlled in the I/S compositional range
of 0 to 80% illite layers, at which point the I/S attains the RI interstratified
-
structure. For long reaction times ( > lo7 yr) the RI structure is reached at
90-100°C; for short reaction times ( < lo6 yr) the RI structure is not reached until
temperatures of 130" to 150°C; in either case, independently of reaction time, the
RI I/S persists to - 175OC, above which temperature the further reaction towards
illite and R 2 3 structure may be equalibrium dominated. The final smectite layers
-
may persist to 350°C (Weaver and Associates, 1984). Our limited data suggest the
R 2 3 stage may also be kinetically or time controlled. The illite in deeply buried
lower Paleozoic shales of the Anadrarko Basin, Oklahoma, have a K.I. of 3.0
(boundary between diagenesis and anchizone). Present-day well temperatures, at the
depth of the samples, are 230" to 250°C. However, the temperature could have been
higher in the past. Illites in Lower Cretaceous samples from a deep Texas well have
similar K.I. values but a well temperature of 300" to 310°C. Both these illites
contain several percent smectite layers.
More recently, high magnification transmission electron microscopy (TEM) has
been used to study the layer-by-layer arrangement of the physils. These observations
provide additional insight into diagenetic processes. TEM not only permits direct
observation of the grain morphology but also permits direct resolution of images of
individual layers of structure. The technique is particularly useful in showing
changes in crystal perfection, dislocations and layer heterogeneity during diagenesis.
Samples are commonly prepared by making thin sections perpendicular to bedding.
433

Fig. 7-6. High-magnification lattice fringe image from the 1750-m depth of a Gulf Coast well. Sample
shows typical discontinuous and wavy smectite layers having variable orientation. Edge dislocations are
indicated by arrows. From Ahn and Peacor, 1986. Copyright 1986 The Clay Miner. Soc.

A portion of the thin section is further thinned in an ion mill in order to obtain
electron-transparent thin edges (Blake and Peacor, 1981).
An excellent series of TEM pictures of Gulf Coast mud-shales showing the
development of illite layers with depth have been published by Lee et al. (1985) and
Ahn and Peacor (1986a). One of the major disadvantages of the TEM is that the
sample must be examined in a vacuum. This very commonly causes the smectites to
lose their interlayer water and collapse to 10 A, thus making it difficult to dis-
tinguish from illite. Smectite lattice fringed images can be identified by their general
irregularity and variation in width from 10 A to 13 A along one layer; subparallel
layers are generally thin and discontinuous; layers are curved with a general wavy
appearance; layer terminations are common (Fig. 7-6). In contrast, illite layers
appear as straight lattice fringes with constant 10 A interlayer spacings that are
continuous and straight over considerable lengths; layer terminations are relatively
rare; packets of layers have a “mottled” contrast (Fig. 7-7). Analytical electron
microscopy (AEM), which can be used to obtain chemical analyses of the layers,
provides a definitive method of characterizing the various layers.
Fig. 7-6 is a typical low-magnification TEM image of smectite from a depth of
1750 m from a Gulf Coast well. X-ray analysis indicates the predominant physil at
434
435

this depth is I/S with 20% illite layers. The general state of imperfection is typical
of smectite in the early stages of diagenesis. The anastomosing lattice fringe images
and lens-like shape are characteristic of smectite. These packets are appawntly clay
crystals rather than aggregates.
At a depth of 2450 m, where the I/S contains 40% illite layers, 50-100 A thick
illite packets occur within smectite (Fig. 7-7). Illite layers are parallel to the
surrounding smectite layers but locally are characterized by low-angle grain-
boundary-like features. Smectite is present to the left of the illite packet; however,
the smectite and illite layers are not continuous and there is an “along-layer”
boundary between smectite and illite.
At 5500 m depth the I/S has a 4:l (x-ray) ratio typical of deep diagenesis
samples. TEM images (Fig. 7-7) indicate illite packets are wider and thicker than
those in shallower samples. Separate packets of illite appear to be relatively
defect-free, but each has a slightly different orientation, presumably inherited from
the clusters of precursor smectite layers.
Bell (1986) treated Gulf Coast I/S samples with dodecylamine, which moves into
the interlayer position of 2:l physils and expands the layers in proportion to the
layer charge. Thus, a d(001) spacing of 13.5 A indicates a low-charge montmoril-
lonite ( < 0.6 per O,,(OH),), 17.5 A indicates a high-charge montmorillonite (0.6 to
0.8 per O,,(OH),), and > 21.5 A indicates a very high charge montmorillonite or
illite ( > 0.8 per O,,(OH),). TEM lattice images indicate that continuous variations
in layer charge in adjacent and within individual layers are common. Fig. 7-8
illustrates some of the combinations of illite and smectite layers that were observed.
The associations of edge dislocations (Fig. 7-8) with the boundary between illite and
smectite along individual layers suggests that much of the illite growth proceeded
laterally. As the smectite structure was destroyed or altered, the illite structure grew
in its place.
Chlorite is believed to form as a by-product of the conversion of smectite to illite
(Hower et al., 1976) and TEM lattice fringe images tend to confirm the idea (Ahn
and Peacor, 1985). Fig. 7-9 shows packets of chlorite interstratified with packets of
10 A illite layers (2450 m sample). The chlorite packets consist of interstratified 7 A
and 14 A layers. AEM data indicate both have a similar composition, Fe-rich. The
14 A phase has the following structural formula:
(A11.5 Fe3.1Mg1.1(Si 2.8 A11.2 P I 0 (OH)*
The 7 A layers are apparently berthierine. It is unlikely the chlorite formed by
solid-state transformation of smectite or kaolinite; more likely, considerable solu-
tion, diffusion, and recrystallization was involved. Much of the Si, Fe, and Mg was
derived from the smectite. A1 was presumably derived from feldspar and/or
kaolinite. The 7 A chlorite is normally a low-temperature phase that converts to the

Fig. 7-7. Top: Hi@-magnification lattice fringe image from the 2450-111depth of a Gulf Coast well. lllite
occurs as 50-100 A packets between subparallel smectite layers. Bottom: Abundant, thick, subparallel
illite packets locally coalesced in the 5500-m depth sample. From Ahn and Peacor, 1986. Copyright 1986
The Clay Miner. SOC.
436

I ,

D ///////

Fig. 7-8. Types of interfaces between smectite layers (striped) and illite layers (open) after saturation with
dodecylamine. ( A ) conservative boundary, i.e., no change in the number of lattice planes; ( B ) non-con-
servative boundary: bifurcating structural layers with a net loss in the number of lattice planes. (C) and
(D) non-conservative boundaries; terminating structural layers with a net loss in the number of lattice
planes. (E) combination of non-conservative houndaries, i.e., complex of edge dislocations where smectite
layers are contributing components to growing illite layers. From Bell, 1986. Copyright 1986 The Clay
Miner. Soc.

14 A polymorph at higher temperatures. The 14 A chlorite in the deepest sample


(5500 m) was the low temperature Ib;, polytype and all 7 A layers were not
converted to 14 A.

Fig. 7-9. Lattice fringe image by transmission electron microscopy obtained from the 2420 p m Gulf
Coast sample packet of 1 0 A illite layers interstratified with chlorite packets. Several 7 A layers are
interstratified with chlorite layers. From Ahn and Peacor. 1985. Copyright 1985 The Clay Miner. Soc.
431

AEM analyses (Ahn and Peacor, 1986a) of the shallow (1750 m) Gulf Coast
smectites indicate they have a highly variable composition, in keeping with their
detrital origin. Tetrahedral A1 is relatively high, commonly between 0.20 and 0.35
per O,,(OH),. The major change between the shallow smectite and the 5510 m illite
is an increase in total Al, tetrahedral Al, and K; and a decrease in Fe. The loss of Fe
relative to Mg is reflected in the high Fe content of the diagenetic chlorite layers
that are interstratified with the I/S.
Na and K are the major interlayer cations in the smectite, and in most layers K is
more abundant than Na. Some K apparently replaces N a before the layer charge is
increased. When competing with Na, K preferentially occupies interlayer sites
(Gast, 1969), particularly the higher charge sites (Marshall, 1955), and the prefer-
ence for K increases very rapidly with decreasing PH,o (Tabikh et a/., 1960).
Based primarily on their TEM and AEM studies; Ahn and Peacor (1986) have
proposed a reasonable mechanism for the smectite-to-illite reaction. Illite layers are
parallel or sub-parallel to adjacent smectite layers, but in general individual smectite
layers do not grade laterally into illite layers. With depth illite packets thicken and
become more abundant as smectite is “consumed”. The sequence of structural
patterns suggests that illite formed by replacement of smectite without any change
in volume.
The pictures indicate that illite formation started at sites within smectite rather
than at the boundaries of smectite packets (Fig. 7-10), as might be expected. As the
smectite is heterogeneous in structure and chemistry, it is possible the illite nucleates
at sites of unusually imperfect structure (e.g., dislocations) or where there are high
concentrations of K and/or Al. As the boundary between smectite and illite is
discontinuous in chemistry, structure, and texture, the formation of illite requires
considerable reconstitution of smectite.
As diffusion and interchange of Si and A1 do not occur in the solid state at low
temperatures, discontinuities in composition, structure, and texture at the smectite-
illite interface requires disruption, disarticulation, and reconstruction (at least in
part) of T-0-T layers. Disruption of the structure does not require that all
cation-anion bonds be broken; it is likely that major portions of the octahedral
sheet and condensed tetrahedral units retain their integrity during the transition.
Most of the K is obtained from an external source (K-feldspar and/or mica) and an
appreciable proportion of it replaces Na in smectite prior to illite conversion. The
rate-determining step is presumably the sluggsh Al-Si diffusion and local reconver-
sion of T-0-T units.
A pathway for diffusion through smectite must exist. Smectite characteristically
has a high density of layer terminations (Fig. 7-6). Where adjacent layers bend
around these layer terminations (edges), a channel about 10 A in diameter must be
present. These channels could serve as conduits for cations and water.
The proposed mechanism for the reaction of smectite to illite of Ahn and Peacor
is basically intermediate, but more specific, between that of Weaver and Beck (1971)
and Hower er al. (1976) (retention of the basic structure of the 2:l layers) and Boles
and Franks (1979) (partial solution of original smectite layers).
In the Gulf Coast, where smectite and illite coexist in direct contact, as individual
438

Fig. 7-10. (a) Lattict fringe image from 2450 m depth sample showing thin illite packet growing within
subparallel smectite layers. Illite layers are not continuous with smectite layers along layer direction as
shown in both circles. (b) Thin packets of illite layers oriented at large angles to one another. From Ahn
and Peacor, 1986. Copyright 1986 The Clay Miner. SOC.

phases and over a wide range of temperature, smectite illite cannot be an +


equilibrium assemblage. The presence of high concentrations of imperfections and
the existence of complex layer stacking relations are consistent with the metastabil-
ity of the two phases. The degree of reaction of smectite to illite must be controlled
by kinetic factors, i.e., temperature, time, activities of chemical components, and
rock/water ratio. In a sequence such as the Gulf Coast, where all other variables are
relatively constant, temperature may be the principal factor controlling the I/S
ratio. (The use of I/S ratios as paleotemperature indicators must be used with
caution.) Ahn and Peacor (1986a) further suggest that the coexistence of smectite
and illite and the abundance of structural imperfections at great depth is due to the
relatively high rock/water ratio, which inhibits transport.
The TEM data and x-ray data are in conflict (Ahn and Peacor, 1986). X-ray
diffraction patterns of the Gulf Coast samples indicate most of the illite and
smectite occur as a mixed-layer phase (interstratification occurs at the layer or
several layer level). In contrast, the TEM data indicate illite exists primarily as
439

packets of layers 10 to 30 layers thick within a matrix of smectite packets. When


samples are prepared for x-ray analysis, they are highly dispersed and only the
fine-fraction is used. It is possible that during preparation the illite and smectite
packets are cleaved into thin particles one and two layers thick. When these thin
units are sedimented on a glass slide they form a mixed-layer arrangement (Nadeau
et al., 1984) (see page 59). Thus, x-ray diffraction patterns may only indicate the
ratio of illite to smectite and not the interlayer stacking relations.
On the other hand, mixed-layer physils d o exist. Lattice fringe images of rectorite
and K-bentonite (Klimentides and Mackinnon, 1986; Ahn and Peacor, 1986b) show
they are composed of regular 1 : l alternations of mica-like and smectite-like units.
When the rectorite is ground it cleaves preferentially along the more weakly bonded
smectite-like layers forming two-layer units. When these units are sedimented the
rectorite structure is reconstituted (Ahn and Peacor, 1986b).
Fig. 7-11 show the change in morphology, as seen with the SEM, as smectite in
bentonites is converted to illite. The morphology gradually changes, with increasing
I/S ratio, from the typical honeycomb habit of smectite to the platy shape of illite
(Keller et af., 1986).
The preceding discussion described the reaction during burial of relatively
low-charged smectite layers, largely derived from volcanic material. Expanded 2: 1
physils are also formed by leaching of K from illites and I/S. In the COST no. 1
well (Recent through Miocene) in the south Texas Gulf Coast the detrital expanded
material is described as a K-deficient I/S. At a depth of 1000 m the I/S has a 3:7
ratio; when the sample is placed in 3N solution of KCI at 60°C for 24 hr, the ratio
shifts to 3:2. This indicates high-charged, vermiculitic layers are present but there is
insufficient soluble K in the natural muds to cause the high-charged layers to
collapse. The differential (30%) decreases systematically to a depth of 2750 m
( - lOO"C), where both the natural samples and the K-treated I/S have a 7:3 ratio.
In this instance the primary function of temperature was to increase the dissolution
of K-feldspar and increase the amount of available K. Over the same interval there
was some conversion of low-charged 2:l layers to high-charged layers by increasing
tetrahedral Al.
My own analyses of samples from this well indicate the shallow I/S has around
40% illite layers (COST no. 2 I/S has 50 to 55% illite layers) as compared to the
- 20% reported in most Gulf Coast wells. The difference is apparently a function of
age rather than geography. The upper portion of the COST wells contains Pleisto-
cene age sediments. Apparently during the wet Pleistocene, more stripped illite from
the Paleozoics to the east and north was transported to the Gulf than in the drier
periods. One cannot assume a homogeneous detrital physil suite has been delivered
to a basin even if the geography has remained relatively constant.
High-charged, vermiculitic layers capable of fixing K are also present in the
Mississippi Delta area but, in the one well studied, persisted to only 2100 m, where
they presumably converted to illite (Weaver and Beck, 1971).
Velde (1984) noted that the layer charge of smectite layers in I/S from different
areas can vary from 0.3 to 0.7 and suggested the differences are related to bulk
chemistry, possibly silica activity, or pH.
440

Weathered illite is a major component of the Tertiary section in the Beaufort Sea
area (Dean, 1985). The stripped illite layers are both high-charged (vermiculite) and
low-charged (smectite); some hydroxy interlayer material appears to be present. The
proportion of expandable layers decreases with depth, with the vermiculitic layers
converting to illite first. Foscolos and Powell (1978) suggested that in a Mesozoic
shale section from the Sverdrup Basin, Canadian Arctic Islands, the vermiculite in
the I/S/V phase formed from smectite (increase charge) during burial and was in
turn altered to illite (fixed K). The mixed-layer physil in a Lower Cretaceous
diagenetic sequence in British Columbia is also a three-component (I/S/V) material
(Foscolos and Kodama, 1974).

Chemistry
There is only a minor amount of reliable data on the chemical composition of the
2: 1 physils involved in the continuous illitization of smectite during burial. Analyses
of I/S physils, primarily K-bentonites, indicate the amount of K and total charge
tend to increase with increasing I/S ratio (Table 2-16). Apparently, because of the
heterogeneity of the starting materials, there is no well developed trend, such as in
the amount of tetrahedral Al or the composition of the octahedral layer.
Chemical analyses of bulk samples from Gulf Coast wells indicate the chemistry
remains relatively uniform with depth, indicating the thick physilite sequence was
essentially a closed system. This is not entirely true as some ions in the interstitial
waters, particularly Mg, Na, CO,, etc. moved from the physilites into the sand-
stones. Chemical analyses of the fine fraction. primarily I/S, show an increase in A1
and K with depth (Weaver and Beck, 1971; Hower et a/., 1976). AEM analyses of
smectite packets in shallow samples and illite packets in deep samples (Ahn and
Peacor, 1986) more accurately reflect the change in chemistry with depth. The
detrital smectites are quite heterogeneous in composition; the average composition
and range of 19 packets is:
(A'l.4XFe0.34Mg0.1X)(si3.75A'0.25 )o,"(oH)2K,,,6Na".13
1.16 - 1.74 0.04 - 0.31 0.12 - 0.34 0.00 - 0.31
0.21 - 0.63 0.00 - 0.25
The composition of the deep diagenetic illite is more uniform. The average of 9
packets is:
(A'l.69Fe0.19Mg0.12)(si3.40A'0.60 )010(0H)2K0.63Na,,.(~l
1.46 - 1.85 0.00 - 0.29 0.52 - 0.69 0.52 - 0.72
0.10 - 0.28 0.00 - 0.06
Samples used for wet chemical analyses include both illite and smectite layers
and are an average of the results produced by electron probe analyses, but show the
same trend (Foscolos and Kodama, 1974; Hoffman and Hower, 1978; Weaver,
1979; Srodon et a]. 1986).
On the basis of several assumptions Boles and Franks ( 1 979) showed that Al-rich
smectite layers convert to illite before Fe- and Mg-rich layers.
441

Fig. 7-11. SEM illustrating the morphological changes that occur as the proportion of l O A layers
increase. (A) Na-montmorillonite, clay spur, Wyoming typical “cornflake” texture. (B) Bentonite from
Mancos Shale containing 65% smectite layers. (C) Bentonite from Mancos Shale containing 45% smectite
layers. (D) Bentonite from Mancos Shale containing 30% smectite layers. Well-developed flakes have a
morphology similar to illite. (E) Tioga K-bentonite, New York, containing 10% smectite layers. (F) Tioga
K-bentonite, Virginia. Sample is essentially pure illite. Well developed, parallel plates are scalloped and
slightly curled at the edges, Bars = l a m . From Keller et a/., 1986. Copyright 1986 The Clay Miner. Soc.
Courtesy W.D. Keller.
442

THOUSAND
3;

200 300 m.y. K / A r AGE BULK SAMPLE


Fig. 7-12. Graph showing the change in K content and K-Ar apparent age with depth of the > 2 p m and
< 0.2 p m size fractions of samples from a Gulf Coast well. Asterisk = age of total sample. Data from
Weaver and Warnpler (1970).

Analyses of sequences of K-bentonites (Silurian, Wales; Cretaceous, Montana;


Upper Carboniferous, Poland) (Srodoh et al., 1986) show similar trends with respect
to K, layer charge, and I/S ratio. The amount of fixed K per Olo(OH), layer
average 0.55 for samples containing less than 50% illite (random); I/S with more
than 50% illite (ordered) have concentrations of about 1.0 per illite layer (Fig. 2-24);
the average for - 100% illite is 0.75 K per illite layer. The total layer charge has a
similar trend. The layer charge increases from about 0.4, for 100% smectite, to 0.5
for I/S with a 1:l ratio. The final 50% of layers have a layer charge of about 1.0 for
an average near 0.8 for illite. The octahedral charge remains relatively constant (0.2
to 0.3) and the increase in layer charge is primarily due to an increase in tetrahedral
Al, as it is in the non-bentonitic smectites.
As discussed earlier, K-feldspar is the primary source of K for illitization of
smectite. Texturally this indicates a shift in the K content from the coarse to the fine
fraction. Fig. 7-12 shows the systematic decrease in the K content of the > 2 y m
fraction and an equivalent increase in the K content of the < 0.2 pm fraction with
depth. The K-Ar apparent age of the > 2 y m fraction remains constant, whereas
that of the < 0.2 pm fraction decreases (Weaver and Wampler, 1970). As a result
the K-Ar age of the total sample decreases.
443

Chlorite
Authigenic or neoformed chlorite is a relatively common constituent of sand-
stones and to a lesser extent limestones; it is discussed in Chapter VIII. Secondary
chlorite is common in anchizone and epizone physilites and will be discussed later in
the chapter. Because of the presence of detrital chlorite in most shales it is difficult
to determine how much authigenic or diagenetic chlorite is present. It is apparent
from the preceding discussion that chlorite can sometimes, but not always, form as a
by-product of the smectite -,illite reaction.
Chlorite and Ch/S, including corrensite, are common in evaporitic rocks, includ-
ing dolomites; they are generally Mg-rich (p. 408). I t is not clear whether these
physils formed syngenetically or diagenetically. It appears that the Mg-rich varities
can form in hypersaline fluids without the need for elevated temperatures. On the
other hand, both chlorite and Ch/S form at elevated temperatures, commonly from
basic volcanics. Corrensite can persist to temperatures on the order of 300°C.
A
Low-temperature chlorites are commonly the 7 variety (berthierine) or the 14
Ib polytype. With increasing temperature they convert to the 14 A IIb polytype (p.
485). In most instances the chlorites are such a minor component of shales that the
polytype or 7 A phase cannot be determined by x-ray analysis; however, TEM
techniques can be used to identify the chlorite phases (Fig. 7-9).
Hower’s et al. (1976) data (Fig. 7-13) indicate chlorite first occurs at 2450 m or
about 70OC. TEM studies (Ahn and Peacor, 1985) indicate some of the chlorite is

Percent Chlorite
Fig. 7-13 Chlorite content of two size fractions of “shale” samples from the same Gulf Coast well
reported in Fig. 7-1. From Hower et al., 1976, Geol. SOC. Amer. Bull., 87. 725-737.
444

intergrown with illite and I/S (but could still be detrital) and occurs as both the
A A
14 and 7 Fe-rich varieties (Fig. 7-9). At 2450 m, where the I/S ratio is 3:7 and
I/S diagenesis has just started, the total sample contains 5.7% chlorite (x-ray). The
chlorite content remains constant through the interval where the proportion of illite
layers is increasing from 30% to about 60%; the chlorite value increases to 7.5% in
the deeper samples where the I/S ratio is 4:l. Chlorite as percent of total physils in
sample systematically increase from 6.7% in the 2450 m sample to 20.8% in the I/S
4:l samples. In all but the shallowest sample approximately half the chlorite is in
the > 10 pm fraction. This suggests it is either detrital or grows much larger than
the diagenetic illite. The first occurrence of chlorite approximately coincides with
the Miocene-Oligocene boundary, which could be coincidence or could indicate
some (or all) of the chlorite is detrital and reflect a change in climate, source, or
environment.
Detrital chlorite, in minor amounts, is present in most Cenozoic mud-shales of
the Gulf Coast and most studies d o not report any significant increase with depth,
though it has been observed in other wells. Our analyses of the < 2 p m fraction of
shale samples from the COST no. 1 well offshore Texas indicate the kaolinite
(5-6%) and chlorite (2-3%) remain nearly constant from 400 to 4000 m. Below this
depth, which is where the I/S reaches a ratio of approximately 4:1, the
kaolinite/chlorite ratio systematically decreases; at 5200 m the suite contains 3%
kaolinite and 6% chlorite and the I/S ratio is 4:l. The distribution suggests both
smectite and kaolinite are involved in the formation of chlorite.
Earlier Fuchtbauer and Goldschmidt (1963), Dunoyer de Segonzac (1969), and
others observed that in thick sedimentary sections kaolinite decreases and trioc-
tahedral chlorite increases with depth and temperature, and they suggest that
kaolinite alters to chlorite. In the Jurassic shales of Lower Saxony, Fuchtbauer
(1967) found that chlorite increased at the expense of kaolinite. Weaver and Beck
(1971) showed that as kaolinite decreased with depth in the smectite-rich Carbonif-
erous Springer shales, chlorite, possibly dioctahedral, increased. The thermal stabil-
ity of the chlorite increases with depth. The shallow chlorite commonly contains
some vermiculite layers.
Sarkisyan (1972) reported that in the Caucasus corrensite (50% expanded layers)
coexists with mixed-layer illite-montmorillonite containing 25 to 30% expanded
layers. This suggests that the temperature to which the rocks were exposed was at
least 120OC. In the Lower and Middle Ordovician shales and shaley limestones of
Newfoundland, corrensite is believed to have formed diagenetically from volcanic
detritus (Suchecki et al., 1977). The associated mixed-layer illite-montmorillonite
contains only 10 to 15% expanded layers. This suggests that the temperature to
which these sediments have been exposed is in the range of 170' to 220'C. In
metamorphosed Alpine rocks corrensite is present in rocks containing well-crystal-
lized illite (Kubler, 1968).
The thermal stability range of corrensite exceeds that of I/S; in sandstones it can
form at shallow depths and can persist to temperatures in excess of 200°C perhaps
as high as 300°C. Suchecki et al. (1977) suggest that this is due to the trioctahedral
character of the chlorite-montmorillonite. The hydrogen bonds of the octahedral
445

OH ions are oriented perpendicular to the clay layer, allowing for greater electro-
static repulsion between interlayer cations and the layer.
In felsic tuffs in the Neogene of Japan (Iijima and Utada, 1971), montmorillonite
changes to corrensite at approximately 100°C (analcime zone) which in turn alters
to chlorite at approximately 120°C (albite zone).
Kossovskaya et al. (1964) report that diagenetic chlorite and illite are abundant
in the floor of semianthracite coal seams in the Petchora Basin. In Australia, Kisch
(1966b) reported the presence of abundant Fe- and Al-rich diagenetic chlorite
(along with mixed-layer illite-montmorillonite-chlorite)in tonsteins associated with
coal of semianthracite rank (90-91.5% carbon). The chlorite is believed to form
from kaolinite when Mg and Fe are present. The Fe is believed to come from
siderite, which, through the loss of Fe, is converted to ankerite (Kisch, 1968). When
Mg and Fe are not present, the kaolinite becomes better crystallized or converts to
dickite. Semianthracite coal is believed to have been subjected to temperatures on
the order of 200°C (Demaison, 1974). A minor amount of chlorite is present in
tonsteins associated with coals having 89% carbon ( - 150°C). The presence of
mixed-layer illite-montmorillonite with (001) spacings ranging from 10.2 to 12 A
confirms that the rocks were probably not exposed to temperatures much greater
than 2OO0C, and further demonstrates that diagenetic chlorites form at lower
temperatures than pure illite (100% 10 A layers).
In a study of the K-bentonite beds of Sweden, Bystrom (1957) found that the
content of chloritic material increases as the tectonic deformation of the beds
increases. In the Appalachian region, K-bentonites commonly contain chlorite
and/or mixed-layer Ch/S. At least some of the chlorite is dioctahedral (Weaver,
1959). Chloritic material has not been reported in the equivalent age K-bentonite
beds from the mid-continent. These data suggest that the formation of chlorite is
related to the depth of burial or temperatures. However, in the Appalachians
chlorite may be present in one K-bentonite bed and not in another a few feet away
in the same outcrop (Weaver, 1953). Something other than temperature must be
involved. A chemical analysis of one of the chlorites from the Appalachian region
indicates that it contains less than 2% Fe,O, and is not the typical low-temperature
Fe-rich chlorite. In addition to temperature, the availability of Mg in adjacent rocks
may be a factor in the diagenetic formation of the chlorite associated with the
K-bentonite beds.
In general low-temperature chlorites tend to be Fe-rich, other than in evaporites;
the Fe/Mg ratio generally decreases with increasing temperature of formation.
In the Salton Sea hydrothermal area, Muffler and White (1969) observed that
Mg-rich chlorite starts to form at temperatures of 130" to 165°C. In the Wairakei
thermal area (Steiner, 1967), Fe-chlorite starts forming from montmorillonite at
temperatures of less than 100°C. Eslinger and Savin (1973) found that the chlorite
stage was preceded by a corrensite stage. Corrensite was observed at 112°C
(shallowest sample examined) and chlorite at 14OOC. Corrensite and swelling
chlorite are present in the basaltic rocks of the geothermal area on Reykjanes,
Iceland, at temperatures ranging from 150" to 280°C. At lower temperatures, only
montmorillonite is present. Normal chlorite was found in one well in the tempera-
446

ture range of 220" to 290°C. Mixed-layer chlorite-smectite dominates at 200" to


230°C (Thomasson and Kristannsdottir, 1972). In the Salton Sea geothermal field,
California, chlorite first appears at about 190°C and converts to biotite at about
+
325°C. The Fe/Fe Mg ratio decreases from 0.70 to 0.44 with increasing tempera-
ture (McDowell and Elders, 1980).
In a TEM/AEM study (Yau and Peacor, 1984) of well samples from the Salton
Sea area, diagenetic chlorite was first detected in samples where the well tempera-
A A
ture was 150°C. The chlorite consisted of interstratified 14 and 7 layers. The
A
7 phase decreased with depth. Chlorite was the dominant physil between 260"
and 310°C; at higher temperatures it converted to biotite. The temperature range of
maximum chlorite development is similar to that found in very low-grade physilites
(Weaver and Associates, 1984). AEM analysis of relatively high temperature chlo-
rites from the Salton Sea ( - 290°C) indicate they are Mg-rich:

(Mg3.2Fe1 .9A10.9Ca0.1 )(si3.1 A10.9 Iol0 (OH)*


The Mg value is appreciably higher than that reported (2.0-2.5) by McDowell and
Elders (1980) based on microprobe analyses. Chlorite and mixed-layer chlorite-
montmorillonite are commonly formed in wall rock adjacent to hydrothermal veins
and igneous intrusions (Sudo et al., 1957; Bundy and Murray, 1959; Chen, 1972;
Harvey and Beck, 1960; Blatter et al., 1973; Shirozu, 1974; Peters and Hofmann,
1984). The chloritic material is generally at the outer (cooler) rim of the alteration
halo. In the low-grade metamorphic portion of the Ouachita and southern Appa-
lachian systems, thin veins of pure-Mg-rich chlorite are abundant (Weaver, 1961;
Weaver and Associates, 1984).

Kaolinite

It has been convincingly demonstrated that in sandstones kaolinite can crystallize


during both shallow and deep burial; in turn, it can be altered to chlorite or illite
when the water chemistry changes from acid to alkaline (Chapter VIII). There is
little, if any, data to show that kaolinite develops in shales during burial, though this
may happen in physilites adjacent to coal or other organic-rich beds where apprecia-
ble CO, is generated. Kaolinite is commonly present in veins in coal and in pyrite
clusters in both coal and organic-rich shales. The kaolinites presumably formed at a
considerable depth.
C 0 2 generated from organic material in shales by thermal maturation is consid-
ered to be responsible for an increase in acidity in enclosed sands and the solution
of feldspars and carbonates and the formation of kaolinite (Chapter VIII). Under
some conditions one might expect that these CO, waters would cause the formation
of kaolinite in shales, particularly fresh water shales and mudstones. This reaction
likely occurs but has not been recognized.
Based on the kaolinite/gibbsite stability fields, Curtis and Spears (1971) con-
cluded that there is enough soluble silica (10 to 40 ppm) in dilute interstitial waters
to cause gibbsite, formed during weathering, to convert to kaolinite. This alteration
is likely to be completed during the early stages of burial.
Most of the shale studies have been concerned with the loss or disappearance
(x-ray) of kaolinite with depth and temperature. However, a new problem arose with
the advent of TEM and lattice fringe image studies of muds-shales. I t is apparent
-
that in low temperature ( < 150°C) physilites 7 A chlorite, presumably berthierine,
as well as 14 A chlorite, is commonly present. If the 7 A physil is berthierine
(001 = 7.05 A) it should not be confused with kaolinite (001 = 7.16 A), though
variations in composition affect the 001 spacings. Where the 7 A phase occurs
randomly interstratified with the 14 A phase, the pattern is characterized by weak
OOr! reflections, with the OOr!=odd reflections weak and broad (Ahn and Peacor,
1985).
Kaolinite in shales disappears during deep burial, but there is little firm informa-
tion on what happens to it. The depth and temperature at which it disappears vary
widely and the factors which determine at what temperature it is destroyed or
transformed are not well known.
A review by Dunoyer de Segonzac (1970) indicates that kaolinite in shales
disappears at temperatures ranging from 90" to 190°C. The persistence of kaolinite
to a temperature of 180°C at Pierrefeu, he ascribed to a very low rock porosity and
the relatively young age (Oligocene) of the formation. In the Cretaceous Logbaba
section kaolinite starts to decrease at 70°C and is gone by 90°C; however, the loss
of kaolinite coincides with a formation boundary and may never have been present
in the underlying section.
In the Oligocene of the Upper Rhine Graben, kaolinite is present to total depth
in a well with a bottom hole temperature of 169°C (Doebl et al., 1974). In Oligocene
sediments of the Gulf Coast (Hower et al., 1976), kaolinite content decreases from
approximately 19% of the total physil suite to 14% over a temperature interval of
about 50" to 174°C (BHT) and apparently was not strongly involved in or created
during the I/S-chlorite diagenetic reactions. In the Gulf Coast Eocene (Boles and
Franks, 1979) kaolinite is completely destroyed between 170" and 210"C, long after
the I/S 4:l phase was formed.
In the Carboniferous " kaolinite-coal-claystones" of the Ruhr district (Eckhardt,
1963, the b-axis disorder of kaolinite decreases with an increasing temperature
(increased coal rank). At temperatures less than 130°C there was little change in the
crystallinity of the kaolinite (in claystones). No b-axis disorder was found in
kaolinite exposed to temperatures above 200°C. In Russia (Kossovskaya et al.,
1964) and Australia (Kisch, 1966), kaolinite is present in clay beds associated with
bituminous coal, but only illite and chlorite are present in semianthracite coals. On
the basis of the coal rank-temperature relations of Demaison (1974), kaolinite
associated with coal beds is destroyed between approximately 150" and 200°C.
During regional metamorphism of the Cambrian Conasauga shale, Georgia,
kaolinite was completely destroyed at a temperature of approximately 200°C.
Though there are probably exceptions, it appears that in shales kaolinite com-
-
monly persists to 200°C but starts to decompose at around 150°C or slightly less.
Kaolinite does not appear to be able to exist at temperatures in excess of 200°C.
448

Under some conditions it may convert to dickite and where it is a major component
of the rock it will alter to pyrophyllite.
Most of the decrease in kaolinite occurs at burial temperatures higher than those
where smectite converts to I/S 4:l. Some of the A1 released during the breakdown
of kaolinite may be incorporated in the I/S as it is converted to illite, but it appears
that the dissolved kaolinite more commonly participates in the formation of
chlorite, of which some may occur interstratified with I/S and illite.
Winkler (1979) reported that in rocks composed largely of kaolinite and quartz
(and containing only minor amounts of K and Mg) the kaolinite can persist to
temperatures in the range of 325" to 375°C. At higher temperatures it can be
converted to pyrophyllite. The temperature of transformation decreases as the H ,O
pressure decreases. When kaolinite occurs as a minor constituent in shales (less than
30 to 50% of the clay suite), it is usually destroyed or converted to some mineral
other than pyrophyllite at an appreciably lower temperature.
In addition to massive pyrophyllite, solutions can be mobilized and pyrophyllite
deposited in fissures produced during dynamometamorphism.
Kubler (1967) and Kisch (1974) found that pyrophyllite was present in anchizone
and early metamorphic stage physilites, which probably indicates a minimum
temperature of formation of approximately 250°C. Frey (1970) found pyrophyllite
in sediments which had been exposed to a maximum temperature of about 200°C.
Henley (1959) established the stability fields of kaolinite, pyrophyllite, mica, and
K-feldspar in terms of temperature and [K+]/[H+]. If the K + concentration and
pH are sufficiently high, kaolinite can convert to illite at temperatures encountered
at moderate depth. Such a transformation occurs in porous rock but it is not known
if it occurs in shales.
In an investigation of the alteration of physils under hydrothermal conditions,
Frank-Kamenetzky et al. (1971) found that in the presence of KCI, kaolinite first is
converted to montmorillonite and disordered kaolinite ( - 225°C) and then to a
K-hydromica at 250°C. K-hydromica is the only phase present at a temperature
higher than 400°C. In the presence of NaCl the sequence is similar, except
Na-hydromica develops and minor amounts of montmorillonite persist to a temper-
ature of 5OOOC. In the presence of MgCl,, kaolinite persists to 375°C. In the
absence of quartz and in the presence of K and Na, kaolinite is converted to
micaceous minerals. Laboratory studies by Lagache et al. (1963) established that
illite alters to kaolinite at relatively low temperatures ( < 200°C) in the presence of
CO,. Kaolinite in Gulf Coast muds was destroyed at 100°C when the mud was
heated in reactor bombs (Hiltabrand et al., 1973). The sediment-seawater ratio was
1:5, which does not duplicate natural conditions.
In nature, kaolinite may be destroyed or modified at temperatures possibly as
low as 70°C or may persist to 375"C, though 200°C is a more reasonable maximum
for shales in which kaolinite is not the predominant physil. The temperature at
which it is destroyed or transformed depends on the pH, chemistry, porosity/per-
meability, and probably time. The temperature at which changes occur appears to
decrease as pH and/or porosity increase.
Kaolinite is less abundant in Paleozoic and older shales than in younger shales.
449

Though numerous theories have been proposed to account for the small content of
kaolinite in older shales, it appears that the main reason is burial diagenesis. Most
thick sedimentary rock sections were deposited in geosynclines and basins and a
relatively large percentage of the Paleozoic and older shales have been exposed to
temperatures higher than 150" to 200°C.

BEYOND DIAGENESIS

Introduction

Most of the burial effects described in the preceding pages were obtained from
studies of Cenozoic and Mesozoic sediments in basin and geosynclinal areas that
have not been subjected to uplift, thrusting, and other major tectonic activity. Most
oil wells are drilled to depths where the maximum burial temperature is < 150°C,
occasionally 200°C, and rarely 300°C. To obtain information on what happens to
shales at temperatures beyond - 200"C, it is usually necessary to look at areas
where deeply buried sediments have been uplifted and eroded, e.g., Alps and
Appalachians.
Changes in physil type and polytype can be used in a gross way to estimate
changes in paleotemperature; however, these changes are usually difficult to mea-
sure and tend to occur abruptly over short temperature ranges. During a study of
the Ouachita Mountain region of Texas, Weaver (1960, 1961) observed that the peak
sharpness and peak width of the 10 A illite peak changed continuously with
increasing metamorphic grade. I suggested a measure of peak character (Fig. 7-14)
which I called the sharpness ratio (SR) (sometimes called the Weaver Index, W.I.).
Fig. 7-15 shows the distribution of SR values in the area where the Ouachita
structural belt is in contact (thrust) with foreland basinal shales. The SR values were
related to metamorphic grade determined petrographically by Pete Flawn (Table
7-3)(Weaver, 1960).
Later, Kubler (1967, 1968) suggested using the peak-width at half-height (in mm)
as a measure of metamorphism or crystallinity. He called this measure the Crystal-
linity Index (CI) (sometimes called the Kubler Index, K.I.). Crystallinity is a poor
choice, as the 10 A peak-width is usually a measure of the proportion of 12 to A
14 A layers interstratified with the illite. Using Weaver's boundaries, Kubler set K.I.
boundaries for three stages of metamorphism. Kubler's boundary values vary,
depending on the settings used to produce the x-ray pattern. Some of the values that
have been used are listed in Table 7-4. Other values are tabulated by Kisch (1983).
Later Kubler (personal communication) recommended reporting the peak-width in
degrees 28 (Table 7-4). This eliminates machine variation and is the best measure to
use. Both the W.I. and K.I. are strongly influenced by physil size (Fig. 7-16), and it
is recommended that measurements be made on the Na-dispersed < 2 p m size.
Actually, it is desirable to use several size fractions. The K.I. is a more precise and
450

A
Sharpness ratio =

Fig. 7-14. Method of measuring sharpness ratio. From Weaver, 1960; 1961. Copyright 1960 Texas Bur.
Econ. Geol.

faster measure than the W.I.; however, when there is a shoulder on the 10 peak, A
due to the presence of I/S, the W.I. may be more meaningful.
The measure of peak width or sharpness is of value primarily when the starting
(pre-metamorphosed) physil is smectite or I/S. Some low temperature, neoformed
illites in sandstones and evaporites have very sharp 10 A peaks. Lithology also
influences “crystallinity”, but in what way is not well established. Kubler (1968)
found the illites in limestone had higher K.I. values than adjacent shales, whereas
Weaver and Associates (1984) found the opposite relation. The presence of parago-
nite, with a basal reflection of 9.6-9.7 A, and mixed-layer paragonite/phengite will
A
broaden the 10 peak; however, for low K.I. values the 5 A peak commonly has a
K.I. value similar to the 10 A peak.
Another simple ratio measurement that has been reported to reflect increasing
metamorphic grade is the 002(5 A ) / O O l ( l O A) peak-height ratio (Esquevin, 1969;
Dunoyer de Segonzac, 1969). The ratio is presumed to be a measure of the
AI/(Mg + Fe) ratio, with the 002/001 ratio increasing as the (Mg + Fe) increases.
Weaver (1965) proposed that the ratio was directly proportional to the K content
(basically the I/S ratio). Actually, K,O and Fe,O, are inversely related in illites
(Weaver and Pollard, 1973), so if one is related to the 002/001 ratio the other will
be also. Most studies indicate a relatively poor relation between the 002/001 ratio
and degree of metamorphism (Dunoyer de Segonzac and Hickel, 1972), though such
relations do exist (Fig. 7-23).
I

SHARPNESS RATIO ~ 2 . 5TO 5.0

SHARPNESS RATIO - 5 . 0

CONTROL WELL

m i o n m

-
Fig. 7-15. Distribution of sharpness ratios of 10 A peaks along the frontal portion of the Quachita Belt
in Texas and Oklahoma. From Weaver, 1960; 1961. Copyright 1960, Texas Bur. Econ. Geol.

Rather than attempt a review of the fragmented literature on low-grade metamor-


phism, I will use our study (Weaver and Associates, 1984) of the southern Appa-
lachians as an example to illustrate some of the reactions that occur in the 150" to
400°C temperature range. This area has the advantage of being structurally simpler
than most other areas, particularly the Alps, that have been studied, and it is a study
with which I am familiar. Other than the early study of the Ouachita System, most

Table 7-3
Average Sharpness Ratios in Rocks of the Subsurface Ouachita Belt in Texas.
Average SR
Maximum - 20
Low-grade metamorphism 12.1
Weak to very weak metamorphism 6.3
Incipient to weak metamorphism 4.5
Incipient metamorphism 2.3
Unmetamorphosed 2.3
452

Table 7-4
Comparison of Sharpness Ratio, Various Crystallinity Values and Diffractorneter Speeds
Zone Sharpness Zo/rnin. X.Y. Degree 2 8 1"/min.
Ratio 1600 mm/hr 8"/min. CuK 381 mm/hr
(Weaver, (Kubler, (Kubler, (Kubler, (Weaver,
1960) 1966) 1968) 1978) 1984)
mm mrn 28 mm
Diagenesis
2.3 7.5 4.0 0.42 3.0
Anchizone
12.1 4.0 2.5 0.25 1.50
Epizone

studies of the transition from shale to slate have been in areas other than North
America. Some publications on diagenesis - metamorphism transition mineralogy
are: Kossovskaya and Shutov (1963), Kubler (1964, 1967, 1970), Dunoyer de
Segonzac (1969, 1970), Karpova (1969), Frey (1970, 1974, 1978). Frey and Niggli
(1971), Kisch (1968, 1974, 1980), Weber (1976). Frey ef al. (1980), Liewig et al.
(1981), Frey (1987), etc.
The Upper Cambrian Conasauga shale in the Valley and Ridge Province of
Georgia, Tennessee, and Alabama (Fig. 7-17) is bounded on the east by the
southeastward dipping Cartersville-Great Smoky fault and on the south by the
nearly horizontal Emerson fault; both faults are of late Paleozoic age. The Paleozoic
rocks in this part of the Valley and Ridge form a wedge-shaped deposit thickening
from around 300 m to the northwest, near the Cumberland Plateau, to approxi-

K.I. ~ W.I.
1 1 1 1 1 1 1 1 1 1 1 1

3-

2-

1- - 8

0-
<,2
1
<2
1
2 4
1
4-31
1
31- >
1 pm<.2
1 1 1 1 1 >I
<2 2-.4 4-31 31- rm
10

62.5 62.5 62.5 62.5


Fig. 7-16. Variation of K.I. and W.I. values of two Conasauga physilites as a function of flake six. The
presence of aggregates is responsible for the reversal in the coarser fractions. From Weaver and
Associates. 1984. Copyright 1984 Elsevier.
453

K.I. <2pm

NORTH

scale. kin

Fig. 7-17. Regional distribution of 10A Kubler Index (K.I.) values of the less-than-2 p m fraction of
Conasauga physilites. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

mately 6000 m to the southeast, near the Great Smoky fault. In the southeast area
the pre-erosion thickness was around 15,000 m and the Conasauga was overlain by
approximately 10,000 m of sediments. In addition, the study indicated there was
additional burial due to the extension of the now eroded thrust faults westward over
much of the southeast area.
454

Illite-Mica

Fig. 7-18 shows how the character of the 10 - A


x-ray peak changes with
increasing metamorphism. Fig. 7-17 show the distribution of the K.I. for the < 2
pm fraction. In Tennessee the contours follow the structural grain with values
decreasing from west to east as the depth of burial increases. In Georgia the
contours deviate from the structure pattern and have an arcuate shape. The
contours intersect the trace of the Great Smoky and Emerson faults at a high angle.
The pattern presumably reflects an increasing depth of burial and temperature in a
southeastern direction. The effect of the faults appears to have been minimal except
in the immediate vicinity of the faults. If the motion along the faults caused a
decrease in the K.I., then the K.I. values, uninfluenced by the faults, should be
larger and the contour pattern would be even more arcuate. Another complication is
that some samples relatively close to the fault contain paragonite (9.6 A). The K.I.
values for these samples were obtained from the 5 A
rather than the 10 peak. A
Measurements of the peak-width at half-height of the 5 A and 10 peaks of x-ray A
patterns of samples not containing paragonite indicates they have similar values.
The localized effect of the fault can be seen in a series of samples from near
Chatsworth, Georgia. The westernmost sample was collected 1.6 km west of the

b
-7.2mm -3.7mm -2.0mm -0.9mm

MID DlAGENESlS LATE DlAGENESlS


P
ANCHIZONE
J
EPIZONE
Fig. 7-18. Sequence of 10 A peaks showing decrease in peak width at half height (K.I.) with increasing
grade of metamorphism. Values are not at half-height position. From Weaver and Associates. 19x4.
Copyright 1984 Elsevier.
455

Table 7-5
Variation in K.I. and W.I. Adjacent to the Great Smoky Fault.
1.6 km West Fault
K.I. 3.7 3.0 2.4 1.2
W.I. 2.0 2.7 5.2 12

fault and the easternmost one meter under the Great Smoky fault. As Table 7-5
shows there is a systematic decrease in K.I. and increase in W.I. approaching the
fault.
As the K.I. values in Georgia can be contoured to form a near arcuate pattern it
is possible to make a plot of the various parameters based on the distance from the
point of maximum metamorphism (OW-0, asterisk south of Cartersville, Georgia).
Fig. 19 and 20 show plots of the K.I. and W.I. values of the < 2 p m fraction as a
function of distance from O W - 0 . Several values are shown for most locations. The
various values for each location may represent different samples and/or different
portions of the same sample. Further, the size fractions and x-ray slides were
prepared by several different people. Thus, the data in the plot represent a near
maximum range of error. Nevertheless, a well established trend is evident.
The K.I. values range from an average of 1.1 to 5. Not only d o the K.I. values
decrease with apparent increasing depth of burial, but there are major changes

OC
0.42 400,

0.25
*t P.

EPIZONE ANCHIZONE OlAGENESlS


456

6-
-

4- -

2-
e m
W
- (1

A
Fig. 7-20. Relation of 10 W.1. values of the less-than-2 pm fraction to distance from sample OW-0.
From Weaver and Associates, 1984. Copyright 1984 Elsevier.

separating the diagenesis zone from the anchizone and the anchizone from the
epizone. These increases in rate of change represent significant changes in the rate
of illite modification. The decrease in K.I. between 60 and 45 km is apparently due
to the loss of a discrete mixed-layer phase ( - 10 percent smectite) in the fine
fraction ( < 0.2 pm), though expanded layers ( < 5 percent) are present in the
coarser material. Between 25 and 15 km the decrease in K.I. values coincide with the
loss of the last few expanded layers. This further reinforces the significance of these
boundary values.
The contour pattern and location of the zone boundaries based on the W.1.
values are similar to those of the K.I. values. Fig. 7-20 is a plot of the W.I. versus
distance from OW-0. Values range from about 2 to 15. Again there is a change in
slope at the diagenesis-anchizone boundary. There is no change in slope marking the
anchizone-epizone boundary; however, there is a significant change in slope at a
value of approximately 6.0. This coincides with Weaver’s (1960) boundary between
incipient and weak metamorphism.
Though the K.I. and W.I. values of the < 2 pm illite-mica can be used to
determine the grade of metamorphism it is important to note that these values vary
with grain size. Fig. 7-16 shows how these values vary as a function of size. K.I.
values decrease and W.I. values increase as the grain size increases from < 0.2 pm
to 4 to 31 pm. The values reverse for the coarser size fractions. This suggests the
coarse material consists of aggregates of finer material. The K.I. contour patterns
are similar to those of the < 2 p m fraction, but metamorphic grade boundary
457

K.I.

EPIZONE ANCHIZONE DlAGENESIS


1 I I I I I I I I
10 20 30 40 50 60 70 80 90 KM

Fig. 7-21. Relation of 10 A K.I. values of various size fractions to distance from OW-0. Values converge
in epizone where complete recrystallization occurs. From Weaver and Associates, 1984. Copyright 1984
Elsevier.

locations vary by as much as 50 km ( - l0OOC). Fig. 7-21 shows how the K.I. values
of all size fractions converge with increasing metamorphism and are similar in t h e
epizone where pervasive recrystallization (to phengite) occurs.
Fig. 7-22 shows the percentage difference in K.I. values of the 10 A peak before
and after heating (300°C). The decrease in K.I. values in the anchizone samples
confirms the suggestion that some expanded layers are present even though ethylene
A
glycol has little or no effect on the shape or position of the 10 peak. Through the
anchizone the percentage of expanded layers decreases from approximately 10% to
0. No expanded iayers are present in the epizone micas (phengite) and the K.I. value
actually increases when the samples are x-rayed while still hot. This increase in K.I.
is presumably due to thermal expansion and should be an excellent method of
establishng whether or not expanded layers are present in illite-phengite.
The micaceous materials in the carbonate rocks interbedded with the physilites
have K.I. values consistently lower than for those in the physilites at a similar
distance from the reference point. A residue ( < 2 p m fraction) from the least
metamorphosed area of the middle diagenesis zone in Tennessee (OR-7) is com-
posed of I/S with approximately 10% smectite layers. The K.1. value of 5.7 is
similar to that of the adjacent shale, 5.3. Residues from carbonate rocks in the zone
of diagenesis have K.I. values considerably lower than those of corresponding
physilite samples, but the K.I. values of the physilite and carbonate illites converge
in the anchizone and become nearly identical in the epizone.
458

IZONE' I ANCHIZONE 1 DlAGENESIS I


1
30
I
40
I
50
I1
60
I
70
I
80
I
90
A
KM

Fig. 7-22. Percent change in 10 A K.I. value when samples are heated at 300°C for 2 hours. The peak
width of the epizone samples increases due to thermal expansion; the others decrease as the expanded
layers are contracted. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

In the Conasauga the 002/001 ratio increases through the diagenesis zone up to
the beginning of the epizone (Fig. 7-23), then reverses and decreases. The increase in
the ratio in the diagenesis zone and anchizone is presumably due in part to the
conversion of detrital biotite to chlorite and phengite, thus lowering the average Fe
content of the micaceous minerals. The conversion of I/S and 1Md illite to 2M
white mica also would cause the ratio to increase. The calculations of MacEwan rt
al. (1961) indicate there is a major increase in the (002)/(001) ratio of 1/S
(10 A/12.4 A)as the ratio increases from 8/2 (0.30) to 9/1 (0.62). Thus, it appears
likely that most of the increase in the (002)/(001) ratio with increasing metamor-
phism is primarily due to a decrease in the number of expanded layers. When all the
A
expanded layers are contracted to 10 the (002)/(001) ratio is more apt to reflect
changes in composition.
A sample in a window in back of the fault trace and about 3 km south of OW-0
has a (002)/(001) ratio of 0.15 and is composed predominantly of biotite. The
reversal at the beginning of the epizone may be due to the early crystallization of
some biotite and/or the crystallization of phengite with a relatively high Fe content.
Several studies have demonstrated that the 2M/1 Md ratio increases with increas-
ing degree of metamorphism (temperature): Reynolds (1963), Maxwell and Hower
(1967), Kisch (1968), Karpova (1969), Dunoyer de Segonzac (1969), and Foscolos
and Kodama (1974), to list a few.
O'x/O'6 studies (Eslinger and Savin, 1973) of Precambrian Belt sediments ( < 0.5
p m fraction) suggest that burial temperatures ranged from 225°C for a shale having
459

EPIZON E ANCHIZONE 01AGENESIS


I I I I I I I I

- -
%
100 ;= -- 0 FAULT -

80 - -

60 - -

40 - -
0

20 - -
EPIZONE ANCHIZONE OlAGE%ESIS
0
0
I
10 20 30 40 50
I
60
- 1
75
I
80
-90 KM
I.

Fig. 7-24. Relation of relative amount of 2M mica to distance from O W - 0 . The relatively high value for
OW-3 is because of its close proximity to the Great Smoky fault. 0.2-2 p m fraction. From Weaver and
Associates. 1984. Copyright 1984 Elsevier.
460

a 2M/lMd ratio of 1/3 to 310°C for a shale with a ratio of 3/2. Extrapolation of
their data indicate that a temperature of 400°C would be required to obtain 100%
2M illite. These results are similar to those obtained for the Conasauga.
In the Conasauga the relative amount of 2M illite-mica (2.80 A/2.58 A) increases
with increasing grain size except for epizone samples which contain only the 2M
polymorph. Fig. 7-24 shows the distribution of 2M illite-mica in the 0.2 to 2 p m
size. The relative amount of 2M mica systematically increases through the anchizone
and in the epizone reaches 100%. One sample (OW-3) in the anchizone has 100%
2M mica. This sample is close to the Great Smoky fault and suggests the conversion
to 2M mica may have been influenced by local temperatures and/or pressure
generated by movement along the fault. The data show that essentially all of the
conversion of 1Md to 2M occurs in the anchizone or through the temperature range
of approximately 280” to 360°C. The 2 to 44 pm fraction of diagenesis zone
samples contains appreciable 2M mica. The 2M polymorph does not develop until
fewer than 10%expanded layers are present in the fine fraction.
The (060) values for the illite-phengite in < 2 pm fraction range from 1.499 A to
1.510 A. Micas from the anchizone have the lowest values; micas in both higher and
lower grades of metamorphism have higher values.
The “illite” from the late diagenesis ( < 0.2 p m fraction, I/S with 10% smectite
layers) has a typical ilite structural formula:
.59 Fe0.24Mg0.17 (si3.45 0.55 )OIO (OH)ZK0.62 Na0.03Ca 0.06

The micaceous material from the epizone, which is completely recrystallized, has a
phengite composition:
(A’l.39Fe0.12M~0.51
)(Si3.53A10.47 )010(0H)2K0.93Na0.01Ca0.01

Phengite starts to form by replacing detrital biotite during late stage diagenesis
but does not become the predominant micaceous material until the epizone stage of
metamorphism.
“Illite” is the primary matrix physil and primarily responded to the depth of
burial or temperature. Most of the other mineralogic changes are influenced by
tectonic stresses, which will be discussed next.

Petrology

The interpretation of the mineral, chemical, and isotope data is to a large extent
dependent on the petrographic interpretation though a large amount of synergism is
involved. The least metamorphosed samples are typical shales with moderate
orientation of the physils. With increasing metamorphism and tectonic deformation
mineral rotation and recrystallization occurs and a slaty cleavage developed. In the
greenschist facies a second cleavage is superimposed on the slaty cleavage.
The origin of slaty cleavage has been investigated for many years and there is still
wide disagreement. For reviews of the subject see Etheridge et ul. (1974) and
Williams (1977). There are two major opinions. One is that foliation forms normal
46 1

Fig. 7-25. SEM of Conasauga shale sample from zone of moderate diagenesis (TN). Orientation is fairly
well developed. Large flakes are muscovite and biotite, smaller flakes are largely illite and I/S. Bar = 10
pm. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

to the shortest axis of the strain ellipsoid by mechanical rotation of the physil plates.
The other is that foliation forms by crystallization and growth of mineral grains
with a preferred orientation which defines the foliation. The prevailing opinion
appears to be that rotation is the major mechanism for the formation of foliation
during very low-grade metamorphism and crystal growth becomes increasingly more
important with increasing grade of metamorphism. The present study tends to
confirm the latter opinion.
It appears that the relative importance of the two mechanisms is dependent on
the timing of the tectonic deformation. If lateral stress is applied early (shallow
burial), temperatures are low and crystallization is at a minimum and rotation at a
maximum. If lateral stress is applied to deeply buried, high temperature physilites,
the role of crystallization should be more important.

Diagenesis
In the zone of diagenesis orientation of the large mica flakes (biotite and
muscovite) and the finer-grained illite-I/S is moderately well developed (Fig. 7-25).
The first obvious (optically) mineralogic changes occur in samples in which an
incipient slaty cleavage developed during the late stage of diagenesis. Detrital biotite
462

Fig. 7-26. Incipient slaty cleavage in high grade diagenesis zone, OW-7 near Chatsworth, Georgia. The
bedding is approximately horizontal. Large mica flakes have been rotated 30° to 90". Pod in center is
chlorite. Compare with Fig. 7-25. Bar=lO pm. From Weaver and Associates, 1984. Copyright 1984
Elsevier.

partially alters to a white mica and green phengite. The latter appears to have
formed first. The altered biotite has been slightly rotated and lies at an angle to the
cleavage. As the Great Smoky fault is approached the secondary cleavage becomes
more pronounced. In the vicinity of Chatsworth the secondary cleavage is fairly well
developed. There is a less well developed cleavage at around 30" to the major
cleavage. The two cleavages produce a pencil fracture pattern. At this stage chlorite
porphyroblasts developed (7 to 10% of sample); chlorite also replaces trilobite
fragments. The K.1 value (3.7) and K-Ar data indicate there was little recrystalliza-
tion or neoformation of the mica phase at this stage; the new orientation is
primarily due to rotation and folding, primarily of the illite and I/S.
Fig. 7-26 is a low magnification view of a sample from the zone of incipient
metamorphism. Fig. 7-27 shows chlorite porphyroblasts. Fig. 7-28 and 7-29 are
higher magnification pictures showing the presence of small kinks and folds ( 5 to 20
pm in length) in the matrix illite-I/S. The axial planes are aligned in the general
direction of the main cleavage. In Fig. 7-29 a sphere of chlorite (center) is overlain
by a tight fold of illite and bounded on the right by a splayed package of biotite.
463

Fig. 7-27. Chlorite pods in sample (OW-7) from high-grade diagenesis zone. Crenulations are not
observed at this magnification but are present on a small scale. Bar = 0.025 mm. From Weaver and
Associates, 1984. Copyright 1984 Elsevier.

EDX analyses indicate the chlorite porphyroblasts are commonly surrounded by


biotite.
The thin section and SEM pictures indicate the chlorite pods altered from mica,
largely biotite, apparently in areas where there was the potential for void space to
develop, i.e., small folds (slippage), splayed books, and areas supported by a
framework of rigid grains. Once chlorite growth started in small voids, continued
stress caused the voids to re-open and growth of the chlorite continued until the
chlorite pod was aligned perpendicular to the direction of stress.
Williams (1972) believes that slaty cleavage starts with the development of
crenulations. The local increase in strain energy causes recrystallization and the
growth of new crystals parallel to the axial planes of the host crenulations and
preferentially grow to form the cleavage lamellae. White and Knipe (1978) con-
cluded from an electron microscope study that mechanical rotation of physils into
the plate of cleavage lamellae is rare. The SEM picture of OW-7 indicates there has
been some crystallization but there has been more rotation and bending than
crystallization. Much of the crystallization involved the development of porphyrob-
lasts rather than the development of new flakes in the cleavage lamellae.
In addition to Si, Al, Fe, and Mg had to have been mobilized to produce chlorite
ovoids and for chlorite to have replaced calcite. Data from other studies (Hower et
464

Fig. 7-28. Small, sharp folds of illite (OW-7). Axial planes parallel to incipient slaty cleavage. Folding
influenced by quartz grain in center. Bar = I p m . From Weaver and Associates, 1984. Copyright 1984
Elsevier.

al., 1976; Boles and Franks, 1979) indicate that as smectite alters to illite Si, Fe, and
Mg are released and probably form chlorite. The development of chlorite
porphyroblasts is accompanied by only a minor increase in the proportion of illite
layers. The K-Ar data also indicate there has been only a relatively minor mobiliza-
tion of K. There does not appear to have been enough smectite converted to illite to
account for all the chlorite that was formed. Most of the chlorite apparently formed
from biotite and perhaps some from muscovite.

Anchizone
The anchizone (K.I. 3 to 1.5, W.I. 2.3 to -
10) occurs in an arcuate belt
approximately 40 km wide. The anchizone-epizone boundary occurs approximately
20 km from the area of maximum metamorphism (OW-0). Most of the physilites
have a slaty cleavage that becomes more pronounced as the Great Smoky fault is
approached. Well developed slaty cleavage extends 10 to 20 km west of the fault
trace.
46 5

Fig. 7-29. Small sphere of chlorite surrounded by splayed flakes. probably biotite (OW-7). Bar =5 pm.
From Weaver and Associates, 1984. Copyright 1984 Elsevier.

The physilites consist primarily of illite-phengite-white mica, a lesser amount of


chlorite, and a minor amount of I/S withrapproximately 5% smectite layers. In the
western part of the area samples contain up to 5% biotite and minor muscovite. The
biotite disappears to the east as the Great Smoky fault is approached. Porphyrob-
lasts increase from west to east and decrease from north to south as the metamor-
phic grade increases. Paragonite has a similar distribution.
In samples from near the anchizone-epizone boundary and at a considerable
distance from the Great Smoky fault porphyroblasts are not present. Ten to 50% of
the biotite laths are oriented (or partially bent) at an angle to main cleavage (Fig.
7-30) and partially replaced by phengite and/or white mica. The phengite and white
mica commonly occur along one or both sides of the biotite and may have grown in
pressure shadows epitaxly on the biotite. Replacement is more pronounced in bent
biotites and those at an angle to the main cleavage (Fig. 7-31). Large arcuate
patches of mixed green phengite and chlorite are present. They appear to have
replaced trilobite fragments. In the sandstone laminae some of the high green biotite
has altered to chlorite.
Eastward, towards the Great Smoky fault, porphyroblasts develop and increase
to a maximum of 23% (OW-26) a few km from the fault (Fig.7-32). The flakes in the
466

Fig. 7-30. Bimodal distribution of rotated biotite and muscovite packets in anchizone samples. Bar = 10
pm. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

pods are aligned at a high angle to the slaty cleavage. In the lower part of the
anchizone 50 to 70% of the pods are pure chlorite. The others are phengite (green)
and two phase mixtures of phengite, white mica, and chlorite in all combinations.
All three physils are present in some pods. White mica increases relative to phengite
(green) as the fault is approached. The white mica may be phengite with less Fe or a
different ferric/ferrous ratio than the green phengite. As the metamorphic grade
increases, southward, the porphyroblasts are destroyed or decreased in size by
pressure solution and largely converted to white mica.
In the least metamorphosed areas some of the quartz is elongate. With increasing
metamorphism most of the quartz occurs in lenses. It consists of both subequant,
silt sized quartz with slightly sutured boundaries and elongate, recrystallized grains.
Both types are commonly partially replaced by chlorite. The white mica appears to
replace both the quartz and the chlorite. Some lenses have beards of white mica.
Rotation, bending, and kinking are observed in the least metamorphosed physi-
lites. As the Great Smoky fault is approached the slaty cleavage lamellae, largely
white mica, are marked by dark anastomosing planes that outline lenticular domains
that contain the physil pods described above (Fig.7-32), quartz, and some calcite.
467

Fig. 7-31. Same sample as in 7-30 showing a bent mica book. Vertical portion is biotite. Bent portion has
altered to white mica. Bar = 10 pm. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

The distance between lamellae are mostly in the range of 0.02 to 0.1 mm. South-
ward, with increasing metamorphic grade, the dark cleavage lamellae become
subparallel and much closer together (0.01 to 0.02 mm). The material in the
lenticular domains has been dissolved by pressure solution and much of the material
recrystallized in the cleavage lamellae.
Some of the cleavage in the pods has a fan-shaped distribution, but most are
oriented in one direction. Some are kinked with the axial planes oriented parallel to
the main cleavage lamellae. Others show slippage and minor bending. The pods
show various degrees of modification, ranging from round with vertical cleavage,
through various degrees of flattening and decrease in the cleavage angle, to the stage
where they are thin elongate lenses with the axis and cleavage traces parallel to the
slaty cleavage. Compression-strain apparently developed extension sites between the
parallel chlorite plates and white mica grew by mimetic growth on the cleavage
faces. In some instances chlorite pods with serrated boundaries contain bands of
468

Fig. 7-32. Lenticular domains of chlorite with flakes oriented at a high angle to the slaty cleavage. The
anastomosing lamellae are composed largely of white mica (illite-phengite). Bar = 10 pm. From Weaver
and Associates, 1984. Copyright 1984 Elsevier.

phengite adjacent to the boundaries, indicating chlorite has altered to phengite. In


other pods phengite has altered to chlorite. In some chlorite-phengite pods the end
portion of plates of chlorite that are oriented nearly perpendicular to the slaty
cleavage are bent in the direction of the cleavage and the bent portion altered to
white mica. This is an excellent example of how the physils in lenticular domains
recrystallize and are incorporated into the cleavage lamellae. Other lenticular
domains are composed of aggregates of irregular quartz grains partially replaced by
chlorite and phengite.
During this early stage in the development of slaty cleavage (low anchizone)
white mica, phengite, and chlorite continued to crystallize. Most of the white mica
grew in extension sites and the chlorite and phengite replaced quartz, matrix physils,
and perhaps each other. The source materials were presumably the remaining
detrital illite and biotite. The decrease in the apparent K-Ar ages of the coarse
fraction confirms that appreciable detrital mica has been destroyed and/or new
mica created.
469

Closer to the Great Smoky fault, planar slaty cleavage is well developed. Some
chlorite pods have rotated and become parallel to the cleavage, but in most
instances they appear to have dissolved at the junction with the cleavage lamellae.
The dissolved material migrated along the pod boundary and new chlorite crystal-
lized on the ends of the pods between the cleavage lamellae. The chlorite forms long
thin lenses (aspect ratio- 3) and the cleavage lamellae coalesce. In many instances
white mica, rather than chlorite, crystallized at the end of the pods and occasionally
replaced part of the pod. Many of the chlorite lenses have bent and mildly kinked
cleavage fractures subparallel to the main cleavage direction.
Fig. 7-33 shows that at least two types of micaceous material are present in the
anchizone physilites. One type consists of thin, smooth, tightly packed flakes similar
to those in chlorite pods; the other type consists of discrete, thicker, smaller, less
tightly packed flakes. The thick flakes have a higher Ti and Si and lower K content
than the thin flakes. The thin flakes would appear to be newly formed metamorphic
mica and the small thicker ones detrital or diagenetic illite.
With a further increase in burial temperature (high anchizone) the physil, quartz,
and calcite pods are progressively dissolved and the anastomosing dark lamellae
become subparallel. Physil pods become thin, scarce, and consist largely of white
mica or chlorite partially replaced by white mica. Pods of white mica, calcite, and
pyrite commonly have beards of quartz and/or chlorite and occur as lens-shaped
domains. Larger patches of calcite are partially replaced by chlorite and quartz.
Cross-cutting veins and fractured pyrite clusters are present in high anchizone
slates. The veins contain quartz, calcite, and Mg-rich chlorite. Chlorite replaces
quartz but not calcite, which presumably formed later. Pyrite aggregates and
fractured clusters of pyrite are relatively abundant, chlorite laths (0.2-2 p m wide),
and in some instances quartz, extend from the edge of the pyrite and are approxi-
mately parallel to the slaty cleavage. In some instances the laths are slightly curved
indicating some rotation occurred during growth (Fig. 7-34). The Fe/Mg ratio of
this new generation of chlorite is less than that of the chlorite porphyroblasts.

Epizone
The anchizone-epizone boundary intersects the Great Smoky fault at a high angle
and is roughly parallel to the younger Emerson fault 20 to 25 km to the southeast.
K.I. values range from 1.5 to 1.2 and W.I. values from 12 to 15. The physils consist
largely of white mica and chlorite. Chlorite decreases to the south as the metamor-
phic grade increases. The chlorite and micaceous material apparently reacted to
form phengite (chemical analyses). In the lower part of the epizone slaty cleavage is
well developed. The dark cleavage lamellae come increasingly closer together with
increasing metamorphic grade (Fig. 7-35) and physil porphyroblasts are progres-
sively lost and converted to white mica, largely phengite. They comprise less than 1
percent of the rock. A second phase of chlorite (Mg-rich) is present. It partially
replaces calcite and quartz lenses and layers. It occurs as ptygmatic veins roughly
perpendicular to the slaty cleavage and in fractures parallel and at a n angle to the
slaty cleavage. T h s zone may be equivalent to the chlorite zone of the metamorphic
geologists.
470

Fig. 7-33. Upper picture shows small, relatively thick flakes of detrital-diagenetic illite. Lower picture
shows their tightly packed flakes of neoformed illite-phengite.Bar = 1 pm. From Weaver and Associates,
1984. Copyright 1984 Elsevier.
471

Fig. 7-34. Curved chlorite laths extending from the side of a pyrite cluster (right). Curved orientation
indicates chlorite growth over a period of time as slaty cleavage developed. Bar = 10 pm. Weaver and
Associates, 1984. Copyright 1984 Elsevier.

Individual quartz grains are scarce and most of the quartz occurs as elongate
lenses and veins. Most of the quartz aggregates resemble chert. Pressure solution
apparently was the mechanism by which the quartz was mobilized.
To the south as the Emerson fault is approached kinks are observed and 4 km
north of the fault trace (OW-1) a second planar foliation ( S , ) develops approxi-
mately perpendicular to the slaty cleavage. This foliation is presumably due to
movement along the Emerson fault. In the same area the temperature was increased,
due to deeper burial, and much of the mica recrystallized, forming subequant
packets. Recrystallization is confirmed by the K-Ar data. Closer to the fault (OW-0)
S, crenulation cleavage is well developed with the fold limbs oriented nearly parallel
to each other and the mica packets are more elongate.
A sample from a window a few hundred meters south of the fault trace the
phengite has partially recrystallized to light tan biotite. The biotite has been forced
472

Fig. 7-35. Dark cleavage lamellae in slate from low epizone. Dark color is due to small vacuoles, not Ti,
Fe, or organic material. See Figure 7-36. Bar = 0.06 mm. From Weaver and Associates, 1984. Copyright
1984 Elsevier.

into parallel sharp folds and the plates have a near parallel orientation. This sample
appears to represent an early stage in the development of a biotite schist.
Fig. 7-35 and 7-36 show the well-developed slaty cleavage in the lower epizone
samples. The dark cleavage lamellae are largely parallel and more closely spaced
than in the anchizone samples. The dark cleavage lamellae are composed of parallel
phengite flakes. They are separated by light lamellae composed of small lenticular
particles of quartz, white mica, and possibly very light green chlorite. The flakes i n
the lenticular domains are commonly oriented perpendicular to the slaty cleavage.
These are apparently the remnants of the coarser lenticular domains seen in the
lower-grade metamorphic rocks. The dark lamellae are relatively pure phengite but
473

Fig. 7-36. Alternating dark bands of white mica and light bands containing concentrations of lenticular
quartz, white mica and chlorite. Voids are presumably caused by the removal of lenticular material.
Bar =10 pm. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

do contain some thin, elongate lenses of quartz. There has been some migration of
silica but the differentiation is not as complete as it appears at first glance. The
distance between the dark cleavage lamellae decreases from an average of 50 pm
(lower anchizone) to 3 pm (high epizone). This reduction cannot all be due to
dissolution. Apparently differentiation occurs within the light lamellae and new
dark and light lamellae are created.
Some of the physillites in the lower epizone are literally permeated with very light
green chlorite. It fills fractures at an angle to the slaty cleavage; it fills fractures
parallel to the slaty cleavage; it replaces calcite lenses and layers; it replaces quartz;
it occurs as lenses and irregular patches containing white mica and quartz inclusions
suggesting it has replaced the matrix mica; it occurs as ptygmatic veins roughly
perpendicular to the slaty cleavage and parallel to the original bedding. The
ptygmatic veins indicate shortening was on the order of 50 percent. Some of the
folded veins contain patches of calcite suggesting they were originally calcite veins
or beds. Chemical analysis of the chlorite indicates the Fe/Mg ratio is lower than in
the chlorite in the zone of diagenesis.
474

Fig. 7-37. Planar foliation (S,) superimposed on slaty cleavage. Bar = 50 pm. From Weaver and
Associates. 1984. Copyright 1984 Elsevier.

Much of the chlorite appears to have crystallized after the slaty cleavage
developed. The kinking indicates there was a second stage of deformation following
the development of the slaty cleavage. Chlorite growth preceded or was contempora-
neous with this second stage of deformation, apparently related to the development
of tension fractures. The late stage of deformation is most likely related to the
northward thrusting motion of the Emerson fault.
In slightly higher metamorphic grade physils chlorite is scarce and phengite is the
dominant physil (OW-1). Kinking increases and a second planar foliation (S,) is
observed (Fig. 7-37). SEM pictures of the planar foliation show some of the mica
plates are rotated as much as 45" and have slipped parallel to the basal planes.
Some bundles of mica plates wedge-out as they are rotated into planar foliation;
others thin by as much as 70 percent. This suggests extensive solution took place in
the second planar foliation shear zone. EDXA spectra indicate there is no signifi-
cant difference between the white mica in the faults and that in the matrix.
Fig. 7-38 shows pictures of another crenulation cleavage lamellae. In this example
the mica plates appear to be thinner and smaller than those in the matrix. The
chemistry is similar. Visually it appears that the difference in morphology is due to
475

Fig. 7-38. Higher magnification view of S, planar foliation. Material in foliation is finer-grained than
unrotated material and mylonized and/or partially dissolved. From Weaver and Associates, 1984.
Copyright 1984 Elsevier.

slippage, breakage, and perhaps solution. Crystallization or recrystallization may


have occurred, but there was little change in composition. EDX maps show that Fe
and Ti are concentrated in the faults formed by the crenulation cleavage. Both Ti
and Fe apparently were mobilized during the development of the second set of
crenulations; however, some of the concentration may be due to solution and
volume reduction.
It is evident that much of the second planar foliation was produced by simple
rotation (Fig. 7-37); however, other features (Fig. 7-38) suggest that the rotated
flakes have been mylonized, perhaps recrystallized, and possibly dissolved. Con-
ceivably all three processes were operative.
Other samples from the same general area contain no distinct cleavage lamellae
or kinks. The mica units are no longer plate-shaped but occur as subequant packets
or aggregates with irregular interpenetrating boundaries (Fig. 7-39). They average
0.01 to 0.03 mm in width. The mica has the appearance of the secondary nests of
mica that are found in pores in sandstones. As the K-Ar data confirms (Fig. 7-51)
the mica has been completely recrystallized. This pervasive recrystallization occurs
near the beginning of the greenschist stage of metamorphism (epizone).
476

Fig. 7-39. Recrystallized subsequent packets of phengite from high epizone. From Weaver and Associ-
ates, 1984. Copyright 1984 Elsevier.

One sample contains an abundance of white, irregular veins and clots of coarse
(up to 0.07 mm), clear secondary mica. Some patterns resemble echelon tension
fissures. The mica flakes in the veins are large and form an open honeycomb pattern
(Fig. 7-40). Flakes are bound edge to edge and edge to face. The open network of
flakes appears to be real, though it is surprising that such a fabric should form at

~~ ~

Fig. 7-40. Veins of secondary white mica in epizone physilite. Upper picture shows large flakes in veins
aligned perpendicular to oriented matrix mica. Bar = 1 pm. The lower picture shows a honeycomb
arrangement of white mica flakes in secondary mica vein. Bar =10 pm. From Weaver and Associates,
1984. Copyright 1984 Elsevier.
477
47 8

depth. The veins appear to have formed after the second stage crenulation cleavage
developed. The veins are apparently a product of thermal metamorphism and the
contortion suggests the rocks behaved “plastically” due to lowered viscosity at high
P,, (Spry, 1969, p. 205).
The mica in samples from a window a few hundred meters south of the trace of
the Emerson fault is organized into subparallel bands in which the orientation is
slightly different. The bands are due to sharp folds with opposing flanks forming
angles ranging from 45” to nearly 0”. Much of the mica has been rotated to the
extent that it is nearly parallel and a differentiated crenulation cleavage developed.
Most of the mica occurs as well developed plates and books and has the appearance
of well crystallized mica. A few areas still contain subequant packets of white mica
from which the platy mica crystallized. The mica oriented parallel to the new
cleavage is slightly cleaner than that at a significant angle to the cleavage. Much of
the mica has a very light brown color and is pleochroic. X-ray analysis indicates the
mica is primarily biotite. The second stage orientation of mica is nearly complete in
this sample. A slightly more advanced stage of metamorphism would produce well
oriented coarse mica and a schistose texture.

Layer Silicate Films


Slates characteristically contain thin dark lamellae (Fig. 7-35) which, among
other things, have been called layer silicate films. It is along these films that the slate
cleaves. These films or lamellae are rich in parallel oriented layer silicates and
oxides of Fe and Ti (Hobbs et al., 1976; Roy, 1978; and others). Electron
microscope studies (White and Knipe, 1978) indicate the central part of the lamellae
are relatively coarse grained, well oriented flakes of chlorite and mica. The margins
contain finer grained, less well oriented physils.
A number of origins have been proposed for layer silicate films. It has been
suggested they form by passive concentration of physils and dark insoluble oxides
(Fe, Ti) as quartz and other minerals are removed by solution (Williams, 1972),
somewhat equivalent to stylolites in limestones. Roy (1978) suggested the films
formed in fractures that were normal to the direction of shortening. The fractures
formed in mica layers formed by rotation during the soft sediment stage of
deformation. White and Knipe (1978) found that cleavage lamellae (layer silicate
films) developed in the most intensely deformed zones where grains are most
sharply bent, kinked, and folded and where internal strain energy is highest. They
believe the development of the cleavage lamellae is primarily a metamorphic
crystallization process in which metamorphically stable physils preferentially de-
velop in zones of intense deformation. The cleavage lamellae migrate laterally by the
breakdown and regeneration of the lenticular domain physils at their junction with
the cleavage lamellae. They suggest that fluid transport is localized along these
junctions. They also note that the dark color of the cleavage lamellae is lost when
thin sections are less than 10 p m thick. This color is not due to Fe and Ti oxides.
In slates the clay films appear as an anastomosing network outlining lenticular
domains (Hobbs et al., 1976). With increasing metamorphism the lenticular domains
become less prominent and the films broader. Eventually the domains are lost and
479

there is a strong preferred orientation. The Conasauga has a similar sequence of


fabrics. Over a distance of approximately 60 km the distance between dark cleavage
lamellae systematically decreases from an average of 0.1 to 0.005 mm. The lenticular
domains and pods (phengite, chlorite, quartz, calcite, and pyrite) decrease in width,
in some instances increase in length, and progressively assume the orientation of the
cleavage lamellae.
The dark cleavage lamellae have a revealing feature: they are white in reflected
light. This indicates the dark color is not due to Fe oxides or organic carbon. It
could be due to Ti oxides, but the EDXA data indicate Ti is not concentrated in the
dark lamellae. Small fluid- or gas-filled vacuoles or bubbles produce a black color in
transmitted light and a white color in reflected light. The study suggests water-filled
vacuoles are concentrated in the physil films (layer silicate films) and are responsi-
ble for many of the features observed.
In the Conasauga samples from the zone of middle diagenesis (shales), the dark
material, which is white in reflected light, is relatively uniformly distributed. Round
to subround, what appear to be bubbles or vacuoles are concentrated in the dark
material. Where the slaty cleavage is well developed the dark lamellae have an
anastomosing pattern and essentially all the vacuoles are elongate. The elongate
vacuoles commonly are 5 to 10 p m long and have an aspect ratio of 5 to 10. As the
lamellae become more nearly parallel and closer together, the elongate vacuoles tend
to be larger, ranging from 5 to 30 pm long. As the second stage of kinking develops,
elongate vacuoles are rotated into the newly developed cleavage. When extensive
recrystallization occurs, well into the epizone, the dark material is no longer
observed. Vacuoles are scarce and are concentrated in the material that is not
recrystallized (grayish color). The recrystallized mica is clear. Most of the fluid has
apparently been concentrated in veins where white mica crystallized.
It appears that during flake rotation and recrystallization much of the fluid is
concentrated in elongate vacuoles parallel to the newly developed cleavage. The
parallel fractures produced during the development of crenulation cleavage prob-
ably serve as channelways for water movement. As the fractures heal, some of the
fluid is trapped in isolated bubbles and remain as a liquid residue. Chlorine is
relatively abundant in the physils enclosing the bubbles. The concentration of these
fluid-filled vacuoles in parallel sheets may account for the ease with which the slates
can be split.
At a more advanced stage of metamorphism, where pervasive recrystallization is
beginning, the dark and light bands are destroyed and kink bands or planar
foliation develops (Fig. 7-37). Apparently, as the kinks developed and rotation,
solution, and recrystallization occurred, the fluid-filled vacuoles, along with the Fe
and Ti, concentrated in the kinks. The high concentration of Ti suggests there has
been considerable solution of the physils in the vicinity of the kink axes. Thus, it
appears that dark cleavage lamellae developed at high temperatures can have high
concentrations of Ti and Fe. Solution of physils is probably more extensive at high
temperatures than at low temperatures. Thus, whether the dark color of cleavage
lamellae is due to vacuoles or Ti and Fe may depend on the temperature at the time
of formation of the lamellae.
480
48 1

Though it is difficult to prove that vacuoles containing fluids are present and
concentrated in the dark lamellae, fluids should be present in the metamorphosed
physilites and this is the most logical location. In addition to any original pore water
remaining in the rock, water is continuously released as the more hydrous minerals
are converted to less hydrous minerals. In the Conasauga, kaolinite, I/S, and
chlorite are destroyed or altered to a less hydrous mica. Though some of the
released water escapes through fractures, it seems plausible that some would be
trapped between the newly crystallized mica flakes, particularly in voids produced
during the formation of crenulations.

Lattice Fringe Images


In an outcrop at Lehigh Gap, Pennsylvania, the Martinsburg Formation grades
from mudstone to slate over a distance of 100 m (Lee el al., 1984, 1986). W.I. values
indicate that the metamorphic grade ranges from high anchizone to high epizone.
Perhaps the most interesting aspect of the various TEM studies by Lee et al. (1984,
1985, 1986) is the common interstratification of various types of layers.
Fig. 7-41 of the mudstone sample shows subparallel packets of illite and chlorite.
The higher magnification picture shows small scale interlayering of illite and
chlorite layers and the presence of ordered 1 : l 24 A chlorite/illite units. Kaolinite
also occurs interlayered with illite. Interlayered kaolinite, in thinner packets, is also
present in the slate. This is an advanced stage of metamorphism for kaolinite to be
present and rather than its presence indicates a low temperature, it is more likely
that the 7 A matrial is dickite. Crystal imperfections (edge dislocations, low-angle
grain boundaries, and stacking disorder) are characteristic of the mudstone samples.
The stage of diagenesis of the mudstone is aproximately equal to that of the deeper
Gulf Coast samples.
In transitional samples (Fig. 7-42) the physils contain well defined packets of
illite and chlorite layers; however, random interlayering of 10 and 14 A layers is
common. A 7 A phase, berthierine, is also present and grades laterally into 14 A
chlorite and 10 A illite (Fig. 7-43). The regularly and randomly interlayered illite
and chlorite occurs only in the cleavage plane and not in the bedding plane.
Presumably it is newly generated material and indicates a delicate response to
pressure solution and layer growth.
At the slate stage there are no randomly interlayered chlorite/ illite units. Illite
and chlorite occur as well-developed packets several thousand Angstroms thick
(Fig. 7-44). Packet boundaries are commonly parallel but some stacking disorder (of
7 A and 14 A layers) occurs in thin zones at the interface between chlorite and illite.

Fig. 7-41. (A) A portion of a typical phyllosilicate grain containing subparallel packets of illite and
chlorite layers in mudstone samples fro? Lehigh Gap, Pennsylvania. (B) A lattice fringe image showing
an interlayered relationship between 10 A illite and 14 A chlorite from the same area as A. From Lee et
al., 1984. Copyright 1984 Springer-Verlag.
482

Fig. 7-43. A hi* a


resolution lattice fringe image showing interlayer transitions between 10 A illite, 14
chlorite and 7 A berthierine. The sample is intermediate between mudstone and slate. From Lee ei a/..
1984. Copyright 1984 Springer-Verlag.
483

Fig. 7-44. A lattice fringe image of slate showing highly perfect chlorite and illite interleaved as packets
of layers. From Lee et al., 1984. Copyright 1984 Springer-Verlag.

Chlorite

Chlorite crystallizes in pelitic rocks ranging in grade from very-low-grade to


low-grade metamorphism (Winkler, 1979; Turner, 1981). The Conasauga study and
other studies indicate significant amounts of chlorite can form during the late stages
of diagenesis.
In a study of the Hunsruckschiefer slates of West Germany, Roy (1978) found
the initiation of chlorite neogenesis was the earliest evidence of metamorphic
change. He believed there was a phase of deformation prior to the growth of
chlorite-muscovite porphyroblasts. The deformation caused the detrital mica to be
rotated, facilitating the growth of chlorite in the kinked and fractured mica. Chlorite
first formed in slates only slightly metamorphosed and not far away from the
domain of diagenesis. Weber (1976) found that chlorite porphyroblasts started
growing by overgrowth and intergrowth on detrital micas at a stage near the
boundary between deep epigenesis and early metagenesis (late diagenesis and early
anchimetamorphism).
In the Conasauga shale, chlorite, along with kaolinite, is present in the < 2 pm
fraction of samples from the middle diagenesis zone in amounts of 0 to 15 percent.
Petrographic examination indicates some is detrital but secondary chlorite is present
in sandstone laminae. Kaolinite ranges from 0 to 25 percent.
484

TN. 10
\ GA.

%CHLORITE 2 4 4 p m

w- &@N
FP.0L7 O km l5

Fig. 7-45. The percent of chlorite in the 2 to 44 p m fraction. Most of the chlorite west and north of slate
line (beginning of flake rotation) is detrital. From Weaver and Associates. 1984. Copyright 1984 Elsevier.

Secondary chlorite, in any abundance, is first encountered in the zone of late


diagenesis. The chlorite developed as relatively large pods and is concentrated in the
coarser fractions. Fig. 7-45 shows the distribution of chlorite in the 2 to 44 pm
fraction. The amount of chlorite decreases with decreasing particle size. The areal
distribution of chlorite in the finer fractions does not have a well-developed pattern.
Coarse chlorite develops very abruptly and then systematically decreases in
abundance to the south as the metamorphic grade increases (Fig. 7-45). The
petrographic investigations show the chlorite is progressively replaced by phengite
and white mica. The percentage contours are aligned at nearly right angles to the
Great Smoky fault and roughly parallel to the younger Emerson fault. The slate-line
(beginning of crenulations) appears to serve as a mineralogic fault. As was suggested
in the petrography section, lateral compression and the development of crenulations
apparently aids in fluid movements, pressure solution (differential stress), and the
development of voids. These factors favor the development of chlorite porphyrob-
lasts.
The configuration in Fig. 7-45 and the petrographic data (development of a
second generating of crenulations) suggest that following the first tectonic event,
burial continued, temperatures increased, and the chlorite porphyroblasts (Fe-rich)
485

K.I.
5

4-
I
3- -

2- -

1- 8
EPIZONE ANCHIZONE DlAGENESIS

were progressively replaced by phengite andd white mica. The parallelism of the
chlorite contours to the Emerson fault and the development of a relatively Mg-rich
chlorite in veins and fractures suggests the alteration of chlorite to white mica may
have been facilitated by the northern thrust of the Emerson fault, though an
increase in temperature due to overthrusting and deeper burial was presumably the
major factor. The second stage chlorite was apparently produced largely from illite
and the first stage primarily from biotite.
The chlorite in the late diagenesis zone, the anchizone, and well into the epizone
contains a small percentage of vermiculitic layers. Chemical and heat treatments
indicate that the chlorite in the porphyroblasts is quite complex. At least two types
are present. One phase is Fe-rich and the other, possibly corrensite-like, is Mg-rich.
The expanded layers may contain some hydroxy interlayer material.
Fig. 7-46 shows the peak width at half-height of the chlorite 7 A peaks and the
K.I. of the 10 A peak as a function of distance. The chlorite peak decreases in width
with increasing metamorphic grade, but the magnitude of the change is considerably
less than for the mica-I/S. Chlorite peak widths decrease from a maximum of 2.2
mm to a low of about 1.1 mm. Kaolinite is commonly present in the samples
subjected to a low grade of diagenesis and interferes with the measurement of peak
width. It is of interest to note that the peak width values of chlorite and illite
converge and apparently intersect in the epizone. Presumably, this is a result of all
expanded layers being removed from both minerals.
The earliest formed chlorite porphyroblasts are the relatively rare (in the liter-
ature) l a polytype; that near the anchizone-epizone boundary is the high tempera-
486

Table 7-6
Chlorite Structural Formulas (Half Unit Cell). Listed in Order of Increasing Metamorphism. From
Weaver and Associates (1984).
Sample Si Al(tet) Fe Mg Al(oct) Fe/Fe+ Mg Oct Cations
Diag
OW-23 3.36 0.64 2.09 1.46 1.84 0.59 5.39
OW-7 3.27 0.73 2.00 1.36 2.00 0.60 5.39
Anch
OW-27 3.11 0.89 2.29 1.74 1.61 0.57 5.64
OW-29 2.90 1.10 2.50 1.88 1.44 0.57 5.82
ow-1 1 2.88 1.12 2.26 2.15 1.43 0.51 5.84
Ep i
OW-9(LS) 3.19 0.81 1.38 2.50 1.69 0.36 5.57

ture IIb polytype. Brown and Bailey (1962) believe the Ia chlorite is stable at slightly
higher temperatures than the more common Ib chlorite. They found that the stable
chlorite in “normal chlorite grade metamorphism” was the IIb form. The Ia chlorite
in OW-7 (high-grade diagenesis zone with incipient slaty cleavage) occurs primarily
as porphyroblast and was subjected to a maximum temperature of approximately
250°C. As this is the lowest temperature at which chlorite porphyroblasts formed, it
is unlikely there was a Ib chlorite precursor. The IIb chlorite in OW-3 (lower
epizone, slaty cleavage) was subjected to a temperature of approximately 330”C,
perhaps coinciding with the beginning of the anchizone ( - 280°C). The petro-
graphic data suggest the IIb polytype did not form directly from the Ia polytype.
The Ia chlorite may be relatively common. Most chlorites for which the polytype
has been determined have been subjected to burial temperatures less than 150°C or
in excess of 300°C. The Ia type may be common in physilites subjected to
intermediate temperatures.
A variety of techniques, x-ray diffraction, EDX, and differential solution, were
used to obtain information about the chemical composition of the chlorites. The
chlorite structural formulas in Table 7-6 were calculated from acid dissolution data.
For a discussion of the technique, which appears to be reliable, see Weaver et NI.
(1984).
The data from the structural formula are plotted on the classification grid of
Foster (1962)(Fig. 7-47). If OW-9, which is primarily an argillaceous limestone. is
neglected, the trend is for the amount of tetrahedral Al to increase significantly
(0.64 to 1.12) (diabantite to brunsvigite) and the Fe2+/R’+ to decrease slightly (0.60
to 0.51) with increasing metamorphic grade. Octahedral Fe is slightly more abun-
dant than Mg and A1 but does not dominate the octahedral layer. The chlorite
cement in sandstones from the zone of middle diagenesis have a much higher Fe
content than the higher grade chlorites in Table 7-6.
The area outlined by dashed lines shows the range of composition of 14 chlorites
from the metasedimentary lower greenschist facies of the Tennant Creek area.
Central Australia (Ramamohana Rao, 1977). The compositional range is similar to
487

1.00-
THURlNGlTE CHAMOSITE
I
0.80 - (11.0) (1.3) (0.61
--- - - - - - - - - - - - --- -
I BRUNGSV I G I T E 1
(u

+fx0.60- I
/
/ I
? R I P 1 DOL IT E

<
DIABANT I T E
11
z 0.40- I
0 9( Ls) I
(258) ‘\ (9.0) (1.3)
-- -- -L-- -
0.201
S H ER I DAN IT E CHL I NOCHLORE PENNITE
(16.8) (27.7) ( 5B 1
0 I I I I I I I

that of the Conasauga samples. Brown (1967) and Black (1975) have also noted that
in low grade metamorpluc rocks, the Fe/Fe + Mg ratios decrease with increasing
metamorphic grade, though in rocks of different compositions, the ratio is con-
trolled by bulk rock composition. A similar decrease in the Fe/Fe + Mg ratio has
been found in geothermal deposits (McDowell and Edlers, 1980). The chlorites in
these studies have a compositional range similar to those in the Conasauga, mostly
brungsvigite and diabantite. Chlorites formed in low grade metamorphosed volcanics
or in sandstones tend to have a wider range of compositions. Composition of
chlorites is apparently more dependent on the compositon of the bulk rock than
temperature (Black, 1974; Velde, 1977). This is evidently true of low temperature
chlorites as well. Because of the availability of Mg, Mg-rich chlorites form in
evaporite deposits and Fe-rich chlorites are commonly associated with low tempera-
ture iron ores and other Fe-rich rocks. However, within a given rock type, the
chemical composition of chlorite may be dependent on temperature.
The metamorphosed physilites characteristically contain chlorites in which neither
Fe nor Mg is dominant. This is presumably determined by the composition of the
typical physilite. The average Fe/Fe + Mg ratio of the Conasauga chlorite samples
from the physilites is 0.57. The average for 9 bulk samples is 0.69. Ramamohana
Rao’s (1977) 14 chlorite samples have an average ratio of 0.49 and the bulk rock
0.69. The value for Clark‘s (1924) average shale is 0.62. Black’s (1975) chlorite from
metasediments has an average near 0.60. Salton Sea geothermal chlorites formed
from sediments, largely sandstones, have an average ratio of 0.53. Thus, it appears
488

I I I I I
2 OCT.
6.0 - -
OCT.

5.8 - -
5.6

5.4
t
-

-
/z -
-
5.3

0.6 0.8 1.0 1.2 1.4 - TET. AL

Fig. 7-48. Relation of tetrahedral and octahedral Al to total Al in Conasauga chlorites. From Weaver and
Associates, 1984. Copyright 1984 Elsevier.

that chlorites formed by very low grade and low grade metamorphism of physilitic
+
rock commonly have a Fe/Fe Mg ratio in the range of 0.5 to 0.6, slightly lower
than that of the bulk rock.
Foster (1962) stated that most chlorites (presumably high temperature) have
more tetrahedral A1 than octahedral Al; however, the very low grade and low grade
metasediment chlorites in the Conasauga and most of those described in the
literature have less tetrahedral Al than octahedral Al, but the amount of tetrahedral
Al increases with increasing temperature. The total amount of A1 from both types of
chlorite is similar except for those with excessively large amounts of either Fe or
Mg. A value of 1.2 tetrahedral A1 generally distinguishes between high and low
temperature chlorites.
The Tet.Al/Oct.Al ratio of the Conasauga chlorites systematically in creases with
increasing metamorphism (Fig. 7-48) with a nearly ideal compensatory relation. The
temperature range is from approximately 250°C to 35OOC. The latter temperature is
from the high anchizone. Presumably a relatively small increase in temperature
would produce a high temperature chlorite with a Tet.Al/Oct.Al ratio larger than 1.
However, in the Conasauga, at higher temperatures ( - 400"),the chlorite is lost and
presumably converted, in part, to biotite. McDowell and Elder's (1980) data for
geothermal chlorites similarly show an increase in the Tet.Al/Oct.Al ratio with
temperature; chlorite converts to biotite at about 350°C.
The sum of octahedral cations in the Conasauga chlorite is less than 6.00 (Table
7-6). Values increase from 5.39 to 5.84 with increasing metamorphism (Fig. 7-48).
McDowell and Elder (1980) reported a similar trend for their geothermal chlorites.
Vacant octahedral positions are common in chlorites (Foster, 1962) and are ap-
parently necessary to obtain a charge balance between the octahedral and tetra-
hedral sheets. However, in the Conasauga and probably in most low temperature
489

chlorites the calculated octahedral vacancies are probably due to the presence of
some expanded water-bearing or partially water-bearing layers.
Both octahedral Mg and Fe increase with increasing metamorphic grade (exclud-
ing OW-9LS). Total A1 remains relatively constant, as tetrahedral A1 increases at
approximately the same rate as octahedral A1 decreases. Rather than chlorite being
destroyed and new chlorite created, it appears more likely that Mg and Fe
systematically replace the octahedral A1 and fill some of the vacant positions. This
need not require an increase in the tetrahedral charge but apparently there is an
increase. The epizone chlorite is from a dolomitic limestone and the composition is
presumably a reflection of the high Mg/Fe ratio of the bulk rock.
EDX analyses indicate the range of Fe/Mg peak height values (2.8 to 6.9) in the
upper diagenesis zone sample (OW-7) is greater than in the more metamorphosed
samples. This tends to confirm the presence of two types of chlorite (one containing
vermiculitic layers) as suggested by the x-ray data. With increasing metamorphic
grade and the conversion of chlorite to phengite, the spread in values decrease; the
Fe/Mg ratio decreases, and the amount of chlorite decreases. The Fe-rich chlorite
appears to be preferentially destroyed. It should also be mentioned that the chlorite
compositons calculated from the x-ray data are drastically different from those
obtained by chemical analysis. However, the peak height ratio 002/003 correlates
well with the Fe content of the chlorites.

Oxygen Isotopes

The application of oxygen isotope measurements in geothermometry is based on


the fact that when minerals are in thermodynamic equilibrium with one another the
relative proportions of oxygen isotopes are not the same in the different minerals.
The magnitude of the difference in the 1 x O / ' 6 0ratio between two minerals, in this
study quartz and illite-I/S, is a function of the nature of the minerals and of the
temperature of equilibration. For additional information on technique see Eslinger
et al. (1979) and Weaver, Eslinger, and Yeh (1984). The temperature values in Fig.
7-49 are questionable because only about half the calculated temperatures were
" reasonable".

The temperature values for the Conasauga are similar to those obtained from
other studies. Eslinger and Savin (1973b) analyzed a series of illitic rocks from the
Precambrian Belt Supergroup, Montana. Calculated quartz-illite isotopic tempera-
ture increased from 225°C in the shallowest samples (22 percent 2M illite) to 310°C
in the most deeply buried samples (60 per cent 2M illite). The 2M data indicates
recrystallization was not complete at 310°C. Extrapolation of the data indicates that
the illite would be fully converted to the 2M polytype at 400°C. The equivalent
temperature for the Conasauga would be approximately 360°C. On the basis of
oxygen isotope studies of various mineral pairs from metamorphic rocks from New
Caledonia, Black (1974) suggested that the boundary between weakly meta-
-
morphosed metasediments and the lawsonite zone ( anchizone) occurred at 250°C
and the lower boundary of the epidote zone ( - epizone) at 400°C.
490

4- L - - - - - - - - - -I-----
EPIZONE ANCHIZONE ',iA,E,,sis
100
0
I
10 20
I
30
I
40
I
50 60
I
70
L -1-
80
-1-
90
-
KM
Fig. 7-49. Maximum burial temperatures a5 a function of distance from OW-0. Line plot is based on
oxygen isotope content of quartz-mica and I/S pairs. Dashed rectangles are range of temperatures based
on the color alteration index of Ordovician conodonts (Harris and Milici. 1977). C.C. = -0.75. From
Weaver and Associates, 1984. Copyright 1984 Elsevier.

On the basis of the composition in fluid inclusions from fissure quartz crystals,
Frey et al. (1980) found a minimum temperature of 200°C for the beginning of the
anchizone. The higher-grade part of the anchizone had a minimum temperature of
270°C.
Winkler (1979) states that in pelitic rocks the epizone starts at 350"-370°C at 2
Kb pressure and biotite develops at 425"-450"C. In the Conasauga, the biotite zone
starts at about 420°C.
Numerous studies of thick sections of young sediments have shown that I/S,
with up to 20 percent smectite layers, persists to temperatures in excess of 200°C.
Cretaceous shale samples from the bottom of a deep well in Texas (samples and
temperatures supplied by L. Price), where the bottom hole temperature exceeds
300"C, have K.I. values similar to those for Conasauga samples with similar
temperature values.
The temperature data for mineral reactions in physilites during burial metamor-
phism are far from satisfactory, but the range of values is realistic. The diagenesis-
anchizone boundary temperature is at least 250"C, perhaps as high as 280°C
(present study). This temperature is considerably higher than the value of 200°C
used as the boundary between diagenesis and very-low-grade metamorphism in
zeolite bearing volcanic rocks (Winkler, 1979). It is quite likely the two boundaries
do not represent equivalent stages of metamorphism, as has been suggested by
Kisch (1974). The anchizone-epizone temperature boundary apparently occurs be-
tween 350" and 400°C. Some of the mineral reactions coincide with these boundaries,
491

but many are continuous reactions that extend over relatively wide temperature
ranges. Thus, in physilites the boundaries are arbitrary but perhaps reasonably
chosen. The problem is further complicated by the variation in mineral definitions.
Many believe that I/S is converted to illite by the beginning of the anchizone.
Actually, most anchizone "illites" are I/S with - 5 percent smectite. Some chose to
call this material illite. A true 100 percent non-expanded phase does not form until
near the beginning of the epizone. Some call this material illite; others call it
phengite. In the Conasauga the beginning of the epizone (K.I. = 1.5, 0.25" 20)
coincides with the pervasive recrystallization of the mica phase indicating that it is a
good natural boundary.
Many hydrothermal and geothermal areas have been studied in detail and the
temperatures of formation of various clay minerals are fairly well established.
Unfortunately, in most cases the host rock was some variety of basalt, and it is
unlikely this information is directly applicable to thick sections of physilites.
Fortunately, an excellent study has recently been completed of the sandstones and
shales of the Salton Sea Geothermal Field (McDowell and Edlers, 1980). The
sequence of mineralogic changes is similar to that observed in the Conasauga.
Fig. 7-50 compares the temperatures at which equivalent mineralogic changes
occur in the geothermal system and in the Conasauga (burial metamorphism). The
same sequence of changes occurs in each section. The temperature at which
equivalent changes occur is consistently 50"-70°C higher in the burial meta-
morphosed section than in the geothermal section. The difference in pressure of
approximately 2 Kb and the availability of water presumably account for the lower
temperatures in the geothermal system. The fact that the reactions in the Cambrian
Conasauga occurred at higher temperatures than equivalent reactions in the Holo-
cene Salton Sea samples would tend to suggest that time does not strongly influence
the reactions.

Potassium-Argon

Both the conventional K-Ar dating method and the 40Ar-39Ar method were used
to study K-Ar relationships in the Conasauga shales (Sedivy, Wampler, and Weaver,
1984). The results of the conventional analyses (Fig. 7-51) show that detrital
materials are dominant in the 2-44 p m fractions of specimens from the zone of
diagenesis. The apparent ages for this size fraction are 907 and 1097 m.y. for the
two least metamorphosed specimens and range downward to near 500 m.y. for other
specimens from the zone of diagenesis. Much of the detritus, at least in the silt size,
was presumably derived from Precambrian Grenville-age igneous and metamorphic
rocks.
For each specimen, the finer size fractions have K-Ar apparent ages less than that
of the 2-44 p m fraction (Fig. 7-51). For each particle-size range, the K-Ar apparent
ages become progressively smaller as the degree of metamorphism increases, but the
apparent ages of the finer fractions vary less than those of the coarser materials. For
the most thoroughly recrystallized specimens the apparent ages of the different size
492

OC I
450
GEOTHERMAL BURIAL METAMORPHISM-CONASAUGA
(McDowell & Elder, 1980)
%CHLORITE
(z-,,~) K.I. W.I. 28
400

350
NO CHL.; BIO. INC.
DEC. CHLORITE
EPJZONE
25 - 1.5 8.0 0.25

40
MINOR BIOTITE
- 1.9 4.5

300
’ ILLITE-MICA REXLIZED 50
NO SMECTITE LAYERS ANCH I ZONE
- 3.0 3.4 0.42

/ DIAGENESIS 60 - 3.7 2.0

0-10

Fig. 7-50. Comparison of temperatures of equivalent mineral zones of the Conasauga physilites and
Salton Sea Geothermal mineral zones. Geothermal temperatures are 5O0-7Oo lower than Conasauga
temperatures. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

fractions are all near 300 m.y., and it is clear that the K-Ar relationships in these
specimens have been determined by post-depositional processes. As the 40Ar-39Ar
age-spectrum method provides information about the distribution of radiogenic Ar
relative to K among different constituents of a sample, it will be discussed first.

40Ar-39Ar
The 40Ar-39Armethod of dating is based on the conversion of a known fraction
of the 39K present in a sample to 39Ar, by irradiation of the sample with fast
neutrons. Since the production of 39Arby irradiation of a sample is proportional to
the amount of 39K,and therefore to the amount of 40K(assuming normal K isotopic
composition), in the sample, it is possible to calculate an apparent K-Ar age for the
sample solely from the isotopic composition of the Ar released from the sample. The
principles of the 40Ar- 39Armethod and the procedures used in the irradiation and
isotopic analysis have been described by Dallmeyer (1979).
Dalrymple and Lanphere (1974) showed that an age spectrum displaying a
relatively uniform apparent age “plateau” for the majority of the fractions of Ar
493

I
M.Y. x 100
-

2 4 p m
-

0 BULK
-
C)
OC

0.2-2pm 400 -
300-
<0.2pm

200-

EPIZONE ANCHIZONE DlAGENESlS

released from a sample at higher temperatures is probably indicative of a fairly


simple, single-event, thermal history. These authors and others have also evaluated
age spectra from a wide variety of samples with complex geologic histories (Dairym-
ple and Lanphere, 1971) with varying degrees of success. Huneke (1976) pointed out
some of the problems involved.
The patterns of argon release during stepwise heating of Conasauga shale
samples gave information that was important in the interpretation of the 40Ar-39Ar
age spectra. Argon-release patterns were first obtained from materials that had not
been irradiated and then from irradiated samples.
The results of stepwise-heating experiments on unirradiated 2-44 pm fractions
are shown in Fig. 7-52, as graphs of the relative amount of radiogenic argon ( 40Ar* )
released in each step versus the temperature. Both types of physilite, the meta-
morphosed type represented by OW-3 and OW-9 and the diagenetic type rep-
resented by OW-19 and OW-31, have in common a significant release of 40Ar* at
about 600"-800°C in the heating schedule used, but only the diagenetic type shows
appreciable loss of 4"Ar* below 6OOOC and above 800°C. The differences in the
release patterns are related to basic differences in the mineralogic composition of
the various samples. All four samples contain appreciable amounts of illite and/or
mica, but only OW-19 and OW-31 contain appreciable amounts of I/S and
K-feldspars.
The argon-release patterns are consistent with the mineralogy of the samples and
previous studies of fine-grained materials. The 40Ar* retained in OW-19 and OW-31
to temperatures of 800°C and above was presumably mostly from K-feldspar and
494

100- I00 -
-C OW19 2-44 + OW9 2 - 4 4
+OW31 2-44 + OW3 2-44
80- 80 -
c
x
x
c
._
._ Ln
v) C
?,! 6 0 -
r
.c
-
.
._

' 5
01
40-

>
._
.- c
0
5 20-
(L

400 600 800 1000


Temp, 'C Temp, ' C
Fig. 7-52. Left: Radiogenic argon release patterns for diagenesis grade samples OW-19 and OW-31. 2-44
p m fractions. Right: Radiogenic argon release patterns for epizone samples OW-3 and OW-9, 2-44 p m
fractions. 600" - 800°C Ar is from detrital or recrystallized illite/phengite; in OW-19 and OW-31. low
temperature Ar is from I/S and high temperature Ar from K-feldspar.

perhaps large detrital mica flakes. The significant 40Ar* losses from these samples
at 400" and 500°C are thought to have been mostly from I/S. OW-3 and OW-9 are
composed largely of recrystallized illite-mica, which appears to lose little 40Ar*
below 600°C. OW-3 contains coarse recrystallized mica in porphyroblasts (not
present in OW-9), which may account for the retention of a rather large fraction
(34%) of its 40Ar* until the 800°C step.
Apparent-age spectra for the 2-44 pm fractions from OW-31 (diagenesis) and
OW-9 (epizone) are shown in Fig. 7-53. Over 95 percent of the "Ar evolved from
OW-31 (from the middle diagenesis zone) was released from material having an
apparent age greater than 600 m.y., with more than half of the 39Arcorresponding to
apparent ages in excess of 1000 m.y. This age spectrum is consistent with the release
of Ar from detrital, high temperature mica and feldspar. The argon released in the
temperature range 400" to 600°C may be attributed to illite and I/S.
In contrast, the age spectrum for metamorphosed sample OW-9, Fig. 7-53, is
more plateau-like, with apparent ages in the range 225 to 425 m.y., for all but the
first 2 percent of the 39Ar released. There is a general increase in the apparent ages
as the release temperature increases, but the values are in all cases less than the
depositional age (between 500 and 550 m.y.) of the sample. The generally low
apparent ages observed in the early heating steps for the Conasauga samples are
probably caused by redistribution of '9Ar by recoil. Age spectra for the 2-4 pm
fractions are basically similar to those for the 2-44 p m fraction, except for a minor
high temperature input from K-feldspar in the OW-31 coarse fraction.
The age spectrum for the 2-4 p m fraction of OW-9 is much like that of the 2-44
pm fraction but somewhat flatter. Apparent ages associated with more than 95
percent of the '9Ar released are in the range 345-405 m.y. The apparent ages from
the plateau-like portion of this spectrum average 375 m.y., which is virtually the
495

OW-31 2-44pm (TOTAL GAS AGE 1140 M.Y.1

0OW-3 2-44pm (TOTAL GAS AGE 351 M.Y.1


1500

600
.......................

P
> ...............:...................;.......'..;......':.;......~... ..
I
5
w' 1000
0
a
I-

7
2
W
K
2 ....................................
..................................
0
a ..................................
.............
..............
.............
500 -I 800 1000

1 I I I I I I I I
I
JL 100
RE(-EASED IN PERCFNT
Fig. 7-53. 40Ar-39Arapparent age spectra for OW-3l(diagenesis) and OW-9 (epizone) 2-44 p m fractions.
Uncertainties in apparent ages are indicated by the widths of the bars. Numbers above bars denote
temperature in "C. The well developed plateau for the OW-9 sample indicates nearly complete
recrystallization of the micas. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

same as the conventional K-Ar ages obtained for the 0.2-2 pm and 2-44 pm
fractions of OW-9. The petrographic and x-ray data indicate that the epizone
sample OW-9 is largely recrystallized mica, apparently phengite. The near-plateau in
the age spectrum of both size fractions suggests that the recrystallization occurred
between 300 and 400 m.y. ago (average near 375 m.y.) with little subsequent
disturbance of K and Ar. Either recrystallization occurred over a long period of
time or recrystal lization is incomplete and some layers with inherited K and Ar are
present. The K-Ar data suggest the latter explanation is more likely. Completely
recrystallized mica has an age of approximately 300 m.y.
Age spectra of two bulk samples (OW-11, OW-24) from the anchizone indicate
about 30% of the '9Ar was released from material having an apparent age of
approximately 300 m.y. (recrystallized) and 70% had higher, partially inherited ages.
The older ages ranged from only 350 to 525 m.y., indicating there was considerable
reorganization of the K in the micas that were not completely recrystallized.
The only major differences among the age spectra for OW-9 (2-44 pm), OW-11,
and OW-24 occur at the highest release temperatures, which confirms the steady
decrease in the influence of detrital materials with increased metamorphic grade.
496

Conventional K-Ar Ages

Recent oceanic and deltaic clays (Hurley, 1966) always have K-Ar ages that are
“too old”, indicating that detrital illites and micas with in herited Ar are present.
This would suggest that most shales should have K-Ar ages in excess of their
stratigraphic ages. Analyses of illite-rich shales (Goldich et al., 1959; Evernden et
a/., 1961; Hurley et a/., 1963) showed that the K-Ar age is commonly greater than
the stratigraphic age, but for many shales the K-Ar age is similar to or less than the
stratigraphic age. Evernden et al. (1961) found a number of shales with K-Ar ages
that are 15 to 150 m.y. less than the stratigraphic age. These samples were primarily
from thin shale beds in limestone.
The < 2 pm fractions of eight lower Paleozoic shales analyzed by Hurley et a/.
(1963) have K-Ar ages approximately 15 percent less than the stratigraphic age and
about the same as the age values obtained from glauconites of similar stratigraphic
position. They found clay-size fractions from Mesozoic and Cenozoic shales to have
K-Ar ages considerably greater than the stratigraphic ages. The same relationship
has been demonstrated for Gulf Coast Tertiary shales by Weaver and Wampler
(1970), Perry (1974), and Aronson and Hower (1976), though the studies of
Cenozoic shales indicate the K-Ar age decreases with depth as smectite converts to
I/S (Fig. 7-4). Many Tertiary and Mesozoic glauconites have K-Ar ages very close
to the stratigraphic ages although illitic shales of similar age are much too old. The
K-Ar data indicate that Paleozoic illites have undergone considerably more di-
agenetic modification than glauconites. This may simply reflect the fact that most
Paleozoic illitic shales, largely geosynclinal, have been buried to greater depth and
exposed to higher temperatures than most glauconites. Glauconites are more char-
acteristic of the platform facies. Also, glauconites are likely to be destroyed during
advanced stages of diagenesis.
Fig. 7-51 shows how the K-Ar apparent ages of material in three size ranges from
the Conasauga samples vary with distance from the OW-0 site (site of maximum
metamorphsm). For all three size ranges the K-Ar ages decrease with increasing
degree of metamorphism and converge to K-Ar apparent ages near 300 m.y. The
K-Ar apparent ages of the 2-44 p m fractions change from near 700 m.y. to 320 m.y.
The change for the 0.2-2 pm fractions is from about 500 m.y. to near 300 m.y., and
that of the < 0.2 pm fractions is from about 360 m.y. to about 280 m.y. The highest
age values observed for each size range were for samples from the middle diagenesis
zone in Tennessee. As with the values of the Kubler Index, the K-Ar ages change
most rapidly near the metamorphic zone boundaries. The decrease in K-Ar ages
with increasing metamorphic grade is presumably related to an increase in tempera-
ture.
The stepwise-heating experiments described in the preceding pages show that
with increasing degree of metamorphism the phases containing argon that would be
released in the laboratory at low temperature (I/S) and at high temperature
(K-feldspar and biotite) were destroyed and the resulting phengite-mica contains
argon that is released at intermediate temperatures (500-800°C). These reactions
should have progressed systematically with increasing burial temperature and thus
49 7

-0 10 20 30 40 50 60 70 80%
WEIGHT
Fig. 7-54. Relation of amount of material in three size fraction to K-Ar apparent age. From Weaver and
Associates, 1984. Copyright 1984 Elsevier.

should have led to the systematic decrease in K-Ar apparent age that has been
observed for the 2-44 pm fractions of the Conasauga. The 40Ar-'9Ar age spectra of
anchizone samples indicate that detrital materials remaining, retained much of their
radiogenic argon during the very low-grade metamorphism. Therefore, the relatively
systematic decrease in K-Ar ages of the 2-44 pm fractions with increasing degree of
metamorphism must be largely a consequence of recrystallization rather than of
thermal degassing, except perhaps for samples from the epizone.
Fig. 7-54 shuws how the K-Ar age values relate to the percentages of material in
each of the three size fractions. As the relative amount of 2- 44 pm material
increases, the K-Ar age decreases. For the two finer size ranges, the K-Ar age
increases as the percentage of material increases. These opposing trends are a
consequence of an increase in relative amount of material in the 2-44 p m size range
at the expense of material less than 2 p m as recrystallization proceeded. (The sum
of these mass fractions does not equal 100 percent in any case, because material
coarser than 44 pm was included irr the calculation of the mass fractions.) That the
relative amount of material coarser than 2 pm increased with increasing metamor-
phic grade is presumably due largely to crysfal growth. (An increasing tendency to
form aggregates with increasing degree of metamorphism could be a contributing
factor, but this alone could not account for the large decrease in apparent age for
the 2-44 p m fractions.) The fact that the apparent age of the remaining fine
material is lower in the cases where the amount of such material is low indicates
498

TN

0 RESIDUE
0 BULK PHYSILITE‘

EPIZONE ANCHIZONE DlAGENESIS


2 I I I I I I I I

Fig. 7-55. Comparison of K-Ar apparent ages of bulk carbonate residues and bulk physilites in relation
to distance from OW-0. From Weaver and Associates, 1984. Copyright 1984 Elsevier.

that some material recrystallized but remained fine, or that additional potassiuni
entered fine material, or that some argon was lost from the fine material by
diffusion. In the latter case one would expect to find apparent ages of less than
- 300 my., which were not found.
Fig. 7-55 shows the K-Ar apparent ages of the carbonate residues (from thin
carbonate layers interbedded with the physilites) and those of the total physilite
samples as a function of distance from OW-0. In the zone of diagenesis, the
apparent ages of the carbonate residues are less than those of the physilite. The
apparent ages for the two rock types are similar in the anchizone and epizone. The
K-Ar age of a residue (OR-7) from the medium grade diagenesis zone is 956 ni.y.,
which is similar to the age of the total physilite samples from the same zone. Thus,
the K-Ar apparent ages of the “starting” materials (illite, I/S, K-feldspar) for the
carbonates and physilites were similar.
The K-Ar ages of the residues decreased to 400-450 m y . in the zone of middle
diagenesis. Part of the reduction in K-Ar age is apparently due to the destruction of
some K-feldspar and the incorporation of K into the I/S. As with the physilites, the
decrease in apparent age is presumably due to the formation of new mica layers
rather than to thermal degassing. As also indicated by the K.I. data, the develop-
ment of mica layers starts at a lower temperature or proceeds more rapidly in
carbonate rocks than in physilites. The K-Ar apparent ages of the carbonate
residues do not decrease in passing from the diagenesis zone into the anchizone as
they do for the physilites. This is probably because formation of the illite-mica is
more nearly completed during late diagenesis in the carbonate rock. The K-Ar data
499

I I I
-=C
50

1W

150

200

250

300

350

20 40 60 80 100 %

%'REDUCTION K-Ar AGE - WHOLE SAMPLE


Fig. 7-56. Percent reduction in apparent K-Ar age of whole sample of Tertiary physilites (Aronson and
Hower, 1976) and Conasauga physilites as a function of temperature. C.C. for Conasauga = 0.95. 690
m y . is the assumed detrital age and 270 m.y. (OW-1) the age when recrystallization was completed,% Age
Reduction = [420- (Apparent age-270) X 100]/420. From Weaver and Associates, 1984. Copyright 1984
Elsevier.

indicate that near total recrystallization and resetting of the K-Ar clock apparently
occurred in both rock types in the epizone.
If we assume that the K-Ar age inherited by the Conasauga formation from its
detrital constituents would have been about the same throughout the formation
(before diagenesis), and that recrystallization has been the dominant cause of argon
loss from the rock during diagenesis and very low-grade metamorphism, then the
degree of recrystallization of K-minerals in the rock should be almost linearly
related to the reduction in K-Ar age. Fig. 7-56 shows the relative reduction in K-Ar
age for whole-rock samples of the Conasauga versus the maximum temperature
500

encountered during burial. The reduction has been calculated relative to an average
of the apparent ages of the least-metamorphosed samples from Tennessee (690
n1.y.). Furthermore, the relative reductions are calculated as they would have
appeared 270 m.y. ago, soon after the end of the diagenetic/metamorphic changes,
so that we may compare the relationship between temperature and apparent age
with that observed for samples from a Gulf Coast well that contained sediments of
Miocene to Oligocene age (Aronson and Hower, 1976). The data of Aronson and
Hower (indicated schematically in Fig. 7-56) show a relatively rapid rate of decrease
in apparent age at the beginning of diagenesis. Once a regular I/S phase had
developed, the rate of reduction in K-Ar age became less. The least-metamorphosed
Conasauga samples contain an I/S physil with approximately 80 percent illite
layers, similar to the clay in the lower portion of Aronson and Hower’s well. The
rate of K-Ar age decrease in the Conasauga samples is similar to that for the
Tertiary samples that contain regularly interstratified I/S. These data suggest that
once a regular mixed-layer I/S is formed the reorganization of the K in the bulk
sample procePds regularly with increasing temperature.
Once an ordered mixed-layer phase is formed, the ratio of I/S layers may remain
constant but the degree of order increases. This requires that the K be mobile. The
mobility of K, and therefore its disassociation from radiogenic argon, is also
indicated by the continued increase in amount of K per illite layer with increasing
metamorphism. In the Conasauga, in the less-than-0.2 p m size range, the mica in
the recrystallized epizone contains 0.93 K per layer 0,,(OH)2, but the 1/S, with 10
percent smectite layers, contains only 0.69 K per illite layer. It is evident that K
continues to be added to the mica phase long after the ordered I/S phase forms and
this would cause a continuing decrease in the K-Ar age whether or not there was
any loss of argon by diffusion.
The K-Ar data show a general consistency between the distribution of radiogenic
Ar relative to K in the Conasauga shale and various other indicators of the degree of
diagenesis/metamorphism, but do the data provide useful information about when
the diagenesis/metamorphism occurred?
The stratigraphic and tectonic history of the southern Appalachians provides an
independent basis for estimating the burial history of the Conasauga formation. The
development of an Ordovician-Silurian clastic wedge indicates that burial tempera-
tures for the Conasauga should have progressively increased until about 400 m.y.
ago. There is very little Devonian sedimentary rock in this portion of the Appa-
lachian basin, but it is possible that late Devonian thrust faulting may have caused
even deeper burial of part of the Conasauga (Hatcher, 1978). Portions of the
Conasauga formation likely reached a maximum temperature late in the Paleozoic
Era as a consequence of the overthrusting (Alleghanian orogeny). Erosion of the
orogenic belt should then have led to fairly rapid cooling and termination of the
diagenetic/metamorphic changes.
The stratigraphic/tectonic history outlined above implies that maximum temper-
atures in the Conasauga formation should have been reached between about 400
m.y. ago and about 250 m.y. ago. All the apparent ages for material < 0.2 pm in the
Conasauga samples (Fig. 7-51) correspond to this period of time. A map of these
501

K-Ar <0.2pm

Fig. 7-57. Distribution of K-Ar apparent ages of the < 0.2 p m fraction. Values systemalically decrease to
the southeast. Compare with K.I. map. 7-17. From Weaver and Associates. 19x4. Copyright 1984
Elsevier.

apparent ages, Fig. 7-57, shows a clear pattern in the apparent age distribution that
corresponds closely to the degree of diagenesis/metamorphism. The apparent ages
from the zone of diagenesis range from 390 to 350 m.y. (corresponding to the
Devonian Period). Southeastward in the anchizone the values range from 333 m.y.
to a little less than 300 m y . near the boundary with the epizone. Apparent ages of
the < 2 p m material in the epizone are about 275 my., the lowest K-Ar ages
obtained in this study.
502

2 I/S had been effectively an open system for K and Ar while


If the ~ 0 . pm
diagenesis proceeded, as suggested by the Gulf Coast Cenozoic studies, then the
apparent ages of this fine material would indicate approximately the time when
diagenesis ended.
The K-Ar ages of the < 0.2 pm material appear to fit this scenario. The apparent
ages of the fine material from the area of moderate diagenesis, which correspond to
the Devonian Period, may be related to a thermal maximum caused by deep burial
under the Ordovician-Silurian clastic wedge and/or by the influence of the Acadian
orogeny. The ages of the anchizone samples and epizone samples, which correspond
to Carboniferous and early Permian time may be related to overthrusting and
deeper burial during the Alleghanian orogeny. These latter samples have been
largely recrystallized, particularly those from the epizone, and their K-Ar ages
apparently may reflect the effects of two periods of orogenic activity.
The materials > 0.2 pm from the zone of diagenesis generally have retained some
inherited argon since almost all the apparent ages are near to or greater than 500
m.y. The same is true of the 2-44 p m fractions of samples from the anchizone. It is
probably that virtually all of the inherited argon was lost from materials of the
epizone. The 0.2-2 pm fractions of anchizone samples have, with one exception,
apparent ages between 360 and 405 m.y. This range is almost the same as the range
of ages seen in the near-plateau portion of the age spectrum of OW-9, 2-4 pm. If
these materials are largely new materials, with little or no inherited argon remaining,
then the peak of diagenetic/metamorphic activity in the anchizone must have been
reached in the Devonian Period. If the peak in activity occurred during the
Carboniferous Period, as the apparent ages from the ~ 0 . 2pm fractions might
suggest, then the 0.2-2 pm material must have retained a rather uniform small
fraction of earlier-accumulated radiogenic argon.
Hunziker (1979) described results of several K-Ar studies of the < 2 pm
fractions of Alpine physilites that were affected by Alpine deformation and anchi-
zone to epizone metamorphism. Illite from an unmetamorphosed section had
inherited argon that reflected the provenance of the sediments. Most of the
illitic-phengitic micas from the anchizone and epizone had K-Ar ages that were
interpreted as indicating the dates of metamorphism, but some samples from the
lower anchizone had K-Ar ages that were interpreted to be mixed ages. The mixed
ages, however, were much closer to the ages associated with metamorphism than to
the ages of the unmetamorphosed material, indicating that the < 2 pm material
from the lower anchizone had been largely recrystallized. Hunziker (1979) also
noted that < 2 pm micas and 2-6 p m micas tend to show no difference in K-Ar age
if they are from the same tectonic or metamorphic phase, and that no evidence had
been found for diffusion of argon o u t of the 2 pm grains. The K-Ar results from the
Conasauga are in general agreement with the relationships observed in the Alpine
region. In particular, the consistency of dates near 375 m.y., from 0.2-2 pm
materials of the anchizone and the 40Ar-”Ar ages of the 2-4 pm material from
OW-9, also near 375 my., suggests a peak in very-low-grade metamorphism in the
Devonian Period. The near convergence of the K-Ar ages of the various size
fractions of epizone samples OW-1 and OW-0 near 300 m.y. indicates the strong
503

influence of the Alleghanian orogeny on these samples. These rocks are slates (or
incipient schists) in which most of the coarser material, as well as the finer material,
has undergone recrystallization. There is a very close similarity between the two
samples in the way apparent age varies with particle size, but two quite different
interpretations of this pattern may be made:
1) Recrystallization peaked near the end of the Carboniferous Period and is
reflected most clearly in the apparent ages of the finest fractions (270 and 278
m.y., also 282 m.y. for OW-9). Coarser materials have retained either a little
inherited argon or some argon that had accumulated in earlier-formed diagenetic
material.
2) Recrystallization peaked before the beginning of the Pennsylvanian Period (320
or more m.y. ago) at temperatures high enough that argon was not effectively
retained in epizone materials. Slow cooling after this time led to closure of the
physils to argon loss in order of decreasing particle size.
The apparent ages obtained from different size fractions of epizone samples
OW-1 and OW-0 can be understood, by reference to Giletti's (1974) argon diffusion
studies, if maximum temperature in the epizone was reached about 300 m.y. ago
with fairly rapid cooling thereafter. The thick, early-Pennsylvanian clastic sequence
that accumulated just northeast of the Ridge and Valley province implies an
early-Pennsylvanian peak in orogenic activity (overthrusting) in this portion of the
Appalachians. The apparent ages somewhat greater than 300 m.y. for materials > 2
p m (flake diameters on the order of 10 pm or more) can be understood because
temperatures around 350°C would not be sufficient to drive all the argon out of
mica that had formed earlier, and it is unlikely the recrystallization was 100%
effective in separating K and Ar. This conclusion is in accord with information
about white micas form the Alps obtained by Hunziker (1979) and his colleagues.
But the fine material could not be expected to have retained argon well until the
temperature was decreasing. The 275 m.y. ages for the ~ 0 . 2p m fractions may
correspond to a closure temperature somewhere between 300" and 200°C.

Separating the Effects of Temperature and Tectonics

The regional mineralogic changes are complex, but the major changes form a
systematic pattern (Fig. 7-58). A basic subdivision can be made on the basis of the
presence or absence of detrital biotite. In Georgia this boundary roughly parallels
the trace of the Great Smoky and Emerson faults.
Metamorphic alteration of biotite is first observed in the zone of high-grade
diagenesis where some of the brown biotite altered to a green mica, presumably
some form of phengite, and to a lesser extent white mica and chlorite. All three
minerals started to form at approximately the same time (temperature), but their
relative abundance changed with increasing grade of metamorphism. The phengite
zone is followed by one in which chlorite, primarily as porphyroblast, is predomi-
nant, but green phengite and white mica are also present. The disappearance of
biotite roughly coincides with the increase in white mica at the expense of chlorite
504
,

Fig. 7-58. Map showing mineral zones in the Conasauga. Weaver and Associates, 1984. Copyright 1984
Elsevier.

and phengite (Fig. 7-58). The overall reaction is for biotite to alter to white phengite
through the intermediary stages of green phengite and chlorite. Illite and I/S may
be involved in the reaction.
Fig. 7-59 shows the mineral zone boundaries, the K.I. boundaries of the < 2 pm
10 ,& material, and the boundary for the first occurrence of crenulations (start of
slaty cleavage). The initial development of crenulations coincides with the beginning
of the chlorite zone; however, good slaty cleavage occurs only in the white mica (no
505

Fig. 7-59. Map showing mineral zones beginning of crenulation cleavage (slate), and K.I. contours (solid
lines). Mineral zones and cleavage are related to tectonic movement along the faults; the K.I. values of
the matrix mica are related to depth of burial (temperature). From Weaver and Associates, 1984.
Copyright 1984 Elsevier.

biotite) zone. It is overprinted by a second set of crenulations in the recrystallized


mica zone. The first occurrence of paragonite coincides with the slate boundary; it
increases in abundance towards the Great Smoky fault.
The K.I., W.I., and 002/001 (illite-mica) contours cross the mineral zone con-
tours and the fabric contour at a relatively high angle except for the K.I. 5 contour,
which coincides with the beginning of the phengite zone. The mineral zones are
based on changes in the relatively coarse porphyroblasts or mica flakes, whereas the
changes in the < 2 pm micas reflect changes in the matrix material.
The contours of the porphyroblast mineral zones and the distribution of the slaty
cleavage are roughly parallel to the trace of the boundary faults. This suggests
tectonic forces are responsible for these features. The fact that the contours
outlining changes in the character of the matrix illite-mica are aligned at nearly right
angles to the fault and roughly parallel to the contours representing thickness of
overburden suggests they are more related to thermal effects than dynamic effects.
The relation of porphyroblasts to tectonic metamorphism is apparently related to
the initial development of crenulations. As demonstrated, the porphyroblasts appear
506

to have developed in the void space created in the crenulations or by replacing


biotite that, by position, had been most deformed. Pressure solution and diffusion
were more pronounced under such conditions. Farther from the fault where crenula-
tions and flake rotation was less pronounced, a relatively small number of biotite
flakes were slightly altered to produce thin, elongate porphyroblasts, largely green
phengite. Most of these are elongate parallel to bedding. Flakes that are inclined to
bedding tend to alter to white mica. The development of slaty cleavage coincides
with the development of chlorite porphyroblasts. Under the higher stress state (near
the fault) the biotite altered to chlorite more so than phengite. The tectonic
porphyroblasts were formed during the early stage of the development of crenula-
tions and then progressively destroyed as deformation and temperature increased
and a true slaty cleavage developed. Even with high stress the temperature ap-
parently had to be around 250°C before porphyroblasts formed.
The K.I. contours characterize the trend of thermal metamorphism which pre-
ceded or probably was partially contemporaneous with the period of tectonic
metamorphism. During the period of thermal metamorphism the illite and I/S
recrystallized, decreasing the number of expanded layers and increasing the K
content.
Fig. 7-59 indicates that the development of slaty cleavage does not have a specific
relation to the K.I. Crenulations, incipient slaty cleavage, start at K.I. values
ranging from 4 (late diagenesis) to 2 (late anchizone). On the basis of the previous
discussion this is to be expected. The recrystallization of illite is based on tempera-
ture and slaty cleavage on deformation. They need not coincide.
The percentage of coarse chlorite, largely porphyroblasts, systematically de-
creases from north to south (Fig. 7-45) as does the K.I. of both the mica and the
chlorite. This suggests the destruction of chlorite (alteration to white mica) is also
dependent on an increase in temperature.
The slaty cleavage and the chlorite porphyroblasts apparently developed due to
thrusting during the Acadian orogeny (Devonian). Little sedimentation occurred in
this area after the orogeny. Thus, if the alteration of chlorite to white mica was
dependent on temperature, the increase in temperature had to be due to an increase
in overburden caused by overthrusting during the Allegheny orogeny. Some of the
changes in the matrix illite and I/S presumably occurred before thrusting, but the
K-Ar data indicate the final recrystallization was related to the time of thrusting.
The increase in overburden presumably was also responsible for the increase in
pressure solution, mineral orientation, and decrease in thickness of cleavage lamel-
lae.
If the northeast-southwest trace of the Great Smoky fault in Tennessee is
extended into Georgia, the overburden would increase to the southeast. The trend of
increased overburden parallels the trend for the mineralogic and textural changes
that are likely due to increasing temperature. In the southern part of the area, the
pattern was modified and presumably accented by the northern movement along the
Emerson fault during the late Pennsylvanian (Allegheny orogeny).
507

Stability Range of Minerals

Fig. 7-60 summarizes the occurrence and relative abundance of the major
minerals in the Conasauga as related to the Kubler Index and Weaver Index of the
illite-mica and the oxygen isotope temperatures. The general stability range of the
various minerals is similar but not identical to that reported by Frey (1974) and
Kubler et al. (1979) for minerals in pelitic rocks of the central Alps.
Within the area of study kaolinite and K-feldspar are the first two minerals to
disappear (Fig. 7-58). Kaolinite is not present in shales with a K.I. less than 3.8 and
a W.I. larger than 1.8 (high-grade diagenesis). The medium-grade diagenesis shales
were not studied in enough detail to establish what happened to the kaolinite. There
is no significant increase in chlorite related to the decrease in kaolinite.
K-feldspar in any appreciable amount is not present in physilites with a K.I. of
approximately 3.5 and a W.I. near 2.0. This is close to the accepted boundary
between diagenesis and very-low-grade metamorphism (anchimetamorphism). The
K presumably was involved in the conversion of smectite layers to illite. At higher
grades of metamorphism K was presumably obtained from the destruction of biotite
and perhaps other detrital micas.
Glauconite is relatively abundant in shales with a K.I. larger than 4.0 and W.I.
less than 1.5 but was not found in shales with K.I. less than 4.5. Frey (1973) reports

GLAUCONITE

-
d
KAO.

K-FLD.

I
<140
I
-280
VEIN CHL.

I
-360°C

Fig. 7-60. Stability range of some minerals in the Conasauga physilites. P - phengite. B = biotite. From
Weaver and Associates, 1984. Copyright 1984 Elsevier.
508

the paragenesis of glauconite to stilpnomelane in very-low-grade metasediments. N o


stilpnomelane detected in the Conasauga.
The proportion of expandable layers in the I/S decreases systematically with
increasing grade of metamorphism (Fig. 7-19). Relatively obvious I/S ( - 10 percent
smectite) persists until the beginning of the anchizone; however, expanded layers are
still present until the beginning of the epizone (Fig. 7-22). This suggests that illite
also persists to the beginning of the epizone. It is not known when illite alters to
phengite. As the K-Ar ages of the finer size fractions are relatively constant through
the entire anchizone (Fig. 7-51), it is unlikely that there is any systematic conversion
of illite to phengite through this interval. Though the age decrease in the coarse
fraction reflects the destruction of biotite, and perhaps coarse illite, and the growth
of white mica, the K-Ar values suggest most of the matrix illite is probably not
converted to phengite until pervasive recrystallization occurs (epizone). This also
implies that crystal growth is not the dominant mechanism for the development of
slaty cleavage.
Detrital biotite is destroyed over a relatively wide metamorphic range. Near the
Cartersville fault it is not found (optical observation) in rocks with a K.I. smaller
than 3.0 and a W.I. higher than 4.0; however, farther from the fault it persists until
the beginning of the epizone.
During the late stage of diagenesis biotite laths alter to green mica ovoids.
apparently phengite, and to a lesser extent white mica which is also probably
phengite. At a slightly higher stage of diagenesis biotite alters to chlorite. Green
phengite also replaces quartz and calcite fossil fragments and perhaps chlorite. I t is
present in nearly all anchizone rocks and is absent in the epizone rocks. It alters to
white mica. The green phengite has the composition of a low Al phengite with a
relatively high Fe content. The Mg/Fe ratio increases from 0.4 for the green
phengite to 4 for the white phengite.
Biotite can be observed altering to green and white mica or white mica and
chlorite in the same thin section. This indicates composition on a micro-scale is
more important than temperature and pressure in the late diagenetic-anchizone
stage of metamorphism. Assuming Al is conserved, the alteration of biotite to
phengite basically involves the loss of Fe and Mg from the octahedral sheet (the
Si/Al ratios are similar, - 1.6). The alteration of biotite to chlorite primarily
involves the loss of Si, Fe, and K. The two reactions would seemingly require
different fluid environments.
The white mica, which is apparently a low Fe phengite, crystallized continuously
from the late stage of diagenesis up to well into the greenschist facies. It is unlikely
that this mica has a constant composition. Most likely the composition varies as a
function of temperature and pressure. This would be particularly true of the newly
crystallized mica. For kinetic reasons mica formed at a lower temperature may not
change composition as the temperature increases until the stage (epizone) is reached
where there is complete recrystallization of the micas. It is difficult to interpret the
chemistry of white micas from EDX spectra because they are frequently closely
associated with biotite or chlorite layers.
Velde (1965) suggested that when the starting rock is an argillaceous sediment the
509

following reaction occurs:


illite + chlorite + K-feldspar + H,O -+ phengite + chlorite (high Al) + quartz
The reaction in the Conasauga appears to be:
illite + chlorite + quartz -+ phengite + chlorite
This is probably the reaction which occurs in the epizone when the matrix illite is
recrystallized. But the phengite in the porphyroblasts presumably forms primarily
from biotite. The reaction may be:
+ +
biotite illite quartz + phengite chlorite +
or, simply:
biotite + phengite + Mg, Fe.
Paragonite first occurs in high-grade diagenetic rock (Chatsworth), which is a
lower grade for its first occurrence than reported by Frey (1970, 1974). It is present
in the anchizone and in the lower portion of the epizone. At Chatsworth, paragonite
was found only in the > 44 pm fraction. This coarse fraction is composed largely of
chlorite porphyroblasts. Frey (1969) proposed the reaction:
+
Na-bearing illite paragonite phengite H,O.
-+ +
The apparent association of paragonite with chlorite porphyroblasts suggests that in
the Conasauga the reaction was:
biotite + Na-feldspar -,chlorite + paragonite K +
Na-feldspar is relatively abundant in the paragonite-bearing samples. The Na
content of these rocks is appreciably higher than in the samples without paragonite.
The presence of Na may be the controlling factor but the distribution pattern
suggests deformation, as opposed to temperature, is also a factor. The presence of
some paragonite in the dark cleavage lamellae confirms that some crystal growth
has occurred during the development of slaty cleavage.
In the middle stage of diagenesis, secondary chlorite is present in sandstone
laminae, where it replaces quartz and muscovite. The replacement chlorite has less
A1 and Mg and more Fe than that formed at higher temperatures. Detrital chlorite
is also present.
In the physilites, obvious secondary chlorite (porphyroblasts) is first observed in
the late diagenesis zone (K.I. 3.7, W.I. 2.0). The first occurrence of chlorite crosses
the K.I. and W.I. contours and ranges up to K.I. about 2.0 and W.I. of 4.0. I t is
related to the first appearance of kinks and crenulations. I t forms from biotite and
from phengite which formed from biotite, largely in voids and strain areas produced
by tectonic movement. The lowest grade of chlorite is diabantite which has a
relatively high Fe content and low tetrahedral Al. The chlorite contains some
expandable layers apparently in the form of a Mg-rich corrensite-like material. The
expandable layers decrease with increasing metamorphism but persist into the
epizone.
The EDX spectra are quite variable but in general suggest the Mg/Fe ratio is
higher in the chlorite than in the assocaited biotite. The Mg/Fe ratio and tetra-
510

hedral Al of the chlorite increases with increasing metamorphic grade. The presence
of pyrite patches in some chlorite pods suggests some of the Fe-rich chlorite alters
to Mg-rich chlorite; however, most of the pods are systematically dissolved by
pressure solution and replaced by phengite. The increase in the Mg content of the
chlorite appears to be largely due to the formation of a second and perhaps third
generation of chlorite which grew in pressure shadows adjacent to relatively insolu-
ble grains, particularly pyrite, and as vein and fracture fillings. Biotite is no longer
present so the Mg and Fe apparently come from the “cleaning-up’’ of the phengite
and illite. Some of the Mg could have been derived from the associated carbonate
rocks. The chlorite in the carbonate rocks is especially rich in Mg. At a temperature
near 4OO0C, the phengite and what little chlorite was present reacted to form biotite,
and phengite. Fig. 7-23 suggests minor biotite first started to form at the beginning
of the epizone.
Elongate quartz grains that have in part formed by crystallization are first
detected in the zone of high-grade diagenesis (Chatsworth) where chlorite
porphyroblasts first form. By the beginning of anchimetamorphism pressure solu-
tion has caused most of the quartz grains to become alongate. In addition, small
lenses of chert-like quartz crystallize. The growth of secondary quartz is more
closely related to the development of slaty cleavage than to the K.I. and W.I..
indicating that, at this stage, pressure is more a factor than temperature. With
increased metamorphism, quartz grains continue to dissolve, and elongate lenses.
beards, and veins develop. Separated grains are very angular and have a tabular to
plate shape.
Pyrite is present in all samples. It most commonly occurs as fine disseminated
particles in the diagenetic-grade physilites. Aggregates are more abundant in the
anchizone. Samples from the epizone contain tabular hematite crystal. Some contain
pyrite cores. Apparently with increased metamorphism both Fe and S have been
mobilized and concentrated.
Small idiomorphic tourmaline crystals form within the epizone contempora-
neously with the recrystallization of the matrix illite-phengite. This suggests that the
illite-mica retains an appreciable portion of their boron until they are completely
recrystallized.
I do not have the time nor space to discuss the bulk chemistry of shales and
slates, but one of the major concepts of the metamorphic geologist explains some o f
the concerns of the diagenesis geologist. Weaver and Beck (1971) noted that the
Gulf Coast muds would never end up with a composition similar to that of the
Paleozoic shales unless ions were added from deeper in the sedimentary pile. The
metamorphic geologist has another explanation which, on the basis of the Con-
asauga study, seems reasonable.
Solution features, contorted veins, and other features indicate there can be a 25
percent or more volume reduction during low-grade metamorphism and the devel-
opment of slaty cleavage (Hobbs et ul., 1976). This is believed to be due largely to
the dissolution of quartz (and in some instances carbonate minerals).
The SiO,/AI,O, ratios (Fig. 7-61) of the Conasauga samples are scattered in the
zone of diagenesis but show a systematic decrease through the anchizone and
511

*-
ApZ03
-
MISS.
7-
DELTA

6- -
TERT.
5-
H
X TN
-

4-
MESO-CENOZOIC
' -
PRE C.PALEoZolC
SLATE
SLATE '
PALEOZOIC
0
DEEP
SEA
rn
3- 0 0
0
Y
-
0
EPlZO& ANCHIZONE DlAGENESIS
I I I I I I I I I,
2O 10 20 30 40 50 60 70 80 90 KM

epizone. Petrographic studies show that detrital quartz was dissolved (became more
elongate) in this interval and x-ray data indicate a significant decrease in feldspar.
On the basis of the Si02/A1,0, ratio of the illite-phengite and chlorite and the
decrease in chlorite content with increasing metamorphism, the SiO,/Al,O, ratio
should increase from approximately 1.8 to 2.2 with increasing metamorphism. Thus,
the decrease in the ratio for the bulk rock must be due to the loss of quartz and
feldspar.
Fig. 7-61 also shows the general decrease in the SiO,/Al,O, ratio of muds,
shales, and slates with increasing age or more properly with increasing grade of
metamorphism. It should be pointed out that the trend could be duplicated with the
proper mixture of deltaic and deep sea clays.
The average K2O/Al20, ratio of Mississippi Delta mud samples (0.219) is
essentially the same as that of the Precambrian slates (0.220). If the Mississippi
mud, which is presumably a silty mud, were to lose 35 percent SiO, (quartz,
feldspar, smectite -, mica + chlorite), 2 percent H,O, and 1 percent CaO (calcite),
the product (Table 7-7) would have a composition quite similar to the Precambrian
slates, particularly for Si, Al, Mg, and K. The high Na value is probably due to the
presence of adsorbed Na and Na deposited when the pore water was evaporated.
The high Fe,O, value for the Precambrian slate is presumably related to the
environmental conditions that characterized that period, as i t is not that abundant
in younger slates.
To the clay mineralogist, removing 35 percent SiO, is a major project. However,
this is routine for the metamorphic geologist.
512

Table 7-7
Comparison of Precambrian Slate and Mississippi Delta Mud After Removing 38 Percent of Material.
Miss. Delta I 38% Loss Prec. Slate '
SiO, 69.96 56.4 56.3
Al20, 10.52 17.0 17.2
Fe20, 3.47 5.6 9.5
MgO 1.41 2.3 2.5
CaO 2.17 1.9 1.o
Na,O 1.51 2.4(1.3) 1.2
K2 0 2.30 3.7 3.8
Other 8.7 13.9 8.5
' Clark (1924)
' 35% SO,, 2% H 2 0 , 1%CaO
' Nanz (1953)
0 1 2 3 4 5
"C
I 1 I I
Avo. O n p Sea
A E

-
Miss. Delta 1235)

-
'*
' 0
0 *
0
0
0
Olipocsne -
0
C L. Cretacewr 11131

D Phanerzoic
Avo. Sh.US.
t
*E PI. (31

G R. 1261

* M U. Ordovician 1115)

** H Avp. Slam
* I R e C. Slam

** J
b mh t .

K Archean 1
I
1 L .2 3
Hi* G r d . Mat..,
4 -
170)

K20/AP203

Fig. 7-62. Relation of K,0/A1203 ratio of physilites to measured or estimated maximum burial
lemperature. Stars = Conasauga; circles, Hower et al., 1976; squares, Weaver and Beck, 1971: A,
Turekian and Wedpohl. 1961; B, Clark, 1924; C. Campbell and Williams, 1965; D. Cameron and
Garrels. 1980; E. Clark. 1924; F. Weaver and Beck, 1971; G . Murray, 1954; H. Eckel. 1904: I , Nanz.
1953: J and K. Cameron and Carrels, 1980: L, Shaw, 1956; M, Scotford, 1965. From Weaver and
Associates, 1984. Copyright 1984 Elsevier.
513

Another ratio that is of interest to physilite petrologists and geochemists is the


K20/A120, ratio of physilites. In general it is believed that K 2 0 (Weaver, 1967b)
and the K20/A120, ratio (Garrels and Mackenzie, 1971) increases with age. A plot
(Fig. 7-62) of the K20/A1,0, ratio of physilites vs. estimated maximum burial
temperture, similar to age, indicates there is no consistent change with age and,
further, there is no change with increasing metamorphic grade. This suggests that
both A1 and K are conserved during diagenesis - metamorphism, whereas, Si, and
possibly Ca and Na are lost.

ORGANIC AND PHYSIL PALEOTHERMOMETERS

For some time organic geochemists have measured various parameters of coal
and organic matter in shales in attempts to develop paleothermometers. More
recently efforts have been made to correlate various physil, temperature dependent
diagenetic changes (Kubler Index, Weaver Index, I/S ratio, illite and chlorite
polytypes, etc.) with the organic temperature scale. In 1960 Weaver (1960) published
a graph demonstrating a relation between oil reserves and the relative abundance of
expanded physils in the shales of each geologic period (Fig. 7-63). He proposed that
interlayer water released during the conversion of smectite to illite was a major
factor in hydrocarbon migration. Since that time there has been considerable
interest in physil-organics by oil companies. As the physils and organics are both
sensitive to changes in temperature and usually occur together, we should have some
understanding of organic diagenesis and metamorphism. The organics are similar to
the physils in that both time and temperature influence the rate of reactions. A table
listing the major thermal maturation indicators was compiled by HeCroux et al.
(1979).

Coalification Process

The book “Stach’s Textbook of Coal Petrology” (Stach et al., 1975) provides an
excellent review of coalification. Much of the first part of this discussion is taken
from their review.
The development from peat through lignite, subbituminous coals, and bi-
tuminous coals to anthracites and meta-anthracites is termed coalification. The
chemical and physical changes that take place during coalification indicate that only
the processes that operate up through the stage of lignite (soft brown coal) can be
classed as diagenetic. The alterations that occur from the beginning of the subbi-
tuminous (hard brown coal) stage are so severe that they are regarded as metamor-
phism.
Both chemical and physico-structural changes occur during coalification. De-
creases in porosity, moisture content, and optical anisotropy parallel to the bedding
planes occur during the early stages of diagenesis (peats and brown coals). More
514

O I L PRODUCTION
AND R E S E R V E S RELATIVE SMECTITE
PERIOD PERCENT” CONTENT

PLIOCede

MIOCENE
OLIGOCENE

EOCENE
U.CRETACEOUS

L.CRETACEOUS

JURASSIC

TRIASSIC -- I
PERMIAN
PENNSYLVANIAN

MISSISSIPPIAN

DEVONIAN

SILlidIAN

OKIJdVICIAN 7.1
CAMBRIAN

“ A d a p t e d f r o m H o p k i n s , Jour. P e t r o l . T e c h . , June 1 9 5 0 , P . 7
Fig. 7-63. Distribution of oil and expanded physils through geologic time. The relative amount of
expandable clay in the Upper Cretaceous is probably higher than indicated. From Weaver, 1960,
reprinted by permission of American Association of Petroleum Geologists.

advanced stages of diagenesis and metamorphism cause changes predominantly in


chemical properties and optical properties (dependent on chemical composition).
As the moisture content decreases, the calorific value of the coal increases. The
loss of moisture depends on a decrease in porosity and on the decomposition of
hydrophylic functional groups, particularly OH-groups. Also, carboxyl, methoxyl,
and carbonyl groups are split off, increasing the carbon content. During the lignite
to subbituminous stage the last remnants of lignin and cellulose are transformed
into humic materials. The humic acids condense to larger molecules, losing their
acid character to form alkali-insoluble “humins”. The most striking changes at this
stage are petrographic. Geochemical gelification (vitrinitization) of the humic sub-
stances occur and the coal becomes black and lustrous.
During the bituminous stage the volatile matter (V.M.) systematically decreases
and reflectivity rises. About equal amounts of CH, and COz are released. The
anthracite stage is characterized by a rapid fall of hydrogen content and of atomic
H/C ratio, and a strong increase in reflectivity and optical anisotropy. Large
quantities of methane are released; increasing graphitization occurs, especially
during the meta-anthracite stage. Some of these changes are summarized in Fig.
7-64.
51 5

Rank :al. Value Applicability of Different


EtulIb Rank Parameters
German USA % Vitrite kcallkg)
- -
-02
- 68
Torf Peat

- 64
w - c a 60 -ca It

- -03 - 60
Weich-
J= Lignite
- 0 - 56 c a lr
-1L

Matt- - 52
a Sub- - c-- 0 4
Bit. B - c a 11 ca 2 !
m - 48 t l
Glanz- -
m
-05

-06 - 44
- c a 11 .ca 8
Flamm- v)

as - 6 1
8 - 40
-
5 -08
m
iasflamm- -
0 - 36
PI

-
A z -
D -l[l
I

- - 32
Gas- =
O
Medium - I/
- 28 - c a 8;
r
Volatile -
c
3ituminous -I4 - 24
Fett-

--
-
.-

- Low
-I6
- 20
M
Volatile -
ESS- Bituminous - I * - 16
-
-20
Mager- Semi- - 12
Anthracite -
- 8 - c a 91
Anthrazit -
Anthracite -
- 4
-40
Meta-Anthr.
Meta-A. -
Fig. 7-64. The different stages of coalification and their distinction on the basis of different physical and
chemical rank parameters. D.A.F. = Dry Ash Free. From Stach er d..1975. Copyright 1975 Gebruder
Borntraeger.
51 6

FRANCE ;ERMANY CANAOA USTRALIA

UPPfA C m R
PARIS WC Sl
RUlW GIPPSCA~
BASIN GRAOLN CANAM NJTH P E R T )

CORRCIA iicuuu~~fr STIPLIN DCYAISOH


C I U 'HIS PAPER:

l¶6¶ nto nti mn

60" 60" 55" 65" 60"


' 80" 80" 75" 100"
90" ' 115" 115" " 90" I15"
c0 100" 130"
135" 140"

150" C'120" 170" ' 155O

165"
210" 200"

Fig. 7-65. Relation of coal rank to burial temperature in various areas. D.A.F. = Dry Ash Free. For
references, see original paper. From Demaison (1974). Copyright 1974 Comptes rendus de I'Acadimie
des sciences de Paris.

The coalification process is controlled primarily by temperature and time.


Pressure retards chemical reaction during coalification by inhibiting the release of
gas. Chemical rank normally increases with depth. The rate of rank increase is
dependent on the geothermal gradient and on the heat conductivity of the rocks.
Studies in folded regions indicate that pressure has little effect on coalification. In
most instances the coalification rank was obtained during burial and before folding.
Adjacent to faults along which movement has been rapid, the coal rank (and physil
rank) may be increased due to the creation of frictional heat. Generally, movement
is so slow that frictional heat is lost before it affects the coal.
Temperatures on the order of 100"-150"C will normally produce a bituminous
coal and temperatures on the order of 300°C (200°C is a more common value) will
produce anthracite. Numerous examples indicate that for samples buried to similar
depths those that have been buried for a longer time will have a more advanced
degree of coalification. The influence of time is greater the higher the temperature.
Demaison (1974) has compiled data showing the relation of maximum paleo-
temperature to coal rank (Fig. 7-65). Presumably much of the variation is due to the
influence of length of exposure to heat. Demaison believes that the influence of time
is minor past about 200 m.y. of exposure to maximum heat effects.
517

Calculations have been made showing how coal rank is related to temperature
and time (Huck and Karweil, 1955; Karweil, 1956; Lopatin, 1971; Lopatin and
Bostick, 1973). Lopatin (1971) notes that it is necessary to sum the time a coal was
in the various thermal zones in order to properly evaluate the time-temperature
effects (sum of the adjusted durations of the process).
Lopatin and Bostick (1973) reviewed some examples. Temperatures below 50°C
have little effect regardless of time. Lower Carboniferous coals of the Moscow Basin
have not been subjected to temperatures higher than 25°C and remain lignites. In
contrast, a temperature of 100°C has produced anthracite in the Upper Carbonifer-
ous, Oklahoma (270 m.y.); about 175°C was required to convert coaly material to
anthracite in Jurassic (160 m.y.); and in the Salton Geothermal Field (Late
Pliocene-Early Pleistocene), temperatures of 290" to 300°C were required to pro-
duce anthracite coal. Diagenetic-metamorphic alterations of physils are much less
time dependent.

Shales
Finely divided coaly particles are present in most shales and siltstones. Bostick
(1974) coined the term phytoclasts for detrital plant material in sediments which has
a composition similar to coal (others use the term kerogen). He concluded that 90 to
95 percent of the organic matter in sediments is phytoclasts rather than compounds
soluble in organic solvents. Most of the phytoclasts consist of plant spores and
pollen, leaf cuticles, and fragments that have bordered pits, ribs, fibers, or cellular
structure. Some consist of insoluble humate and flocculant masses that may be of
algal origin.
Rank measurements can be made with the microscope on these organic particles.
Untreated rock samples are more satisfactory than concentrates prepared by chem-
ical or physical methods. However, separation is often necessary because many
rocks have an extremely low content of organic material.
One of the simplest methods of determining the degree of thermal alteration is
through the use of a color index ( S t a p h , 1969; Hood et al., 1975; Epstein et al.,
1976). The color of kerogen generally changes from yellow ( < 60°C) to brownish
yellow to various shades of brown (60" to 125°C) and finally to black ( > 125°C) as
the degree of alteration increases. More quantitative measurements include vitrinite
reflectance (Bostick, 1974), carbon content, and electron spin resonance.
CastaEo and Sparks (1974) found little difference in vitrinite reflectance of
coal-shale pairs except when the shales contained small amounts of vitrinite.
Palynomorphs in limestones have significantly lower reflectance values than associ-
ated coals (Epstein et al., 1976).

Organic Material
Coalification and bituminization are diagenetic processes during which mobile
products (gas or crude oil) are evolved and aromatization and condensation of the
518

solid residual products (coal or kerogen of petroleum source rocks) take place. In
general, the parent substance and the initial decomposition soon after deposition
will determine whether coal or petroleum forms. Coals are formed primarily from
lignin and cellulose from higher plants, and petroleum from lower organisms (algae,
animal plankton, and bacteria). These latter organisms decompose anaerobically to
produce lipids, the most important progenitors of hydrocarbons. During the geo-
chemical bituminization process, the lipids become incorporated into kerogen.
Thermal degradation of organic matter is required to produce petroleum (Philippi.
1965). Degradation forms smaller molecules of increasing volatility, mobility, and
hydrogen content (with methane as the end product). The carbonaceous residue
(condensation) loses hydrogen, with graphitic carbon being the end product
(Dobryansky, 1963).
The results of studies relating hydrocarbon occurrence to rank of associated coaly
matter in various oil provinces were tabulated by Bostick and Danberger (1971). Oil
and associated gas occurs mainly in rocks in which the coaly material has reached
subbituminous to high-volatile bituminous rank. Dry gas (methane) is formed
during the medium-volatile bituminous coal through the anthracite stage. The base
of the oil zone also coincides with 65 to 70 percent fixed carbon and vitrinite
reflectance values of 0.8-1.1.
The effects of time on hydrocarbon generation were demonstrated by Laplante
(1972). In the Gulf Coast Tertiary he found the level of carbonization necessary for
hydrocarbon generation increased from 80°C for Oligocene sediments to 97°C for
Upper Miocene. Pusey (1973a) introduced the concept of a liquid window. He noted
that thermal “cracking” of kerogen and hydrocarbon generation starts at 65°C. and
that liquid hydrocarbon destruction becomes dominant at temperatures greater than
150°C. This temperature interval he termed liquid window. It is over this same
temperature range that most smectite layers are converted to illite.
The various coal-ranking properties (calorific value, moisture content. volatile
content, hydrogen content, and vitrinite reflectance) are not applicable over a wide
temperature range. Hood et a/. (1975) suggested a scale which is continuous over the
temperature range of approximately 50°C to 300°C and which can be correlated
with coal-rank parameters. Though temperature is the major defining factor in the
process of organic metamorphism, the effect of time is important enough that a
temperature scale cannot be used as a general measure of organic metamorphism.
They call their scale “Level of Organic Metamorphism” or “LOM” (Fig. 7-66). The
scale is a linear subdivision of a type-section (Tertiary-Cretaceous of New Zealand)
of organic-rich sediments which was buried at essentially a constant rate and a
constant temperature. Their scale is arbitrary and they note that the coal rank of a
Carboniferous relative-depth column compiled by Bostick and Danberger (1971 )
does not coincide with the one they chose to use.
Lopatin (1971) and others concluded that doubling the reaction rate with each
increase in temperature of 10°C provided a model of the relative effects of
temperature and time in subsurface organic metamorphism. Hood et al. (1975)
combined maximum temperature (T,,;,,) with an effective heating time ( t e l l ) to
develop a simplified method of predicting LOM for petroleum source rocks. Using a
519

HYDROCARBON
OM
+O RANK GENERATION

LIGNITE
DIAGENETIC
METHANE

SUB-
6
BIT.

OIL

35 .
30 '
-

6I
CONDENSATE
6
15 .

-
WET GAS

CATAGENETIC
METHANE

ANTH. l 5o --

!O

Fig. 7-66. Level of organic metamorphism (LOM) as defined by Hood et al., 1975. Modified by Waples,
1984. Reprinted by permission of American Association of Petroleum Geologist and Academic Press,
Inc.

variety of burial histories and a wide range of activation energies (8.4 to 55


kcal/mole), they concluded that the time during which a rock has been within 15°C
of its maximum temperature represents a suitable definition of effective heating
time.
The most commonly measured parameter of particulate organic material in shales
is vitrinite reflectance ( R o , It,,,). Vitrinite reflectance is a physical property of
vitrinite that is related to its chemical structure. The generation of hydrocarbons
from vitrinite is apparently directly proportional to the amount of vitrinite in a
source rock (Tissot and Welte, 1984). Most modeling methods for vitrinite reflec-
tance assume it is directly proportional to a time-temperature integral (Waples,
1980; Middleton, 1982) or absolute temperature (Baker, 1983). Neither approach
completely explains the vitrinite values observed in a basin (Antia, 1986).
520

Vitrinite’s properties are affected by the initial composition of the vitrinite,


microbial activity, pore fluid composition, thermal history, fluid movement, sedi-
ment mineralogy, and pressure history. (This is a long list of factors or variables and
unfortunately the same factors affect the physils.) Further, the value of vitrinite
reflectance is dependent on analytical procedure, handling, storage; its identifica-
tion is subjective; the reflectance of individual grains is not uniform, etc. (Antia.
1986). I am not certain who is working with the more difficult material, the organic
geochemist or the clay mineralogist. Organic material is perhaps more responsive to
temperature than the physils, but it is also more sensitive to a number of interfering
variables. There is no obvious reason to expect that organic material will provide
more accurate paleotemperature data than the physils.

Physil - Organic Paleothermometry

Kisch (1968, 1974, 1987) reviewed the relation between coal rank and burial
metamorphic mineral facies. He concluded that the diagenesis-anchizone boundary
coincides with the change from lean coal to anthracite (V.M. 8 percent, V.R. 2.5).
This change in coal rank occurs at a temperature of approximately 200°C (Demai-
son, 1974; Hood et a/., 1975). The boundaries of the anchizone are based on K.I. for
illite, upper 7.5-lower 4.0 (0.42”-0.25” 2 8 ) and approximately coincides with the
boundaries for anthracite (8 to 4 percent volatile matter and 2.7 to -
4.2 R l , , ) .
Kubler et al. (1979) suggested V.R. boundary values of 2.6-2.8% and 4.0-4.2%.
Later, Kisch (1980) reported on numerous areas where the onset of anchizone illite
(K.I.) was associated with high-rank anthracitic coal. Terchmuiiller el a/. (1979)
found that the diagenesis/anchizone boundary corresponded with the
anthracite/meta-anthracite boundary (German coal classification, 3.5 percent R ).
This would suggest a diagenesis/ anchizone boundary temperature closer to that
reported for the Conasauga Shale (Fig. 7-49).
Kaolinite and mixed-layer illite-montmorillonite are presumed not to be present
in anchizone sediments (Kisch, 1974); however, the compiled data indicate that
kaolinite is present in the underclays associated with the anthracite coals of the
northeastern United States, along with pyrophyllite, illite, and chlorite. Frey and
Niggli (1971) considered that the upper anchizone boundary (7.5 K.I.) coincided
with coals containing 35 to 19 percent V.M. (compared with Kisch’s boundary at 8
percent V.M.). The lower boundary is similar to that proposed by Kisch. Quinn and
Glass (1958) found anthracite and low-rank meta-anthracite of the Narragansett
Basin, Rhode Island, associated with rocks of the muscovite-chlorite subfacies of the
greenschist facies, and a higher-rank meta-anthracite with rocks of the biotite-chlo-
rite subfacies.
For West Germany coals there is no simple relation between illite “crystallinity”
and coal rank; each region is characterized by its own relationships (Wolf, 1975).
Teichmuller et al. (1979), in a study of 420 samples, found that “crystallinity” and
coal rank could not be correlated in all cases. In the Central Alps, a good
correlation between K.I., coal rank, and fluid inclusion data in some areas, but no
521

S.R.
7
1
92.5
I
3
I
85
FIXED CARBON%
I
80
I
70
1
65 I

1’ w A P P R O X I M A T E L Y 70 KM E
1
Fig. 7-67. The relationship between illite sharpness ratio in the Coal Measures (South Wales) terrigenous
rocks and the fixed carbon percentage in the associated coal. From Gill ef a / . , 1977. Copyright 1977
Elsevier.

general relation, was found (Frey er dl., 1980). In part this is due to variation in
lithology which influences the K.I. values.
Gill et al. (1977) found a good relation between the coal rank, fixed carbon, and
the sharpness ratio (S.R.) of the illite-I/S (Fig. 7-67). Anthracites have > 90 percent
fixed carbon. The percent expandable layers in the I/S ranges from 40 to 0 and
decreases with increasing S.R. The carbonate rocks have slightly lower S.R. values
than the shales.
Van de Kamp (1977) attempted to relate LOM values to physil changes. He
states that the montmorillonite zone occurs in the interval LOM 0 to 7; mixed-layer
illite-montmorillonite, LOM 7 to 9-10; illite zone, LOM 9-10 to 14; greenschist
facies, LOM greater than 14. These associations are incompatible with other data. A
more realistic relation would be: montmorillonite, LOM 0 to 9-10; mixed-layer
illite-montmorillonite, LOM 9-10 to 18-20; illite, LOM 16 + . An LOM value of 20
is considered to be equivalent to 2 percent V.M. (meta-anthracite) and the greenschist
facies; however, the physil data do not support this. Hood er al. (1975) report that
the organic matter in the Carboniferous shales from the Shell Rumberger 5
(Oklahoma) should have an LOM value of 20 at approximately 7500 m, which
should be in the greenschist facies. My x-ray analysis of a shale core from 7318 m
(Total Depth) from the same well showed that it is composed of illite and chlorite
(dioctahedral) and 25 percent I/S. The W.I. is 2.9 and the K.I. is 10, indicating
diagenesis rather than low-grade metamorphism.
In any given area, coal rank may correlate well with physil suites, particularly
K.I. and W.I. On an absolute basis there is not a close correspondence between coal
rank and illite-I/S grade. This is because changes in both materials are influenced
522

Table 7-8
Correlation of K.I. and Coal Rank Data from Various Areas (Frey ef al.. 1980).
~

Beginning Anchizone End of Anchizone


0.42’ 29 % R,, 0.25’ 2 8
- 2.25 - 4.0
2.65 5.0
2.6-2.8 4.0
3.4 5.5
3.5

by a number of factors other than temperature, probably coals more so than


illite-I/S. In most studies it is assumed the illite “crystallinity” values represent the
true grade of metamorphism (presumably temperature) and the mean reflectivity
(R,,,) measures variations in coal rank within the metamorphic grade. Table 7-8
contains a list of R , values from various areas for the beginning and end of the
anchizone (based on K.I.). Fig. 7-68 shows two examples from the Central Alps. In
one area there is a good linear relation between illite crystallinity (K.I.) and mean
reflectivity (%R,,,)and in the other there is none (Frey et al., 1981).
Fig. 7-69 shows the relation between vitrinite reflectance and sharpness ratio
(W.I.) and crystallinity index (K.I.) for Carboniferous shale samples from the
Ouachita Mountains, Oklahoma and Arkansas (Guthrie et al., 1986). The W.I.
values indicate most samples are from the anchizone. The correlations are relatively
good and the authors suggest they may be used to identify the zone of oil generation
and preservation. However, this is likely to be true only if the illite peak values are
first calibrated with an organic parameter.
Albrecht and Ourisson (1969) studied a well penetrating the Cretaceous Souala
shales of the Cameroun and established that extra stable heavy saturated hydro-
carbons were most abundant in the depth range of 1300 to 2200 m. Present-day well

FALKNIS W P E
Falhisbreccien
Neokanflvsch
OTristeischichten

4
s.

4 6 8 10 I2 14
ILLm CRYSTALLINIn INDM
Fig. 7-68. Correlation of illite crystallinity versus the mean reflectivity (R,) of two studied areas from
the Central Alps. From Frey et a]., 1980. Copyright 1984 Eclogde Geol. Helv., Birkhauser Verlag AG.
523

= 0.28 + 0.08X
- LOG Y
N = 54 r = 0.640 A
ATOKA
J A CKFORK
STANLEY

0
t
<
u
10 1
-
-- 0 .
..
y;:w 5 -
-
z
a
u -
<
I
m

1
-
.
1 1 : 1 I I I ( I I I

0 1 2 3 4 5

lLo;Y;l.o;-o.;, , 0
; , A T O,K A ,
N = 54 r = 0.566 A JACKFORK
a
0 STANLEY
1
0 1 2 3 4 5
MEAN VITRINITE REFLECTANCE (% in oil)
Fig. 7-69. Relation between mean vitrinite reflectance and illite sharpness ratio (W.I.) and crystallinity
index (K.1) for Carboniferous shales from the Ouachita Mountains, Oklahoma and Arkansas. Metamor-
phic grad range from high diagenesis to near beginning of epimetamorphism. From Guthrie et al., 1986,
reprinted by permission of American Association of Petroleum Geologists.

temperatures over this interval increase from 60’ to 105’C. The x-ray data of
Dunoyer de Segonzac (1970) indicate that through this interval the proportion of
illite layers in the montmorillonite increases from approximately 10 percent to 70
percent.
Pusey (1973b) established a correlation between ESR-kerogen data and the
“ first-collapse” of montmorillonite in the Cretaceous of South Texas. Oil is pro-

duced above the “first collapse” of montmorillonite (120°C) but gives way to dry
gas below the “first collapse”.
524

Foscolos and Kodama (1974) and Foscolos et ul. (1976) studied the physils and
organic material in a series of Lower Cretaceous shales form northeastern British
Columbia and established diagenetic relations. Vitrinite reflectance values show
considerable variation in the shales where diagenetic reactions have been minor, but
increase systematically as mineral diagenesis increases (K.I., W.I., I/S ratio, etc.).
The interval of maximum yield of extract and hydrocarbons (mg/g organic C )
coincides with the interval where the proportion of illite in the I/S increases from
25 percent to 75 percent. The relation, particularly the upper boundary, is not good.
In part, this is because the assumption is made that the starting expanded physil was
a pure smectite.
In the North Sea area illitization of smectite starts at a vitrinite reflectance level
of 0.4 percent R,,; by R , , 0.6 percent the I/S ratio is 4:l (Pearson et ul., 1982). In
the Cretaceous shales of the Powder River Basin, Wyoming (Burtner and Warner.
1984), there is a close correlation between the beginning of I/S diagenesis and the
onset of oil generation as indicated by the Rock-Eva1 pyrolysis 7;,,;,, values and the
Rock-Eva1 production index, whatever they are. In the Cretaceous shales o f the
Montana disturbed belt, analyses of organic matter in the shales, including pyrolitic
hydrocarbon, temperature of maximum pyrolysis. hydrogen and oxygen indexes,
vitrinite reflectance, and thermal alteration index, all show a poor correlation with
the proportion of illite layers in 1/S (Schultz, 1984).
There is a lot that remains to be done in relating organic and physil reactions
during diagenesis and metamorphism, but the results should be well worth the
effort. The “model” will involve synthesizing the weighted interactions of a large
number of variables. Eventually the relations will prove to be relatively simple and
the interpretive potential staggering (second definition).
525

Chapter Vlll

PHYSILS IN SANDSTONES

Introduction

When I first started trying to use physils, then known as clay minerals, to solve
stratigraphic problems, I soon learned to stay away from sandstones. Almost
invariably sandstones have a different physil suite than the surrounding shales. It is
now well established that many, perhaps most, of the physils in sandstones were
formed during burial by fluids migrating through the permeable sandstones. In a
study of 785 samples, Wilson and Pittman (1977) found authigenic physils were
present in 90% of the samples. In deeply buried sandstones the physils are altered
by metamorphic processes, and tend to become similar to those in the adjacent
physilites.
Analyses of adjacent sand and mud samples in recent sediments from various
environments indicate the physil suites are similar. Water currents are not particu-
larly effective in segregating physils by size. In part t h s is because the physils
commonly occur as mixed-mineral aggregates and pellets; even if the physils are
dispersed they commonly are deposited by percolation, on and in, sands during
conditions of slack water after the sand is deposited.
Size of the sandstone units is not a factor. Analyses of 1 to 5 mm thick
Cretaceous sandstone laminae with various configurations - regular, irregular,
mottled, and slumped - showed that they contained 2 to 7 times more kaolinite
than the adjacent shales (Weaver, 1967). Further, in at least some instances the
amount of secondary physils in sandstones is proportional to the porosity. Fig. 8-1
shows the close relation between an electric log pattern of a Cretaceous beach
sandstone, Washakie Basin, Montana, and the kaolinite content of the physil suite.
The proportion of kaolinite increases with increasing porosity and permeability. In
the Washakie Basin there is a fairly good relation between the kaolin content of the
physil suite and the initial oil production rate. The kaolinite is secondary and the
relation to porosity is apparently due to the destruction of feldspar and micaceous
rock fragments.
In addition to kaolinite, chlorite, illite (I/S), smectite, and Ch/S are common
authigenic physils in sandstones. I (Weaver, 1959) originally tried to determine if the
various sandstones, quartzites, arkoses, and graywackes had distinctive physil suites.
In view of the diagenetic origin of so many of the physils in sandstones it is not
surprising that there is no consistent relation between physil suites and sandstone
type.
526

FOREST ARCH. NO. 3


I X KAOLINITE
0 50 100

Fig. 8-1. Resistivity (right) and self potential log of a Cretaceous beach sand, Washakie Basin, Wyoming.
Markers are at 10 foot intervals. High resistivity zone at the top of the sand is due to calcite cement.
Right graph shows the percentage of kaolinite in the physil fraction.

Petrologists have long observed authigenic and diagenetic physils filling pores
and replacing minerals in sandstones. Since the development of the scanning
electron microscope (SEM) and the recognition by oil companies that physils
strongly influence productivity and secondary recovery there has been a major
(explosive) increase in the study of physils in sandstones.
The extensive development of secondary physils has complicated the job of the
sandstone petrologist. Feldspars can be completely replaced by physils and turn an
arkose into a quartzite, after burial. The development of secondary physils increases
the amount of matrix material and decreases the average grain size, thus complicat-
ing the interpretation of the textural data (Wilson and Tillman, 1974).
Wilson and Pittman (1977) were among the first, along with Fuchtbauer and
Muller (1970), to stress the importance of authigenic physils in sandstones. In part,
this awareness was a result of the development of the SEM. The term authigenic is
widely used to refer to both authigenic and diagenetic physils. It is used to describe
physils which grew in the place where they occur.

Recognition of Detrital and Authigenic Physils


The physils in sandstones are either detrital (allogenic) or authigenic-diagenetic.
Table 8-1 and Figures 8-2 and 8-3 show the modes of occurrence of allogenic and
secondary physils in sandstones. Granular detrital physils can be contemporaneous
mud fragments and clasts or shale, slate, and schist rock fragments. The size of the
physil flakes and the degree of orientation give a clue to the rock type. In general,
flake size and orientation increase with increasing diagenesis and metamorphsm.
527

Table 8-1
Reliability of Criteria for Authigenic Origin of Physils, after Wilson and Pittman (1977).
Criteria Relia- Frequency
bility of
Occurrence
Composition
1) Absence of impurities 0 C
2) Clay assemblage monomineralic X R
3) Significantly different from associated shales + U
4) Concentric color zoning 0 R
Morphology
5) Crystal outlines + C
6) Delicate projections + C
7) Undeformed 0 C
8) Pseudomorphous replacement + R
Structure
9) Low-temperature polytypes + ?
10) High degree of crystallinity 0 U
Texture
11) Gap in grain-size distribution (silt fraction lacking) X C
12) Large particle size 0 C
Distribution
13) Pore linings missing at grain contacts + C
14) Scattered pore fillings X C
15) Fracture fillings + R
16) Absent in early diagenetic concretions + C
17) Cover diagenetic components formed at earlier stage + C
18) Laminae with abrupt lateral terminations + R
19) Radial alignment of individual flakes 0 C
20) Medial sutures 0 R
21) Bridges between detrital grains near points of contact + C
X = unreliable; 0 = generally reliable; + = very reliable; R = rare; C = common; U = ubiquitous; ? =
unknown.

The physil fragments, because of their lower specific gravity, are usually coarser
than the associated quartz grains. Plate-shaped clasts can be significantly larger than
the quartz grains. Biogenic pellets, sometimes converted to glaucony (Fig. 8-3), are
another common granular form. Studies of recent sediments suggest pellets may be
more abundant than commonly realized.
Dispersed physil flakes may filter into pores during slack water periods in any
type of water environment or form a thin coating on detrital grains (cutans) in
continental soil-like environments. Dispersed or flocculated physils are deposited as
thin laminae during periodic low energy conditions in both continental and coastal
marine environments. These laminae may be preserved, but in many environments
burrowing organisms commonly homogenize the laminated sediments.
Authigenic physils crystallize from solution (neoformation), lining pores, filling
pores, fractures, cleavage faults and vacuoles, and by alteration of detrital physils
528

DISPERSED
MATRIX

INTERCALATED
LAMINA

DETRITAL
MICA
RIP-UP
CLASTS

INFILTRATION
RESIDUES

Fig. 8-2. Mode of Occurrence of allogenic clay in sandstone. From Wilson and Pittman, 1977. Copyright
1977 Soc.Econ. Paleo. Miner.

(transformation). Apparent replacement of minerals by physils occurs in at least two


ways. One process involves dissolution of a mineral and crystallization of physils in
the space created by the dissolution, e.g., kaolinite “replacing” calcite. Another
form of replacement is a result of the alteration (hydration) of the surface of a

PORE LINING PORE FILLING

CLEAVAGE
TRACES

PSEUDOMORPHOUS REPLACEMENT FRACTURTFILLING


Fig. 8-3. Modes of Occurrence of authigenic clay in sandstones. The sizes of the individual flakes or
aggregates have been exaggerated for purposes of illustration. From Wilson and Pittman, 1977.
Copyright 1977 SOC. Econ. Paleo. Miner.
529

silicate mineral; an amorphous phase is usually formed(?) which reorganizes, with or


without the addition of external cations, to produce a physil, e.g., kaolinite and
illite, replacing feldspar. These processes are discussed in Chapter 111.
Wilson and Pittman (1977) compiled a list of criteria (Table 8-1) which can be
used to distinguish between neoformed and detrital physils. The criteria are grouped
into five categories. The criteria have varying degrees of reliability. Some of the
diagnostic features can be observed in thin section; others require SEM or x-ray
diffraction analysis.
Composition: Authigenic physils are “glass-clear”, whereas detrital physils gener-
ally have a darker, impure appearance due to the presence of Fe and Ti oxides and
organic material. The “impurities” are usually deposited with the physils in the
original depositional environment, though some of the oxides may be produced by
weathering in the source area. As mentioned previously a major difference between
the physil suite of the sandstone and adjacent physilite usually indicates the former
is authigenic. The contrast is less in volcanogenic sediments where the physils in
both the sandstones and physilites may have formed authigenically.
Morphology: Morphology, as seen with the SEM, is perhaps the most distinctive
feature of authigenic physils. Some of these physils occur as interlaced mats of
fibers with branching patterns and delicate tips extending into pores. The more
flake- or plate-shaped physils commonly occur in the form of delicate rosettes, and
vermicular or honeycomb aggregates. The pseudomorphic replacement of detrital
grains, commonly feldspar and volcanic material, is a reliable indication of authi-
genic physils. If the replacement occurred in the source area the product would
normally break LIP during transport.
Structure: Some physils have low-temperature and high-temperature polymorphs.
Type Ibd and Ib chlorite and 1Md and 1M illites are the low-temperature authigenic
polytypes; however, these polytypes can also be detrital. Most shales contain some
1Md illite (or I/S). If a sandstone contains a low temperature polytype and the
adjacent shales a high temperature polytype, the low temperature variety is most
likely authigenic. Also, during deep burial ahd high temperatures ( - 150” to 25OOC)
the low temperature varieties are transformed to the high temperafure varieties (IIb
chlorite and 2M mica). The degree of crystallinity as indicated by the sharpness of
the x-ray peaks is generally good for authigenic physils but not distinctive.
Texture: Sandstones containing authigenic physils may have a bimodal size
distribution. The authigenic physil flakes and fibers are generally larger than the
majority of detrital physils.
Distribution: Where physils line pores but are absent in areas where there is
grain-to-grain contact, it is likely the physils are authigenic. Physil growth starts at
the grain surface and extends outward, in some instances completely filling the
pores. In other situations a second type of physil or another mineral, usually quartz
or calcite, or both, may grow pore-ward of the clay coating.
Physils filling fractures, coating authigenic minerals or aligned radially with
respect to the surfaces of detrital grains are usually authigenic. With the SEM
authigenic physils can often be observed forming bridges between detrital grains.
Table 8-2
Characteristics of Authigenic Clays, from Wilson and Pittman (1977). w
1/1
0
Morphology Form of Relationship Thickness of Special
of Flakes Aggregates to Sand Size Coating or Features
Detrital Long Dimension
Grains of Aggregates
(microns)
Kaolinite and pseudohexagonal stacked plates pore filling 2-2500 flakes notched or embayed
Dickite (book) (generally (twinned?)
2-200)
pseudohexagonal vermicule pore filling 10-2500 flakes notched or embayed
(generally (twinned?)
20-200)
pseudohexagonal sheet pore filling 0.1-1 flakes notched or embayed
(twinned?)
Chlorite pseudohexagonal plates (two- pore lining 2-10
dimensional
cardhouse)
curled equi- honeycomb pore lining 2-10
dimensional with
rounded edges
equidimensional rosette or fan pore lining 4-150
with angular or and pore (generally
lobate edges filling 4-20)
fan-shaped cabbagehead pore lining 8-40
fibrous bundles and pore
filling
Illite irregular with sheet pore lining 0.1-10 bridging between
elongate spines sand grains
Smectite not recognizable wrinkled sheet pore lining 2-12 bridging between
or honeycomb sand grains
Mixed-layer subequant with imbricate sheet pore lining 2-12 bridging between
smectite/illi te stubby spines to ragged sand grains
honeycomb
531

Characteristics of Authigenic Physils (modified from Wilson and Pittman, 1977)


Kaolinite and Dickite: Kaolinite and dickite are the most common recognized
physils in sandstones. They most commonly occur as pore fillings. The plates are
usually well developed with a hexagonal or near-hexagonal shape. However, in thin
section it is sometimes difficult to distinguish between chert and packets of
kaolinite. They commonly have a higher degree of crystalline order (x-ray) than
associated detrital kaolinite. Dickite forms at higher temperatures than kaolinite,
sometimes from kaolinite. The characteristics of the major types of authigenic
physils are summarized in Table 8-2.
In most cases kaolinite fills only scattered pores in sandstones. It occurs as
packets or small stacks of pseudohexagonal kaolinite plates randomly arranged with
respect to each other (Fig. 8-4). Individual plates range from 3 to 20 p m in
diameter. In some instances large vermicular stacks or books can be observed.
Chlorite: Chlorite occurs in a variety of forms. It occurs as individual plates
coating sand grains, as rosettes, honeycombs, or cabbagehead-like growths (Fig.
8-5). In the plate type, individual flakes are oriented perpendicular to the surface of
sand grains. Individual flakes range from 2 to 10 pm in diameter. The flakes are
arranged primarily in a face-to-edge orientation. In the honeycomb type, the
face-to-edge arrangement is more fully developed and flakes are commonly curved
to produce a cellular pattern.
Rosettes and fan-shaped clusters occur both as pore linings and as pore fillings.
Flakes are oriented both face-to-edge and face-to-face (Fig. 8-5). Individual rosettes
or fans generally are 5 to 20 p m in diameter. The rare cabbagehead variety consists

Fig. 8-4. (A) Stacked plates of kaolinite in porous sandstone. Note the face-to-face arrangement and
pseudohexagonal outlines of individual plates, Eocene Frio Sand, Texas. (B) Vermicular authigenic
kaolinite in porous sandstone. The kaolinite also occurs. as individual isolated plates. Sand grains are
coated with illite which apparently formed prior to the kaolinite. Permian Rotliegendes Sandstone, North
Sea. From Wilson and Pittman, 1977. Copyright 1977 Soc.Econ. Paleo. Miner.
532
533

of relatively tightly packed, subparallel oriented flakes that form subequant grains.
Individual grains range from 8 to 40 pm. Lath-shaped, 7 A, Fe-rich chlorite
(berthierine?) coating sand grains has been observed (Dean, 1983).
EDX analyses indicate the Fe concentration increases in the following sequence:
plates and rosettes > honeycomb > cabbagehead. The authigenic chlorites in sand-
stones have no distinctive composition (Hayes, 1970), but on average they are more
Al- and Fe-rich than igneous and metamorphic chlorites. In contrast, authigenic
chlorites in evaporite rocks are Mg-rich. X-ray analyses of a large number of
authigenic chlorites in sandstones (Hay, 1970) indicate the vast majority are the Ib
polytype. These are apparently transformed to the IIb polytype at temperatures of
150' to 200°C.
Illite: In thin section illite appears as high birefringent packets, filling pores or
replacing feldspar and as coatings on grains. Critical point drying (an elaborate
process that involves replacing the pore fluid with methanol and then replacing the
methanol with liquid CO, which is then allowed to evaporate) indicates illite occurs
as free-standing, tape-shaped fibers of variable length (2-70 pm), width (1-2 pm),
and thickness (0.0002-0.0001 pm). The fibers line pore walls or completely bridge
pores (Fig. 8-6). When samples are air-dried (normal procedure) the fibers become
interwoven or matted, producing flake-like or honeycomb structures (McHardy et
al., 1982; Cocker, 1985). These hairy illites have an important influence on sand-
stone permeability. Rochewicz and Bakun (1980) (quoted by Srodoh and Eberl,
1984) suggested a progressive growth pattern whereby hairy illites aggregate or stack
to form sheets that develop as honeycomb-like pore filling. When illite replaces
kaolinite the general morphology of the kaolinite plates is preserved but the sharp
edges are lost (8-18).
X-ray patterns of authigenic illites in sandstones are commonly similar to those
for I/S physils that have 10 to 30% smectite layers (Fig. 2-20). Some of the I/S
material is ordered with superlattice peaks at 27 A (Hurst and Irwin, 1982). At least
some deeply buried samples have 10 A illites (Giiven et al., 1980). The laths have a
1M stacking sequence (Weaver, 1953; Giiven et al., 1980). There is some indication
that the x-ray patterns are deceiving. McHardy et al. (1982) observed with the TEM
that the authigenic illite laths averaged only three 2:l layers thick (Fig. 8-7). They

Fig. 8-5.(A) Chlorite coating sand grain. Individual plates are attached on edge to the detrital sand grain.
Pennsylvanian Siro Sand, Oklahoma. (B) Honeycomb growth form of chlorite as coating on sand grains.
Dark spots on grains represent points of contact between adjacent grains, some of whch were removed
during sample preparation. See C for an enlargement of chlorite. (C) Honeycomb chlorite that has
formed as a coating on a sand grain. Individual chlorite crystals curve and intersect to create the
honeycomb appearance. Jurassic Norphlet Sand, Florida. (D) Chlorite in the form of rosettes occurring
as a pore filling. Note the lobate edges which contrast to the pseudohexagonal outline of kaolinite that it
superficially resembles. Miocene Discorbis Sand, Louisiana. (E) Low magnification view showing chlorite
in the form of small cabbagehead structures attached to sand-sized detrital grains. See F for a magnified
view of cabbageheads. (F) Cabbagehead form of chlorite. Note the individual layers of clay particles that
comprise the structure. Eocene Frio Sand, Texas. From Wilson and Pittman, 1977. Copyright 1977 Soc.
Econ. Paleo. Miner.
534

Fig. 8-6. SEM photographs of hairy illites in sandstones. A-E = possible stages in development of
diagenetic illite in sandstone pores. Lath growth begins from the surface of detrital grains (A); the fibers
aggregate by regular (B) or irregular (C) stacking or by being bound with another, Fe-richer, physil (D);
irregular stacking leads finally to the development of honeycomb-like pore fillings (E). The critical point
drying technique was not used so the sequence may not be real. Photos by A. Rochewicz. From Srodoh
and Eberl, 1984. Copyright 1984 Miner. Soc.Amer. Courtesy D.D. Eberl.

suggested these thin units consisted of three 2:l layers bound together by two planes
of K ions. The basic units stack on top of one another when sedimented on slides
for x-ray analysis; swelling occurred between these unit packets (p. 59).
Some of the fibrous illites have a relatively low Mg and Fe content (Weaver,
1953; Rochewicz and Bakun, 1980), but others are similar to illites derived from
smectite (Srodo6 and Eberl, 1984). Illites replacing kaolinite in tonstein, and
presumably in sandstones, have a low Mg and Fe content and a high A1 content
535

Fig. 8-7. TEM of filamentous illite, three layers thick, from the sandstone from the North Sea x 120.000.
From McHardy et al.. 1982. Courtesy W.J. McHardy.

(Srodon and Eberl, 1984):

.96Fe0.06Mg0.0, (si3.1 3 A10.87 )O10 0.69Na 0.05

The illite contains only 2% expanded layers; a small admixture of chlorite is present.
Smectite: Authigenic smectites occur as a crinkly coating or a cellular mat on
sand grains. The smectite apparently occurs as a thin sheet of very small flakes that
is distorted when the sample is dehydrated (Fig. 8-14).
Mixed-Layer Illite/Smectite: The I/S physils have morphologies intermediate
between the two end members. A smectite-rich I/S may resemble crinkly smectite
but have spiny projections. As the illite content increases, the individual sheets with
lath-like edges develop (Fig. 7-11).
536

Factors Affecting Diagenesis


A number of factors influence the type of physil that will grow in a sandstone:
porewater chemistry, temperature, pressure, detrital mineralogy, organic content
and composition, and tectonics. Water composition, mineralogy, and temperature
would appear to be the most important factors. There are two main sources of pore
water in sandstones: (1) fresh meteoric water, which is dilute (commonly < 1000
mg/l) and acidic; (2) sea water, which is relatively concentrated (up to > 300,000
mg/l) and alkaline (including expressed mud water). Concentrated brines formed
by the evaporation of fresh water are of local significance in some evaporite deposits
in desert sands. The initial pore water chemistry is determined by the depositional
environment. Initially, fluviatile sandstones contain fresh water and marine sand-
stones contain sea water. Kaolinite commonly forms in fresh acid waters; illite,
smectite, and chlorite form in modified sea water or concentrated meteoric waters.
There is an initial association of physil type to sandstone type; however, pore water
chemistry is easily changed.
Tectonic activity can lead to a change in pore water chemistry. Thus, tilted
outcropping sandstones can be a conduit for meteoric water. Faults can allow deep
marine pore water to migrate upward into sandstones containing meteoric water.
Aside from the initial pore water in sandstones, the main source of new water is
interstitial water released from mud-shales as a consequence of compaction and, at
greater depth, water released during the conversion of smectite to illite. Both the
inorganic and organic composition of the fluids must be considered.
Most fresh water is acid enough and dilute enough that it falls in the kaolinite
stability field (Fig3-11). However, the acidity of both fresh and sea water is
commonly increased by the production of CO, from organic material. Both par-
ticulate and dissolved organic matter (aliphatic acids) are present in most sand-
stones (Kharaka et al., 1985). At temperatures < 80°C thermal and bacterial
decarboxylation of organic matter is the main source of CO,. The organic source
has been confirmed by carbon isotopic studies. At < 80°C the most important
inorganic weak acid is H,CO,, with H,S and H,SiO, being of secondary impor-
tance. At temperatures > 80°C weak organic acids are the dominant species, at
least in oil field waters, but HCO; also occurs at these higher temperatures
(Kharaka et al., 1985). In the 80 to 120°C temperature range thermocatalytic
degradation of kerogen produces large volumes of carboxylic acids, such as acetic
acid. Concentrations of acetic acid as high as 10,000 ppm have been found in
Cretaceous oil field waters. At higher temperatures thermal degradation of carbo-
xylic acids produces HCO, and CH, (Surdam and Crossey, 1985).
Theoretical calculations indicate that the peak generation of CO, by thermal
maturation of kerogen should occur at approximately 100°C. Concentrations of
CO, in natural gases in the Gulf Coast systematically increases from a depth of
around 2000 m to total depth ( - 5,000 m), equivalent to a temperature range of 95"
to 115°C (Franks and Forester, 1984). Analyses of interstitial shale waters from a
Gulf Coast well (Weaver and Beck, 1971) indicate a high HCO, content with a
maximum concentration at about 5,000 m, - 100°C. Weaver and Beck (1971)
537

suggested that the CO, was produced by the decomposition of organic matter and
that it dissolved the calcite that was deposited with the shale.
True pH values (calculated) of pore waters can be from 0.5 to 2.5 pH units lower
than the pH values measured at the well head (Kharaka et al., 1985). They provide
an example where surface measured pH values (6.5 and 6.2) indicate chlorite should
be the stable phase. The calculated in situ pH value (4.0) indicates kaolinite should
be the stable phase. Kaolinite is the physil actually present in core samples (Kaiser
and Richman, 1981).
Fig. 8-8, from Curtis (1978), summarizes the relative formation of CO, from
organic matter as a function of depth and various processes. The development of
CO, during Stage IV coincides with the thermocatalytic formation of organic acids.
These CO, and organic acids are presumed to sweep through the sandstones
dissolving carbonates and aluminosilicates, creating secondary porosity. On the
other hand, Bjerlykke (1974) has suggested that in many basins there is not enough
kerogen to produce the necessary CO, to account for the suggested increase in
secondary porosity. He suggested that clay mineral reactions may account for
leaching during deep burial (2-4 km). Weaver (1960) first reported the association

---
rate of Mroduction of aqmic mattor-
&I- corbm dioJdm/ bicabomtr
Fig. 8-8. Carbon dioxide production within different diagenetic zones. Numbers are carbon isotope
composition. From Curtis, 1978. Reproduced by permission of the Geol. SOC.from Jour. Geol. SOC.
London, 135.
538

between hydrocarbon production and expandable clays and suggested that inter-
layer water released during the conversion of smectite to illite was instrumental in
the migration of oil at depth; presumably these fluids also have an effect on the
development of secondary porosity and physil formation.
Surdam and Crossey (1985) discussed the role of organic acids in dissolving
minerals and promoting inorganic diagenesis in porous rocks. The petroleum
geologists are particularly interested in the development of secondary porosity in
sandstones by the dissolution of carbonate and aluminosilicate minerals. Dissolu-
tion of the latter phase provides the raw material for the formation of secondary
physils.
Al solubility and mobility are two of the major concerns in explaining the growth of
most physils in marine-like waters. At 100°C the solubility of many aluminosilicate
minerals can be elevated by the presence of carboxylic acids. Monofunctional
carboxylic acids typically elevate the solubility of A1 by an order of magnitude or
more. Difunctional carboxylic acids and phenols are particularly good chelating
agents for Al, allowing it to be transported. The role of the organic acids is
essentially the same as in soils.
In soils, organic acids originate in the A horizon and transport Al and other
cations to the B horizon. In the subsurface the organic acids originate in the
muds-shales and migrate into sands where they aid in dissolving and transporting A1
and presumably other cations. When the A1 complex is destabilized, kaolinite is
probably formed. This would account for the late stage kaolinite present in many
sandstones. It is also likely that organic acids complex A1 while they are in the shale.
The presence of soluble organic matter, presumably largely humic and fulvic acids,
in the muds-shales of the Gulf Coast is easily observed when the physilite samples
are dispersed in distilled water. The water color increases from light brown to very
dark brown with increasing depth of samples. A distinct increase in color occurs at a
burial temperature of approximately 80°C (Weaver and Beck, 1971).
Surdam and Crossey (1985) suggest that the reduction of the ferric Fe in smectite
as it is converted to illite is balanced by the oxidation of kerogen and enhances the
release of organic acid groups during thermal degradation.
We cannot consider the chemical composition of sandstone water in detail. In
general the salinity of formation waters increases with depth but in some areas there
is a decrease with depth or no significant change. Na is the dominant cation (70 to
90%) in marine sandstones, followed by Ca. Mg concentrations are generally lower
than in sea water and decrease with increasing temperature (Kharaka et a[., 1985).
Little is known about the composition of the interstitial mud-shale water that
migrated into the sandstones. Water squeezed from muds at shallow burial depths is
probably similar to sea water. After a few thousand feet of burial the interlayer
cations in montmorillonitic mud are more abundant than the cations in the pore
water. Analyses of exchange cations and interstitial cations from Gulf Coast core
samples suggest Mg is preferentially transferred from mud to sand (Weaver and
Beck, 1971).
Another factor influencing the composition and concentration of mud water
transferred to sandstones is the semipermeable membrane effect. Compacted clays
539

and shales serve as semipermeable membranes retarding the flow of dissolved ions
with respect to water. There is a difference of opinion as to whether membrane
filtration is effective in the natural system, but the process can be used to explain
the presence of relatively fresh deep sandstone waters in some basins (Kharaka et
al., 1985).
Diagenetic reactions in shales, starting at about 50°C, are believed by some to be
a major source of Si for secondary mineral formation in sandstones. Boles and
Franks (1979) suggested the following reaction for the transformation of smectite to
illite:

3.93 K ++ 1.57 smectite - > illite + 1.57 Na+ + 3.14 Ca2++ 4.28 Mg2+
+4.78 Fe3++ 24.66 Si4++ 57 02-+ 11.4 OH- + 15.7 H 2 0

A1 is conserved as the smectite is partially destroyed. At least a portion of the Si4+,


Fe2+, Mg2+, Ca2+, and H,O probably migrates into adjacent sandstones. In the
Eocene of Texas, much of the Si4+precipitated to authigenic quartz overgrowths. At
shallow depths the Ca2+forms calcite cement. At temperatures > 100°C the FeZf
and Mg2+ react with kaolinite to form chlorite and/or with calcite to produce
ankerite.
In addition to ions from an external source transported into sandstones, free ions
are generated locally. Ions are released by the dissolution of the more soluble
components in sandstones as water chemistry and temperature change. In deeply
buried sandstones pressure solution between quartz grains can be a major source of
silica.
Hurst and Irwin (1982) compiled some of the data on the distribution of
secondary physils in sandstones deposited in various environments. Kaolinite is
commonly the only diagenetic physil in fluvial sandstones; deltaic, brackish-water,
and marine sandstones contain kaolinite and/or chlorite. Illite and I/S are less
common than the other two physils and do not have any particular environmental
affinity.
In pure quartz sandstones the volume of pore space filled by secondary physils,
most commonly kaolinite, is very small. Required A1 must be transported in by
circulating waters. Because the amount of kaolinite in quartz sand is small, relatively
little transported pore water is necessary. The water is mostly meteoric (Blatt, 1979).
Where secondary smectite, illite, or chlorite are present in quartz sandstones the
sands were flushed with ion-rich water from subjacent physilites and/or ions were
produced by dissolution of ferromagnesian heavy minerals. In less mineralogically
mature sands such as arkoses, lithic sands, graywackes, etc., a large portion of the
ions for the formation of authigenic physils are derived from internal sources. As
the needed transport of pore water is minimal, authigenic physils can form at depth
where temperatures are higher. Illite is most likely to form in arkoses and smectite
(and chloritic physils) in volcanoclastic sandstones (Blatt, 1979).
540

Examples

I have discussed some of the factors that influence the authigenic (or diagenetic)
formation of physils in sandstone. Some specific examples will illustrate the com-
plex interrelation of these various factors.

Kaolinite
Sheldon (1964) listed 42 formations, ranging in age from Miocene to Ordovician,
that contained authigenic kaolinite. Most of the sandstones have water salinities less
than sea water and p H values less than 7.0.
Fuchtbauer (1967) found authigenic well-crystallized kaolinite in clean Jurassic
sandstones of Germany. The degree of crystallinity decreases with increasing clay
content. Some of the kaolinite formed from feldspar in an acid environment created
by CO, released during the early stages of coalification. With depth, kaolinite is
believed to be converted to chlorite. Silica is liberated to form quartz overgrowths.
In the Cretaceous of Alberta, Ghent and Miller (1974) found that in quartz-rich
sandstones authigenic kaolinite formed later than quartz cement. In feldspar-rich
sandstones it was associated with authigenic chlorite, calcite, and quartz. Sarkisyan
(1972) reported that when pore waters are acid, regeneration of detrital kaolinite
and kaolinization of clay cements occur at depths of 1,000 to 2,000 m. At depths of
2,000 to 3,000 m kaolinite can form from SiO, (dissolution of quartz) and A1,0,
(dissolution of feldspar).
Kaolinite not only forms from feldspar but from a number of other minerals. In
the North Sea the kaolinite in the Jurassic Brent Sandstone formed from muscovite,
which caused a reduction in porosity (Bjmlykke, 1984). Also in the North Sea area
below 900 m detrital chlorite is replaced by kaolinite (Huggett, 1986). In some of
the North Sea sandstones kaolinite crystallizes before secondary quartz in fluvial
sandstones but postdates secondary quartz in marine sandstones. The sequence is
presumably related to pore water composition (Hurst, 1984).
Isotopic studies of secondary kaolinite can provide information on the composi-
tion and temperature of the formation water. Oxygen isotope data indicate the
kaolinite in the shallow Milk River Formation of southeastern Alberta was de-
posited from meteoric water (Longstaffe, 1984a). Hydrogen isotope data obtained
from kaolinite in the Tuscaloosa Sandstone from Louisiana and Mississippi indicate
that in the shallow sandstones kaolinite formed from meteoric waters and in the
deeper sandstones formed from waters derived from shales during the conversion of
smectite to I/S (Suchecki, 1984).
Oxygen and hydrogen isotopic analyses of kaolinite in the Tertiary sandstones of
the Texas Gulf Coast suggest kaolinite is formed at temperatures higher than 70°C
from diagenetically evolved sea water. The formation of kaolinite coincides with or
immediately follows the alteration of feldspar and dissolution of carbonates. Calcu-
lated pH values were approximately 5 to 6. The low pH was apparently caused by
the generation of CO, from organic matter at a temperature of about 100°C
(Franks and Forester, 1984).
Though kaolinite is most commonly the first physil to crystallize (along with
541

quartz), it can also crystallize after more complex physils precipitate and the water
chemistry is simplified. For example, in some parts of the Lower Tuscaloosa
sandstones (Upper Cretaceous) of Mississippi and Louisiana, kaolinite precipitated
after chlorite (pore liner) (Dahl, 1984); a similar relation occurs in the Cretaceous
Belly River sandstone, Alberta (Longstaffe, 1984), and in the North Sea (Huggett,
1986). In the Neogene sandstones of the Bengal Basin, Bangladesh (Imam and
Shaw, 1985), chlorite and illite, along with siderite and quartz, form early. During
an acid phase generated by the maturation of organic matter, feldspars and
carbonates were dissolved and kaolinite precipitated as pore fillings,
In many cases where kaolinite forms early it is later replaced by chlorite or illite.
In a review of Russian studies Shutov et al. (1970) reported that during burial,
there is a gradual transformation of kaolinite to dickite through the stage of
mixed-layer growths. During deep burial, veins of dickite are formed. Under strong
stress conditions, dickite is transformed to nacrite.
In the Cretaceous sandstones of Cameroon, dickite is formed in the temperature
range of 70 to 90°C (Dunoyer de Segonzac, 1969). Though “elevated” or diagenetic
temperatures are required to produce dickite, the temperature of formation appears
to vary in response to varying chemical and physical conditions.
Kossovskaya and Shutov (1963) reported that in quartz-kaolinite sandstones, the
kaolinite is converted to dickite when the pressure-temperature reaches the stage of
deep-seated epigenesis, and to pyrophyllite at the metagenesis stage (anchizone).
The pH is acid to neutral. They suggest that time influences the transformation
temperature. Dickite was observed in lower Paleozoic rocks at 1,000 to 1,500 m and
in Mesozoic rocks at 2,500 m.
In sandstones that have a complex diagenetic sequence - commonly kaolinite,
illite, and chlorite - kaolinite usually formed earliest and is present in the shallower
sandstones. An example is the Cretaceous Muddy Sandstone, Powder River Basin,
Wyoming. Almond and Davies (1979) identified three diagenetic zones in a vertical
interval of about 50 to 70 m. The upper zone contains diagenetic kaolinite and
quartz; the middle zone contains kaolinite plus I/S and traces of chlorite; the lower
zone contains kaolinite plus Fe-chlorite and Fe-rich I/S. The authigenic physils are
developed sequentially within the pores. In the upper zone kaolinite books occur as
pore fillings; in the middle zone detrital grains are lined with kaolinite, which is
covered with a halo of I/S; in the lower zone chlorite rosettes are dispersed
throughout the I/S pore lining or completely cover it.
The vertical zonation and sequential arrangement of the physils indicates they
crystallized from moving pore-fluids. The moving fluids were principally ground
water. As the ground water moves through a sediment pile, the more labile grains,
primarily feldspar, are dissolved and the water becomes enriched in ions. The
process of dissolution will continue until the fluid becomes supersaturated with
respect to a mineral. At this stage a mineral, kaolinite in the Muddy Sandstone, will
crystallize. The concentration of ions not involved in the reaction will continue to
increase until the solution is in equilibrium with another mineral (I/S) which then
crystallizes, etc. Thus, as sands are progressively more deeply buried and water
continually moves from the shallower to the deeper sands (Fig. 8-9), a sequence of
542

DlAGENETlC SEOUENCE

KAOLlNlTE QUARTZ SMECTITE/ILLITE CHLORITE CALCITE

Fig. 8-9. Sequence of formation for diagenetic minerals associated with each of the three diagenetic zones
in the Muddy Formation. From Almon and Davies, 1979. Copyright 1979 Soc. Econ. Paleo. Miner.

physils are precipitated. For chlorite to precipitate, the deepest waters should have a
relatively high pH and relatively high concentrations of Mg and/or Fe.
Vertical sequential arrangements of physils are common. A well-developed se-
quence is present in a well that penetrated Upper Cretaceous to Triassic sediments
in the Sverdrup Basin, Canada (Foscolos and Powell, 1979). At a depth of about
800 m authigenic kaolinite and quartz are first observed. At 2000 m authigenic illite
and I/S are present; in part they replace the earlier former kaolinite. The 2:l clays
increase in abundance, along with chlorite, to a depth of approximately 3000 m,
where chlorite is the dominant authigenic physil in the sandstones.
The diagenetic changes that occur in the surrounding shales roughly coincide
with those that occur in the sandstones. The development of kaolinite in the
sandstone coincides with the loss of amorphous SO,, AI,O,, and Fe,O, in the
shales. The appearance of illitic material in the sandstone coincides with the
conversion of smectite to I/S, a process that involves solution and recrystallization.
The appearance of authgenic chlorite in the sandstone coincides with the conver-
sion of chloritic intergrades to chlorite in the shales and the development of new
fine-grained chlorite. The authors suggest that the change in physil composition in
shales is a continuing process of dissolution and reformation. With continuing
compaction, some of the ions generated move with the shale interstitial water into
the adjacent sandstones where they participate in the formation of authigenic clays.
The idea is interesting, but it is difficult to envision how the byproduct of the
conversion of smectite to illite (Si, Mg, Fe) could move into a sandstone and
produce an illite.

Chlorite and Ch / S
Hayes (1970) analyzed authigenic chlorites from a variety of sandstones and
found they were all the Ib polytype. The range of chemical composition of the Ib
chlorites is similar to that of the IIb chlorites. Hayes suggested that the Ib types
convert to IIb at approximately 15OOC.
543

Chlorite, like kaolinite, can form early or late. In the Lower Tuscaloosa sand-
stone of Louisiana and Mississippi, chlorite rims were precipitated shortly after
burial (500 m) from pore solutions enriched by the dissolution of ultramafic and
volcanoclastic detritus. During deep burial ( - 130 to 170°C) authigenic kaolinite
alters to chlorite (Dahl, 1984). D/H analyses suggest the early chlorite formed in a
mixed seawater-freshwater system (Suchecki, 1984). Microprobe analyses (Beskin,
1984) indicate the Fe/Mg ratio decreases below a burial depth of 3300 m, though
there is no change in the AI/Si ratio. Microprobe analyses (Curtis et al., 1984) of
chlorite samples from 5560 m have the following average composition:
.67Feo3.:0Fe,2.:0Mg1 .35)(Si2.94A11 .06 )OlO (OH)*.

The chlorite is classed as a Mg chamosite. Below 6300 m chlorite occurs as the high
temperature IIb polytype.
Oxygen isotopic analyses indicate the early chlorite in the Belly River sandstone,
Alberta, crystallized from brackish water (shoreline/delta environment) (Longs-
taffe, 1984b). Oxygen isotope analyses of a Pennsylvanian age deltaic sandstone
from north-central Texas suggest Fe-rich Ib chlorite rims formed during shallow
burial, soon after deposition, when amorphous aluminosilicates and iron oxides-hy-
droxides reacted with pore water (Land and Dutton, 1978).
The Oligocene Frio Formation of Texas contains two generations of chlorite
(Kaiser, 1984). Geochemical calculations and petrographic data indicate that in the
first stage chlorite forms at the expense of smectite grain coatings at less than 75OC.
The early formation is favored by the presence of Fe- and Mg-rich volcanic detritus.
A second stage of chlorite is postulated to form from kaolinite under geopressured
conditions (seawater) at greater than 100°C. Fig. 8-10A and 8-10B are activity
diagrams of the reactions
1.4 Kao + 2.3 Mg2- + 2.3 Fe2++ 6.2 H,O = Chl + 0.2 H4Si0: + 9.2 H +
and
1.8 Ca-Mont + 1.85 Mg2+ 1.85 Fe2+ 14.8 H,O+ +
= Chl + 0.29 Ca2++ 4.6 H4SiO: + 6.8 H +
The symbols represent the composition of formation waters. The diagrams illustrate
the shift in stability fields as a function of temperature. For example, for water of a
particular composition kaolinite should be the stable phase at 25°C (1 bar) and
chlorite at 100°C (600 bars). Table 8-3 lists the types of reactions that may occur in
sand stones.
Samples of a sandstone/shale interval from the Oligocene, Frio Formation,
Texas, indicate secondary rim Fe-chlorite is most abundant at the lower sandstone-
shale contact suggesting the chlorite formed from diagenetic fluids expelled from the
shale (depth 3117 m). The absence of chlorite on partially dissolved grain surfaces
suggest it formed before grain dissolution. Kaolinite formation is related to grain
dissolution (Moncure et al., 1984). A typical formula for the authigenic chlorite is:

( 1.67 Fe22.12 Mgl .49 ) ( si2.90 )


1.10 1
'0 (OH)8
544

-
7)
-IS -

U
7)
N
n

-75 -70 -65 -60 - 50 -4s

Fig. 8-10. Activity diagrams of reaction kaolinite = chlorite (A) and montmorillonite = chlorite (B) as a
function of temperature and pressure. SW = log activity product in sea water. Samples indicate the
composition of Frio Formation (Oligocene) waters, Texas. From Kaiser, 1984. Reprinted by permission
of American Association of Petroleum Geologists.
545

Table 8-3
Reactions that may occur in sandstones (from Kaiser, 1984) '
1. CaCO, +0.05Fe2+ = Cao,95Feo,05C03 +O.O5CaZC
2. 1.4Kao+2.3Mg2++2.3Fe2+ + 6 . 2 H 2 0 = Chl+0.2H4SiO: +9.2H+
3. l.8Ca-mont+1.85Mg2+ +1.85Fe2+ +14.8H20 = Chl+0.29Ca2+ +4.6H4Si0j +6.8Hf
4. lllite+ 1.64Mg2++ 1.89Fe2++ 8.24H20= 0.82Chl+0.6K+ + l.37H4Si0: +6.46H+
5. Plagio+1.3H+ +3.45H20 = 0.65Kao+0.3Ca2++0.7Na+ +1.4H4SiO:
6.2Ab + 2H+ + 9H20Kao+ 2Na+ +4H4SiO:
7. Ca-mont + 1.32H+ +4.78H2L = 0.78Kao+0.25Mg2++0.25Fe2+ +0.16Ca2+ + 2.44H4Si0j
8. lllite+1.1H++3.15H20 =1.15Kao+0.6+ +1.2H4SiO: +0.25Mg2+
9. Ca-mont +0.33Na+ = Na-mont +0.16Ca2+
10. Ca-mont+0.41K+ +0.57H+ +2.64H20 = 0.68111ite+0.08Mg2+ +0.25Fe2+ +0.16Ca+ - 2 + 1.62H4Si0j
11. Plagio(An30)+0,6Na+ +1.2H4SiO: =1.3Ab+0.3Ca2+ + 2 . 4 H 2 0
+
12. K-spar + Na+ = Ab K+
Reactions written conserving A1 in solid phase.

In the Eocene Wilcox sandstones of southwest Texas authigenic chlorite ap-


parently forms from kaolinite at depths below 3610 m ( - 165°C) (Boles and
Franks, 1979). The increase in chlorite in the sandstones -coincideswith its increase
in the shale where it presumably formed as a by-product of the smectite to illite
transformation. The ions necessary for the conversion of kaolinite to chlorite, Fe
and Mg, were transported from the shale to the sandstone. The composition of the
chlorite in the sandstone is typically:

(A11.8,Fe~.i3Mg1 2.7gA11 .21 )OlO (OH),

The reaction is:


3A1,Si205(OH)4+ 3.5Fe2+ 3.5Mg + + 9H20
-+ Fe3.5M~3.5A16.0Si6.0020(0H)16 + 14H+
The high A1 content is presumably due to the kaolinite precursor.
In the Carboniferous continental and deltaic sandstones of the Russian Platform
Ib chlorite formed from volcanic fragments. With depth IIb chlorite becomes
predominant and the Mg content of the chlorite increases.
Corrensite and Ch/S with varying ratios are also formed authigenically in
sandstones. Ch/S is a common physil in the West Texas Permian sediments,
apparently because of an abundance of volcanic material. It is abundant in many of
the sandstones and is commonly the ordered variety, corrensite (Dodd et al., 1955).
SEM pictures indicate the shallower corrensite has a honeycomb morphology
similar to smectite. In the deeper samples the corrensite consists of well-developed
bladed crystals similar to chlorite (Tompkins, 1981).
Ch/S is the major and commonly the only secondary physil in a thick section of
Paleogene arkoses in the Santa Ynez Mountains, California (Helmold and van de
Kamp, 1984). Deep-water turbidites are overlain by shallow-water and eventually
continental deposits. Ch/S occurs primarily as grain coatings and was formed early,
during shallow burial, following the incipient dissolution of heavy minerals and
546

feldspars. Ch/S occurs in both a random and an ordered arrangement. In one


section (Fig. 8-11) the proportion of smectite layers decreases with depth and at
about 6 km ( - 200°C) it is completely converted to chlorite. In another section
authigenic chlorite is present throughout most of the section, first occurring at 4 km
-
( l0OOC). Something other than temperature influences the extent of chloritiza-
tion. Water chemistry is presumably a factor. There is no suggestion by the authors
as to the source of the ions, but as the major physil in the shales is montmorillonite
it is likely it was present in the sandstone and was a major source of the ions for the

A
A
A

A 0
0
A

A
A
A

A
A

A
A

A GIBRALTAR ROAD SECTION


RANDOM INTERSTRATIFICATION

A GIBRALTAR ROAD SECTION


ORDERED INTERSTRATIFICATION

8 WHEELER GORGE SECTION


RANDOM INTERSTRATIFICATION

0 WHEELER GORGE SECTION


ORDERED lNTERSTRATlFlCATlON

el I I I I
0 20 40 60 80 I
O h EXPANDABLE LAYERS I N CIS CLAY

Fig. 8-11. Variation in expandability of mixed-layer chlorite/smectite in Paleogene sandstones from


Santa Ynez Mountains, California, with estimated maximum burial depth. From Helmold and van de
Kamp, 1984. Reprinted by permission of American Association of Petroleum Geologists.
547

formation of Ch/S. Extra Mg was presumably derived from the adjacent


montmorillonitic shale and volcanic rock fragments in the sandstone. Fig. 8-12
shows the morphological character of the Ch/S in the sandstone and Fig. 8-13 the
x-ray characteristics.
A thick section of Neogene geosynclinal felsic tuffs from Niigata oil field, Japan,
contains a typical diagenetic zonal arrangement of zeolites. In the shallowest portion
the volcanic fragments are coated with montmorillonite and opal that presumably
formed from the reaction of felsic glass with interstitial sea water. With depth the
-
montmorillonite gradually changes through corrensite or swelling chlorite ( 100°C)
to chlorite ( - 120°C). Authigenic albite crystallizes, from analcime and quartz, at
the same time as the chlorite. The chlorinity of the interstitial water decreases with
depth (Iijima and Utada, 1971).
The chloritization process is similar to that taking place in the geothermal areas
in Iceland. Basalt alters to smectite at temperatures up to 200°C. Interlayering with
chlorite starts at 200°C and by 230" to 250°C the smectite is completely converted
to chlorite (Kristmannsdottir, 1978).
The importance of depositional porewater chemistry on the formation of authi-
genic physils can be seen in the volcanoclastic sandstones of the Upper Cretaceous
Horsethief Formation, Montana (Almon et al., 1976). Corrensite (Fe-rich) is the
only authigenic physil in the deltaic sandstones and montmorillonite in the beach
and shallow marine sandstones. Both formed early. Thermodynamic calculations
suggest precipitation of corrensite occurred in hyposaline waters where the p H was
- 7.5 to 8.8 and the Mg/Ca ratio high, due to the dissolution of volcanic material.
Montmorillonite precipitated in sandstone containing sea water with a typical
marine Mg/Ca ratio. In some way Fe, whch is twice as abundant in the corrensite
( - 20% FeO) as in the montmorillonite, is a factor. Presumably there was more
organic material in the delta than in the marine environments.
Graywackes are sandstones with an abundance of physil matrix, commonly
chlorite. There has long been a difference of opinion as to whether the physil matrix
was detrital or diagenetic. Recent studies indicate that in many graywackes the
physil matrix is primarily diagenetic; this is particularly true of sands that originally
contained an abundance of mafic volcanic material.
In a study of Tertiary graywackes in arc-related basins along the Pacific con-
tinental margin, Galloway (1974) found that at depths greater than 300 m authi-
genic chlorite and montmorillonite formed as coatings and rims (flakes perpendicu-
lar to grain) around detrital grains. With increasing depth (900 to 3000 m) and
temperature, open pore spaces are filled with authigenic zeolite, chlorite, or
montmorillonite. Volcanic rock fragments, plagioclase feldspar, and mafic heavy
minerals were the source material for the clays and zeolites. At advanced stages of
burial, replacement of feldspar and rock fragments by chlorite and other minerals
occurred, and the matrix recrystallized and was partially replaced by chert and
micas.
In Permian volcanic rich sandstones and conglomerates, New South Wales,
chlorite (rims) is intergrown with the zeolite laumontite, calcite, and quartz. The
rocks have been buried less than 1000 m. The secondary minerals formed as a result
548

Fig. 8-12. (A) Authigenic ordered chlorite/smectite (corrensite, arrows) lining detrital quartz grains.
Remaining pore space (p) is filled by epoxy. Thin-section photomicrograph, plane light. (B) Authigenic
ordered chlorite/smectite (corrensite) coating detrital grains. Note intergranular pores (p) and the
absence of coatings at grain contacts (x). (C) Enlarged view of outlined area in B, showing characteristic
cellular morphology of chlorite/smectite. Authigenic chlorite/smectite bridging pore between two
extensively coated detrital grains. (D) Cross section of chlorite/smectite coating on feldspar grain. Initial
clay platelets are oriented parallel to grain surface (arrow), while later platelets are perpendicular to
surface. (E) Chlorite/smectite bridging pore between two coated grains. (F) Chlorite/smectite mold
formed by the intrastratal dissolution of a detrital grain of unknown composition. Samples from
Gibraltar Road section, Fig. 8-11. From Helmold and van de Kamp, 1984. Reprinted by permission of
American Association of Petroleum Geologists.

of the interaction of connate water with volcanic material at low temperatures


( - 35OC) (Raam, 1968).Chang et al. (1986) described a volcanoclastic sandstone
sequence in offshore Brazil where saponite was converted to chlorite/saponite
549

<2 MICROMETER FRACTION <2 MICROMETER FRACTION


AIR-DRIED SAMPLES GLYCOLATED SAMPLES
,?A 1 3

D
8 10 14 18 22 28 30 34 2 6 10 14 18 22 26 30 34

DEGREES 2 8 DEGREES 2 0
Fig. 8-13. Representative X-ray powder diffraction patters of the < 2 pm fraction of four sandstones
from Gibraltar Road, Fig. 8-11, C/S = chlorite/smectite, I = illite, Q = quartz, P = plagioclase, K = K-
feldspar. (A) Randomly interstratified chlorite/smectite (88% expandable interlayers). (B) Randomly
interstratified chlorite/smectite (50% expandable interlayers) plus minor illite, quartz, and feldspar. (C)
Ordered chlorite/smectite (corrensite; 42% expandable interlayers) plus quartz and feldspar. (D) Chlorite
plus randomly interstratified illite/smectite (20% expandable interlayers) and minor quartz and feldspar.
From Helmold and van de Karnp, 1984. Reprinted by permission of American Association of Petroleum
Geologists.

during burial. Ordered interstratification occurred at 60" to 70°C.


A physil-rich sandstone, graywacke, can be produced from sands that do not
contain volcanic material. In the Upper Proterozoic graywackes of southern Sweden,
feldspars and biotite have altered to chlorite, I/S, and illite to produce a physil-rich
matrix. The chlorite is Fe-rich, as is the biotite. The illite replacing biotite contains 4
+
to 7 times as much FeO MgO as the illite replacing K-feldspar (2 to 3%) (Morad,
550

Fig. 8-14. Diagenetic montmorillonite formed in a semiarid alluvial and eolian complex (Cenozoic, High
Plains, Nebraska). A. Montmorillonite-coated surface of a sand grain. B. Cross-section of thicker cement
coating showing montmorillonite “cornflake-like” particles that increase in size away from the grains
surface (g). C. Photomicrograph of vitric siltstone showing bubble-wall and rod-shaped shards, biotite
with attached glass, and clay filling voids between grains that is difficult to interpret as either detrital or
authigenic in thin section. D. Scanning electron micrograph of sample C showing that clay is well
developed authigenic montmorillonite growing on grain surfaces. No detrital clay is present, although
size analysis of this sample indicates about 25% clay-size material. From Stanley and Benson, 1979.
Copyright 1979 SOC.Econ. Paleo. Min.

1984). In the Lower Paleozoic graywackes from southeast Ireland the matrix physil
is primarily sericite formed by the replacement of feldspar and rock fragments
(Shannon, 1978).
The secondary physils in marine volcanoclastic sandstones are commonly either
chloritic or smectitic or both, or Ch/S. Those sandstones deposited under marine
conditions tend to develop chloritic minerals and those deposited under continental
551

conditions are more likely to develop secondary smectite; there are many excep-
tions. Tertiary-Holocene continental volcanic sediments in southern Guatemala
have montmorillonite as the major authigenic physil, along with hematite and
heulandite. Much of this alteration occurred within 2000 years of deposition, caused
by the introduction of groundwater (Davies et al., 1979).

Smectite
In the High Plains area of the western United States, extending from Wyoming
to New Mexico, Cenozoic semiarid alluvial and eolian sandstones have been leached
by periodic dilute groundwater solutions. Most water analyses plot in the
montmorillonite stability field. These solutions have partially dissolved heavy miner-
als, volcanic glass, and feldspars to produce secondary montmorillonite and varying
amounts of calcite, opal, and chert (Stanley and Benson, 1979). The alteration of
unstable minerals in continental sandstones starts shortly after deposition. The same
is probably true, to a lesser extent, of sands in marine environments.
Smectite along with clinoptilolite, silica, and scattered celadonite occurs in
vesicles and fractures throughout a 1500 m core of the Columbia River basalt. The
smectite, nontronite, was the first phase to form. It has the following average
composition:
(A10,23 Fei.:2 Fei.lOMgl.I2 ) (Si?..64A10.36)OlO O.l~% ~a
0.46K0.15

The authors, Benson and Teague (1982), believe the basalt alteration occurred under
the low temperature conditions that exist today.

Illite
Illite and I/S are also common authigenic physils in sandstones. Whereas
kaolinite, smectite, and chlorite can form at shallow depths and surface tempera-
tures, illite and I/S apparently require an elevated temperature to form.
In shale-rich sequences where smectite is converted to I/S with the proportion of
illite layers increasing with depth and temperature a similar reaction commonly
occurs in the interbedded sandstones. The I/S in the sandstones commonly contains
more expanded layers than that in the adjacent shales. The differences shown in Fig.
8-15, from Boles and Franks’ (1979) study of the lower Eocene Wilcox Group of
southwest Texas, is typical of the Gulf Coast Region. Note that in both the shales
and sandstones there is a considerable range of I/S ratios for a given temperature.
This indicates that composition as well as temperature is a factor in determining
how much montmorillonite will change to illite.
As in the Wilcox, the sandstones in the lower Eocene Queen City Formation of
the Texas Gulf Coast commonly contain I/S physils with 10 to 15% more smectite
layers than the I/S in the adjacent shales. When the I/S physils were saturated with
K an appreciable proportion of the smectite layers in the sandstone collapsed to
10 A, indicating the presence of highly-charged vermiculitic layers; the same treat-
ment collapsed a few layers of the shale I/S. The Li-saturation test and chemical
analysis indicate the expanded layers in the sandstone I/S have a relatively high
tetrahedral charge. This suggests that the illitization of smectite is a two-step
552

60.

00.

100. -
0.5 2 y m fraction
Hower and others (1976)
-
:
0,
120.
3
L 0
+.
L
0 0 3b
140.

f
E
H 0 < 1pm from shale
0 < 1 p m from sandstone

200
0 0 4 . 0
0 20 40 60 80 100
' 1 . Expandable layers in 1 / S clay
Fig. 8-15. Variation in percent expandable layers of I/S clays with temperature. Open circles = < 1 pm
from shale; solid circle = < 1 pm from sandstone; Eocene, S.W. Texas. Probably errors + 3 to f 6
percent smectite layers. From Boles and Franks, 1979. Copyright 1979 SOC.Econ. Paleo. Miner.

process. First, a high layer charge is created primarily by an increase in A1 in the


tetrahedral sheet; second, K moves into the highly-charged interlayer space. The K
activity in sandstones may not be less than in shales, but there may be more
competing ions in the sandstone waters (Howard, 1981).
In the Cretaceous sandstones of the Green River Basin, Wyoming, the proportion
of illite in the I/S increases with depth. This is accompanied by an increase in
discrete illite and a decrease in the weight percent of I/S. Pollastro (1985) suggests
that some I/S was destroyed by the selective cannibalization of smectite layers in
I/S to provide the components needed to make a more illitic I/S. SEM pictures
show illite fibers on earlier I/S substrates, tending to confirm the cannibalization
suggestion; however, this appearance may be the result of sample preparation.
Most of the I/S in the preceding examples apparently had montmorillonite as a
parent material and evolved in a manner similar to montmorillonite in shales. Illitic
material that grows from solution or replaces kaolinite or feldspar is commonly a
1M illite or an I/S with more than 80% illite layers. Whereas the montmorillonite >
I/S reaction is strongly temperature dependent, assuming the necessary ions are
553

Fig. 8-16. SEM of sandstone cores (North Sea) after critical-point drying, showing interrelation of illite
ribbons and quartz overgrowths. In D ribbons can be seen emerging from holes in the quartz suggesting
the quartz overgrowth formed later than the illite. Arrows in A show location of B and C. Scale in pm.
From McHardy et al., 1982. Copyright 1982 the Macaulay Land Use research Institute. Courtesy W.J.
McHardy.

present, the other reactions are primarily controlled by porewater composition and
are somewhat, if not entirely, independent of temperature. Authigenic illites gener-
ally form at higher temperatures than the other authigenic physils in sandstones but
have been reported forming at < 6 0 ° C (Hurst and Irwin, 1982) and ~ 2 0 ° C
(Longstaffe, 1984).
Fig. 8-16 shows the relation of thin IM illite ribbons (x-ray indicate I/S, 4:l) to
quartz overgrowth in a sandstone from the North Sea area. The emergence of
ribbons from holes (Fig. 8-16d) suggests the quartz overgrowths formed after the
ribbons. Many of the individual laths are only 2-3 mica layers thick. McHardy et af.
554

(1982) suggest that the smectite or expandable layers are actually a reflection of the
spaces between individual laths which are formed as they pile on top of each other
during sedimentation for x-ray analysis.
Illite and kaolinite are the major authigenic physils in Lower Permian Rotlie-
gendes desert sandstones of the southern North Sea (Almon, 1981; Rossel, 1982).
Illite, along with dolomite, quartz, and hematite, formed during shallow burial. The
illite coating is formed of small plates parallel to grain surface. This early illite may
have formed from pedogenic smectite coatings. A second stage of diagenesis is
characterized by the formation of pore filling kaolinite at the expense of feldspars.
The acid waters were probably formed by the addition of CO, derived from the
underlying Carboniferous coals. Fibrous drusy and pore-bridging illite formed
during the period of intermediate to deep burial. Much of the illite formed at the
expense of kaolinite (Fig. 8-17). The K was apparently obtained from waters derived
from the overlying Zechstein evaporite sequence, which contains both halite and
K-salt. The illite has an abnormally high Na content (average Na/K = 0.23 vs. 0.07
for other illites) which would tend to confirm that the pore brines were derived from
the Zechstein salts. The fibrous illite, in this and other sandstones, has a strong
influence on permeability. Though illite only reduces the porosity from 20 to 16%,it
reduces the permeability from 375 to 2 md.
Authigenic illite is abundant in the Triassic Keuper sandstones, Irish Sea, at an
average depth of 1000 m. The illite appears to have developed a morphological
sequence tangential (to grains) -+fibrous + platy. The composition is that of a
phengitic muscovite with Mg being less abundant in the platy illite. The illite
possibly formed from the action of brines with kaolinite (Macchi et a[., 1986). A
similar origin is proposed for the illite in the Middle Jurassic Brent Sand Formation
(estuarine and deltaic) in the northern North Sea (Hancock and Taylor, 1978;
Blanche and Whitaker, 1978). Authigenic kaolinite, from feldspar, is abundant in
the upper part of the formation. With depth kaolinite decreases and illite (with a
few expanded layers) increases. Thin section and SEM pictures indicate the illite
replaced the kaolinite (Fig. 8-18 and 8-19), preserving the kaolinite stacks but
distorting the sharp crystal boundaries. The growth of kaolinite occurred at the end
of Jurassic time when the area was uplifted and exposed to subarea1 erosion and
rain water. Following subsequent burial the sands were exposed to alkaline condi-
tions and elevated temperatures. The authors suggest that illite diagenesis occurred
simultaneously with the migration of oil into the sandstone ( - l0OOC).
On the other hand, Kantorowicz (1984) claims the illite in the Brent Sand formed
early from interstitial marine water. Later kaolinite formed from mica when
freshwaters from the overlying fluvial sediments were introduced. In the Ravenscar
Group sandstones in Yorkshire, authigenic kaolinite, dickite, and chlorite (chamo-
site) formed in non-marine sediments and illite in both marine and some non-marine
sandstones. In the marine sandstones illite was presumed to have formed from the
interaction of sea water with K-feldspar or muscovite. It was suggested the illite in
the non-marine sandstones was in some way related to the K released by the
dissolution of K-feldspar and muscovite. Presumably the water composition was
very close to the phase boundary between kaolinite and illite.
555

Fig. 8-17. Clay minerals in Rotliegend sandstones. (a) kaolinite crystal (K) surrounded by drusy illite (I)
growing from coating into the pore space; X4400. (b) Drusy illite developed at the expense of kaolinite,
X 1000. (c) Mite (I) growing from the existing illite coating around booklets of kaolinite into the kaolinite
aggregates; X 2000. (d) Illite (I) replacing outer rims of kaolinite aggregates and forming a druse of short,
leaf-like platelets (honeycomb structure); X 330. (e) Drusy illite not completely filling pore (P); x 150. (f)
Pore-bridging illite (I; X 330). From Rossel, 1982. Copyright 1982 London Clay Min. SOC.
556
557

Fig. 8-19. Kaolinite book from the Jurassic Brent Sand with illite growths developed on the edges,
x 5700. From Hancock and Taylor, 1978. Reproduced by permission of the Geol. SOC.from Jour. Geol.
SOC.London, 135.

Secondary illite and I/S (up to 15% smectite) is abundant in the eolian sand-
stones of the Permian Rotliegendes, southern North Sea (Lee et al., 1984). Delta 0l8
values of the diagenetic fluids involved in the illite formation were calculated from
the oxygen isotopic compositions of the illites, and the formation temperatures
estimated from the inferred depths at which they formed and estimated geothermal
gradients. The values are low, implying a meteoric component in the diagenetic
fluid. Inferred temperatures of illite formation are 96 to 136°C and I/S < 60°C.
K-Ar ages of the illitic physils fall in the range of 100 to 175 m.y., which is within
the time period of the development of the major graben system in the southern
North Sea. The authors concluded, on the basis of the K-Ar ages, that the diagenetic
illite formed over a large range of depths. The narrow range of ages among coarse
and fine clay size fractions indicated that the formation of illitic material occurred
during a short time period.

Fig. 8-18. Photomicrographs (A,C,E) and scanning electron micrographs (B,D,F) of authigenic clay
mineral textures in the Middle Jurassic Brent Sand Formation. A,B: Abundant interlocked ‘books’ of
authigenic kaolinite fill the pore space near the top of the Brent Sand. Note the pseudo-hexagonal platy
crystals. A: photomicrograph, X200. B: scanning electron micrograph, X1500. C,D: An ‘island‘ of
authigenic kaolinite is surrounded by later-formed illite, and some penetration of illite into the kaolinite
occurs. The shape of this kaolinite aggregate suggests that it may represent a replaced feldspar grain. C:
photomicrograph, X 175.D: scanning electron micrograph, X 730.E,F: Authigenic illite pseudomorphing
pre-existing kaolinite. The ‘book’ texture characteristic of kaolinite is clearly displayed (cf. A,B), but is
now composed of illite. Birefringence colors in thin section are those of illite. Note at high power the
ragged nature of individual (illite) crystal plates (F) as compared to B. E: Photomicrograph, X400. F:
SEM, X 1500. From Hancock and Taylor, 1978. Reprinted by permission of the Geol. SOC.from Jour.
geol. SOC. London, 135.
558

Fig. 8-20. SEM photomicrograph of authigenic minerals in St. Peter Sandstone, Ordovician, Cook
County, Illinois. A: Small kaolinite crystals on authigenic pyrite; B: K-feldspar, kaolinite and incipient
quartz overgrowths; C: filament-like illite around pressure-solution pit, on surface of encrusting quartz
overgrowth, and bridging to terminate overgrowth; D: partially illitized and dissolved F-feldspar with
quartz-grain impression on left; E: euhedral kaolinite; F: illite formed around pressure-solution pit.
From Odom et al., 1979. Copyright 1979 SOC.Econ. Paleo. Miner.
559

Authigenic illite is the principal authigenic physil in the shallow (300-400 m)


Milk River sandstone of southeastern Alberta. The temperature of formation is
estimated to have been + 15 to + 20°C. Oxygen isotopic data suggest the formation
fluids in which the illite formed was a mixture of meteoric water and sea water
(Longstaffe, 1984).
The clean ( > 99% quartz), Ordovician St. Peter Sandstone, Wisconsin and
Illinois, contains at least nine diagenetic minerals, which formed in five separate
stages (Odom et af.,1979). The first stage involved the precipitation of K-feldspar
crystals and quartz as thin overgrowths. During the second stage illite (Fig. 8-20)
and locally smectite and chlorite were precipitated, sometimes with dolomite and
calcite. During the third stage euhedral to subhedral quartz terminations crystal-
lized. The fourth stage is the formation of pyrite and the fifth stage the development
of kaolinite, which did not appear to form directly from feldspar. The mineral
sequence indicates the original pore fluids had a high K/H ratio and Si content
(K-feldspar). The ratio decreased with time as illite, smectite, and chlorite formed.
Eventually fresh water invaded the St. Peter and kaolinite formed. In the fresh water
the authigenic K-feldspar altered to illite. The authigenic minerals were apparently
precipitated from migrating pore fluids derived from underlying and overlying
carbonates.
Diagenetic illites in Eocene sandstones from Kettleman North Dome, California,
are in equilibrium with the pore water (10,000 ppm, mainly Na and Ca) at 100°C
(Merino and Ransom, 1982). The illite replaces kaolinite, quartz, and mica and
occurs as cement. Microprobe analyses indicate the illite has the composition of
phengitic muscovite (0.5 to 1.0 tetrahedral Al). The Fe and Mg values are consider-
ably lower than for shale illites. The free energies of formation were calculated to
range from -1280 to -1320 kcal/mole at 100°C and 150 bars. The free energy
decreases as the K content and the Al-for-Si substitution increases.
Dutta and Suttner (1986) studied three alluvial sandstones (Gondwana, India;
Fountain and Cutler, Colorado) containing meteoric waters, in order to establish the
effect of climate on the formation of early authigenic physils. All three sandstones
have two stages of authigenesis. During the earliest stage kaolinite, chlorite, smec-
tite, and quartz crystallized. The second stage was dominated by replacement
reactions involving the growth of illite, Fe oxide, and carbonate minerals. Early
diagenesis occurred at burial depths of tens to hundreds of meters and within a few
million years of deposition. Consequently, the nature of the authigenic mineral suite
was determined primarily by the groundwater chemistry, which was in turn con-
trolled by climate. During arid periods, ionic concentration of the ground water was
high and smectite and chlorite formed. During times of high precipitation, kaolinite
and quartz formed from the dilute pore waters. Illite occurs only at burial depths
below 1600 m ( - 75°C); most of it replaces kaolinite. The K is presumably a
product of the dissolution of K-feldspars and detrital micas in continental waters.

Discussion
It is apparent from these case histories that the distribution and combination of
authigenic physils in sandstones are extremely variable. This, in turn, indicates
560

various types and sequences of fluids have been present and, in most instances,
moved through the sandstones. There are basically three simple categories of water:
acid, alkaline with Mg, and alkaline with K.
Shallow acid waters are of meteoric origm and deep acid waters are generated by
the maturation of organic matter and the generation of CO, and organic acids. In
both situations the authigenic physil is kaolinite. The Si and A1 is commonly derived
from feldspar and/or micas or can be transported from adjacent shales.
Chlorite also forms both early and late. The early material commonly forms in
brackish to marine waters in sandstones that contain a source of Mg and Fe,
volcanoclastics, ultramafics, and/or montmorillonite. At depth kaolinite may alter
to chlorite in marine waters modified by shale dewatering.
Corrensite and Ch/S, like chlorite, tend to form in marine and brachsh water
where there is an abundant source of Mg, commonly volcanic material. The reason
Ch/S forms rather than chlorite may be because of lower Mg concentrations but it
appears more likely to be due to the mode of formation. Chlorite apparently grows
directly from solution, whereas Ch/S appears to form primarily from pre-existing
smectite. The latter process is temperature-dependent, and the proportion of chlo-
rite layers in Ch/S increases with increasing temperature.
Montmorillonite most commonly forms, in any abundance, in volcanoclastic
sandstones that are leached by groundwater. The high solubility of the volcanic
material allows the groundwater to rapidly shift to the stability field of montmoril-
lonite.
I/S is commonly present in marine sandstones that occur in thick montmoril-
lonitic shale sections. The development of the I/S in the sandstones parallels that in
the shales, but the sandstone I/S characteristically contains 10 to 20% fewer illite
layers than that in the shale; in part, this may be due to a deficiency of K.
Illite forms under a wide variety of conditions. It is found in eolian, desert,
deltaic, and marine sandstones. Waters are reported that range from meteoric to
marine. It can precipitate from solution but more commonly replaces kaolinite and
in some places feldspar. The main requirement for the formation of illite appears to
be a source of K, commonly external, and an elevated temperature. Though some
low temperatures are reported, the data suggest temperatures on the order of 100°C
or higher are required for the formation of illite.
The various authigenic physils are a major factor in reducing the porosity and
permeability of sandstones. Fibrous illite is usually the most effective permeability
reducer.

Limestones
I did not include a chapter on limestones because there has been very little work
done on the petrology of physils in limestones. As a result, we have been looking at
limestones with the SEM since 1986 and have found that many contain beautifully
developed authigenic-diagenetic physils.
561

Chapter IX

EVOLUTION OF PHYSILS AND CONTINENTS

INTRODUCTION

Some years ago, I (Weaver, 1967) published a preliminary estimate of the


composition of shales in the various geologic periods. This estimate was based on
x-ray analyses of North American shales. The data acquired since 1967 and the
published data on European, African, and South American shales indicate the
mineral distribution shown in Fig. 9-1 (an updated version of the original graph) is
basically correct. There are exceptions to the general trends which will be discussed.
It should be pointed out that these data are for physilites (non-metamorphic rocks
containing more than 50% layer silicate minerals) or “shales”. The physils in
sandstones, carbonates and evaporates have a somewhat different pattern but make
up only a small percentage of the total physil population.
Most of the physils in shale-like rocks are detrital in origin and have a relatively
high A1 content and will be referred to as A1 physils. The Mg- and Fe-rich physils
are more commonly authigenic in origin (Chapter VI).
In Fig. 9-1, the term “expanded physils” refers to smectites, primarily montmoril-
lonite, and mixed-layer illite-smectites (I/S). “Illite” refers to lOA physils contain-
ing < 10% expanded layers.
I originally estimated 5% I/S in the Lower Paleozoic shales. This value is
probably low by a factor of two, possibly as much as three. A reasonable value
would lie between 10 and 15%. As shown in the chapter on diagenesis-metamor-
phism, the last few expanded layers can exist to a burial temperature near 350°C
(greenschist facies). It should be noted that most x-ray analyses are made of the
2 pm fraction which included most of the I/S. Much of the illite and chlorite is
coarser than 2 pm. In post-Lower Paleozoic sediments, most I/S contains more
than 20% expanded layers and is easily detected.
Chlorite was estimated to increase from 5% in the younger sediments to ap-
proximately 10% in the Lower Paleozoic shales. This value was probably under-
estimated. Including chlorite-smectite and chlorite-vermiculite with chlorite would
increase these values slightly. A bigger increase is caused by the fact that the chlorite
is generally coarser than illite and is thus relatively more abundant in the > 2 pm
fraction. A realistic value should be somewhere between 10 and 15%.Shales seldom
contain more than 30% chlorite.
The most obvious temporal change in the clay mineral suite occurred in Middle
Mississippian time. The younger rocks have a relatively complex physil suite with
562

100 NORTH AMERICAN PHYSILS


m.y. 00 4
0 -PLIOCENE-
MIOCENE
OLIGOCENE
EOCENE
rALEDCME
IJ
iSMECTITIC ILLITE UAULIN

1
--CRETACEOUJ
-
L
CRETACEOUS

JURASSIC

I
2 -TRIASSIC

PERMIAN

1
DEVONIAN

SILURIAN

ORDOVICIAN
1
5

CAMBRIAN

6
PROTEROZOIi

Fig. 9-1. Distribution of physils in North American shales. Smectitic physils include I/S with < 90% I;
illite includes I/S with z 90% I; chloritic includes Ch/S and Ch/V.

expanded physils, illite, and kaolinite being relatively abundant. Chlorite is a minor
component. The Lower Paleozoic suite is dominated by illite. Chlorite and I/S with
a high illite content are relatively abundant, probably totaling between 20 and 30%.
Kaolinite values average 5% or less.
563

It has been suggested that the temporal changes are due to diagenetic changes
caused by increased burial depth and temperature, with time (Grim, 1953). I
(Weaver, 1967) suggested that the change in composition of the physil suites with
time could be due to changing source materials (tectonics), changes in the weather-
ing regime (development of land plants), and to a lesser extent, changes in ocean
chemistry. Climate and volcanism are two other major factors to be considered. All
of these mechanisms have had an effect. The movement and interaction of the
continental plates may be the controlling mechanism in that they influence most of
these factors. We will evaluate these factors in the following sections. Note that if
burial temperature is the dominant factor, then most Paleozoic physils started as
smectites presumably derived from volcanic rocks. If the Paleozoic illites were
formed by weathering or authigenically, then the temperature effects would have
been relatively minor. Temperature should perhaps be considered a homogenizer.
Temperature is important, but perhaps Paleozoic and Precambrian physilites would
be composed primarily of illite and chlorite regardless of high burial temperatures.
Let’s see.
Numerous studies of thick Cenozoic and Mesozoic physilite deposits have
demonstrated the decrease in kaolinite and the conversion of smectite to I/S and
chlorite with increasing burial depth and temperature (see Diagenesis-Metamor-
phism).
Temperatures on the order of 350°C are required to alter smectites to a white
mica-chlorite suite which contains no expandable layers. Rocks exposed to these
temperatures are usually classed as metamorphic, commonly slates and greenschist.
Thus, unless they are detrital, most of the illites and chlorites which occur in the
typical Lower Paleozoic-Precambrian shale are actually mixed-layer physils, though
they commonly contain < 5-10% expandable layers.
Fig. 9-2 illustrates the type of situation that is commonly encountered on the
flanks of continental plates. Frey (1970) has described a similar sequence from the
Upper Triassic and Lower Liassic of the Alpine region. The graphs show how the
physil suite in the Upper Cambrian Conasauga physilites of the southern Appa-
lachians changes over a lateral distance of 120 km. The changes are similar to what
occurs over a burial depth of 5,000 to 15,000 m, depending on the geothermal
gradient.
The two lower graphs show the change in physil suites of the < 2 p fraction
(clay) and the 2 to 44 pm fraction. As mentioned in Chapter 1, most physils in
Paleozoic physilites are coarser than 2 pm. Thus, most analyses of the < 2 p m
fraction do not give a true picture of the physil suite. The upper graph shows the
percentage of material in the < 2 p m fraction. Approximately 30% of the physils in
the shales are < 2 pm in size. This value decreases slightly until the beginning of
the development of slaty cleavage and then decreases rapidly. Aggregates are a
problem, but SEM studies suggest it is not the major factor up to the slate stage.
The K.I. decreases and the percent of 2M illite-phengite increases from west to east.
Partial x-ray patterns are shown in Fig. 7-18.
The materials depicted in the left portion of the diagrams are typical shales from
the western portion of the Ridge and Valley Province. They are from the western
564

20 - 50

0, 0

180

80

60

40

20
Degree 2 8
0
SHALE SLATE GREENSCHIST

ILLITE-PHENGITE PH ENG IT E

ILLITE

CHLORITE

0 20 40 60 80 100 120 km
3,000 10,800 m Burial
125 200 300 400 C

Fig. 9-2. Variations in Conasauga shale-slate as a function of burial depth and temperature.

edge of the Appalachian basin, have never been deeply buried, and have only been
subjected to mild tectonic movement. The physil suite is typical of that found in
much younger sedimentary rocks. The suite is similar to that of the Gulf Coast
Cenozoic sediments at a depth of 5000 to 6000 m.
To the right, east, and farther into the miogeosyncline, increased subsidence
caused the physilites to be buried more deeply and exposed to higher temperatures;
kaolinite was destroyed and the proportion of expanded layers decreased. These are
“typical” Paleozoic shales. The samples in the slate to greenschist facies are
adjacent to the Great Smoky fault and have been subjected to overthrusting and
deep burial. This has led to the development of slaty cleavage, extensive recrystalli-
zation and the growth of chlorite and phengite. Complete recrystallization occurred
at about 350°C.
Physilites in the stable plate interiors usually have not been deeply buried and the
physil suites reflect their detrital origin. Those along the mobile continental margins
565

30
Cenoz.
-
I Mesozoic I Paleozoic
Total Corbonate

Do lo mite
_/---------- -z

0 100 200 300 400 500 600

TIME (M.Y.)
Fig. 9-3. Relative global abundance of authigenic Mg-rich physils, kaolinite and carbonates through
geologic time. From Weaver and Beck, 1977. Copyright 1977 Elsevier Pub. Co.

are more apt to have been deeply buried and the physil suite modified by
metamorphic processes. With a reasonable geothermal gradient, it would require a
depth of 10,000 to 12,000 m to obtain a temperature of 35OOC. In the mobile belts,
it is unusual for thickness of this order of magnitude to be deposited without some
orogenic activity occurring. Thus, in ancient mobile belts, the sedimentation rocks
that would be illite-chlorite shales in the stable interior, are commonly slates and
schists. Such thickness can develop in interior basins (i.e., Anadarko Basin) and
present-day continental margins (i.e., Gulf Coast of United States) without having
been exposed to major lateral stresses. Relatively few Paleozoic-Precambrian shales
have been exposed to burial temperatures as high as 250 to 300°C. The absence of
kaolinite in most geosynclinal shales indicates they have probably been exposed to
burial temperatures as high as 150 to 200°C.
In another graph (Weaver and Beck, 1977), I attempted to show the temporal
distribution of authigenic Mg-rich physils (Fig. 9-3). The chain structure physils,
palygorskite and sepiolite, are relatively abundant in the Cenozoic and Mesozoic. As
the chain physils decrease in abundance with increasing age, Ch/S (corrensite-like)
physils become the major authigenic Mg-physil. The palygorskite refers to coastal,
as opposed to lacustrine, palygorskite formed in brackish water and the Ch/S is
that formed in hypersaline waters. The total volume of both minerals is relatively
small compared to the other physils. The factors controlling the shift are not known,
but climate and other conditions that effect water salinity are important factors.
566

MEAN GLOBAL MEAN GLOBAL


TEMPERATURE PRECIPITATION
present present
:old warm
Quaternary

10

Oligocene

Cretaceous

I Jurassic

1000 -

2000 -

3000-

I
1

4000L
I
I

I
I

4600

Fig. 9-4. Generalized temperature and precipitation history of the earth. Trends are dashed where data
are very sparse. The curves are drawn to represent postulated departures from present global means. but
only relative values are indicated. Note that the time scale is progressively expanded in younger time
units. From Frakes, 1979. Copyright 1979 Elsevier. Pub. Co.

I have included three graphs which provide some of the background information
needed to interpret the significance of the physil suites through time. Fig. 9-4
contains an estimate of mean global temperature and mean global precipitation
through geologic time. Fig. 9-5 contains an estimate of sea level fluctuations. Keep
in mind these graphs represent only “best guesses” and if the interpretation based
on the physils differs from that based on the data in the graphs, it does not mean
the physil based interpretation is incorrect. For example, the kaolin maximum in the
567

L 1ST OROEA CYCLES 2 N 0 OAOEA CYCLES ISUPERCYCLESI

-RISING -
3EIATIVE CHANGES OF SEA LEVEL
FALLING
RELATNE CHANGES OF SEA LEVEL N O T A
-RISING FALLING- TlONS

9
9
,
’300

1 1‘
L -
CAMBRIAN - M
c-o -
E
-PRECAMBRIAN
Fig. 9-5. First-and second-order global cycles of relative change of sea level during Phanerozoic time.
From Vail el al., 1977. Reprinted by permission of American Association of Petroleum Geologists.

+ + + - + + +
t + + + + t + +
+ + + L A +
lo’; AND O R T H O C N E I S S E S -t .

V V V V V V ” v

SEDIMENTARY ROCKS=

-44500- 3500-2700- f400-600-225-0-

ABSOLUTE T I M E I N MILLIONS OF Y E A R S
Fig. 9-6. Scheme of the proportional change in main rock types of the continental erosion area during
geological time. From Ronov, 1972. Copyright 1972 Elsevier Pub. Co.
568

Cretaceous (9-1) contrasts with the maximum global dry conditions that are
projected to have existed at that time (9-4).
Physils are derived primarily from volcanic rocks (lavas), feldspars (granitic rocks
and gneiss), and layer silicates (shales-slates). Fig. 9-6 shows the proportion of these
rock types (sedimentary rocks are primarily shales and slates) in the continental
erosion area during geologic time (Ronov, 1972). Ronov et al. (1980) also concluded
that during the Phanerozoic, the volume of terrestrial volcanic rocks gradually
increased (with decreasing age), with major increases during the Devonian, Triassic,
and Cretaceous. Note that in the late Precambrian and early Paleozoic granitic
(K-rich) rock were considerably more abundant than mafic lavas. Presumably this
has something to do with the relative abundance of illite in the older sediments.

PRECAMBRIAN

In the Precambrian, the earth’s surface conditions were drastically different from
those of the present time. The physils should provide information as to the nature
and magnitude of the difference. The nature of the earth’s environment in early and
middle Precambrian is extremely speculative, and there is little data on the physils
in unmetamorphosed physilites of this age. There are a number of limiting factors.
The early source rocks were igneous with compositions nearer to basic and
ultrabasic rocks rather than to acid ones (Ronov, 1972). The continental environ-
ment was probably alkaline rather than acid. The continental land mass may have
consisted of one continent, and deposition occurred largely in large inland water
bodies. Glacial deposits and iron deposits have received particular attention though
they are probably of minor importance volumetrically.
Most sediments of the Early Precambrian have suffered some degree of metamor-
phism and one can only speculate as to the original physil suite. The data suggest
chlorite is relatively more abundant than in younger sediments. This is also
suggested by the chemical data (Ronov and Midgism, 1971). This presumably
reflects the abundance of basic volcanic material. The K value (2.3% for Russian
Platform shales and 3.1% for American Platform shales) for the Archeozoic clays
and shales in less than for any other period of geologic time. Fe is more abundant in
the American Platform shales but not in the Russian Platform shales. Mg does not
appear to be appreciably more abundant in these early shales than in younger ones.
The banded iron deposits (older than 1.7 billion years) are characterized by having a
low A1 content and few detrital minerals (Govett, 1966). Presumably, fairly intense
chemical weathering occurred in the source area and Al-rich clays and bauxites
should have been formed. Even though these materials would have been trans-
formed by metamorphism, there is nothing in the chemical data to suggest high
concentrations of Al. Govett (1966) suggested that the high A1 residual material
would be deposited later and on top of the F e deposits. He presented several
examples to indicate shales overlying these early iron formations have higher A1 203
values than the average Precambrian shale and higher than underlying shales (22.5 %
569

versus 16.04%). Using the volume of iron formation, rocks and overlying shales,
Govett (1966) calculated that the source rock had a basaltic composition.
If the bulk shale contains 22.5% A1,0,, the clay fraction should contain ap-
proximately 30%. Illites, kaolinite, and dioctahedral chlorites contain this much
Al,O,; trioctahedral chlorites and chamosites do not. The Fe and Mg values
indicate these latter two clays could not have been abundant. The metamorphic
product is largely illite and the beginning material must have been largely illite.
Though K-feldspar plus kaolinite or montmorillonite can be metamorphosed to an
illite, there is not much K-feldspar in basalt.
Well crystallized illite and chlorite are the only physils in the 2000 m.y. old
continental (glacial) Gowganda Formation of Ontario, Canada (Tank and McNeely,
1970; Lindsey, 1969). The presence of Al-rich minerals (kaolinite, diaspore, pyro-
phyllite, andalusite and kyanite) in the overlying formations suggests the period of
glaciation was followed by a period of tropical weathering and the concentration of
A1 minerals in the soils (Nesbitt and Young, 1982). The argillites of the Middle
Precambrian Rove Formation of Minnesota and adjacent Canada contain musco-
vite, chlorite and septichlorite (?) (Morey, 1967).
There is a fair amount of data on the late Precambrian, but the data is sketchy.
By late Precambrian, photosynthesis was well developed and oxygen had become
relatively abundant. Pangaea was apparently fragmented into several continents and
some semblance of an ocean system developed.
Iron formations younger than 1.7 billion years are nearly all oolitic (rather than
banded as in older deposits), and A1 content is relatively high. Chamosite, which
contains the highest A1 content of the iron alumino-silicates, is the most common
clay in oolitic iron formations, both of late Precambrian and younger formations.
Under conditions when Fe is abundant, Fe-A1 silicates tend to form rather than
Al-silicates. This is also true of recent environments as well as ancient. Much of the
Precambrian chlorite could possibly have originated as bertherine (7 A chamosite).
In his discussion of the sedimentary facies of the Precambrian, Lake Superior
iron-formations, James (1954) reported that greenalite, stilpnomelane, minnesotaite,
and iron-rich chlorite were abundant clay minerals in these deposits. He believes
greenalite is the only primary mineral and the other three are the results of
metamorphism of other Fe-silicate compounds.
In a later study, Bailey and Tyler (1960) found the following clay minerals
associated with the Lake Superior iron ores: dickite and kaolinite (the most
important), nacrite, talc, pyrophyllite, 1M and 2M muscovite, various serpentines,
dioctahedral and trioctahedral chlorite and montmorillonite, palygorskite, regularly
interstratified chlorite-montmorillonite, nontronite, and chamosite. They concluded
these minerals were of hydrothermal origin and that regional variations in the clay
suite were due to differences in the temperature of the hydrothermal fluids.
A study of the 13,000 m thick Late Precambrian Belt series (approximately 1 to
1.7 m.y.) of Montana and Idaho (Maxwell and Hower, 1967) showed that 1Md and
2M illite along with a lesser amount of chlorite comprises the physil suite of the
shallow buried samples. One sample contained kaolinite. The 10 A peak in their
x-ray patterns is broad and indicates smectite or chlorite layers are interstratified
5 70

with the illite layers. With depth, the proportion of 2M illite increases from 22% to
60%. This occurs as the temperature increases from 225" to 310" (Eslinger and
Savin, 1973). The chlorite content of the bulk samples ranges from 5 to 60% and
-
averages 25% (Eslinger and Sellars, 1981) which is high for shales.
The Thick Belt section can be compared to the Gulf Coast Cenozoic where
volcanic derived montmorillonite is being converted to illitic material during burial
diagenesis. Many, probably most, of the thick clay-shale sections of Cenozoic and
Upper Mesozoic age contain a high percentage of volcanic derived montmorillonite
which is progressively altered to I/S and eventually illite with increasing depth of
burial and temperature. The problem we face is whether this scenario can realisti-
cally be extrapolated back into the Paleozoic and Precambrian. We may as well start
now, in the Precambrian, to consider the alternatives.
Various authors (Belt Symposium, 1973) describe the Belt rocks as feldspathic
quartzite, subarkosic, and highly feldspathic (20-30% K-feldspar). Winston (1973)
described sands which contain 15% microcline and 5% orthoclase. Another 20% of
the grains are sericite and kaolinite aggregates that formed from feldspar. Murray
and Duncan (1977) report some of the shales contain significant amounts of
kaolinite.
"The mineralogic characteristics of the constituent grains in the siliceous rocks that compose much of
the Belt Series are compatible with the idea that much of the ultimate source rock was granite gneiss,
in large part rather alkalic," Ross, (1976).

Thus, the petrographic data suggests the main source of he Belt physils was
feldspar and micas rather than volcanics. It is quite likely that the weathering
product of the feldspars in the Precambrian alkaline environment was 1Md illite
and I/S with a high illite content. The starting material was probably much closer
to a true illite than the volcanic derived smectite supplied to the Gulf Coast and
other Cenozoic basins.
A series of x-ray patterns of the late Precambrian Ocoee series shales and
phyllites of the southern Appalachian Mountains (Hurst and Schlee, 1962) indicates
the least metamorphosed physilites contain an I/S with 5 to 10% smectite layers.
With increasing metamorphic grades, chlorite becomes as abundant as the mica. The
sequence is similar to that of the Belt.
Reynolds (1965a,b) found that Precambrian illites from relatively unmeta-
morphosed carbonate rocks and shales were predominantly the low temperature
1Md polymorph. Most of the samples were of Late Precambrian but included
limestone as old as 2.7 billion years. Chlorite was present in some samples.
Reynolds reports samples from the Nonesuch, Rove, and Gunflint formations
contain only illite. Other samples (Weaver) from these same three formations
contain appreciable chlorite and in some Nonesuch samples, chlorite is more
abundant than illite. A sample from the Gunflint is composed of I/S ( 4 : l ) and
kaolini te.
Unmetamorphosed Late Precambrian samples from Australia contain varying
amounts of illite and chlorite. The physils in some samples are nearly all illite and in
others, nearly all chlorite. Six samples, of 13, contain appreciable smectite and one
571

contains kaolinite. The peak ratios indicate the illites have a high Fe content and the
chlorites, a high Mg content (Weaver).
The physils in the Upper Precambrian phosphate deposits of the Upper Volta,
west Africa, are illite and chlorite (Lucas et al., 1980). The Upper Precambrian
glacial deposits of Adrar De Mauritanie, West Africa have a complex mineralogy.
Various amounts of the following physils are present: illite (both di- and trioc-
tahedral), chlorite, vermiculite, smectite, kaolinite, Ch/S and I/S. Smectite com-
monly comprises 90% of the physil suite (Chamley et al, 1977).
When clays and shales, in which the physil suite is composed largely of kaolinite,
are subjected to burial temperatures on the order of 300" to 350"C, the kaolinite
(plus quartz) is converted to pyrophyllite. Thus, the presence of phyrophyllite in
very-low-grade and low-grade metamorphic rocks is a strong indication that kaolinite
was abundant in the original sediment.
Petrov (1958) reports that in the Karelian Republic and adjacent parts of
Finland, and in the Ukraine and Urals, metamorphosed Archaean rocks contain
quartzites and pyrophyllite bands and lenses. Pyrophyllite is also abundant in the
Precambrian rocks of northeastern South America (Dunoyer du Segonzac and
Chamley, 1968). Thus, it appears that as far back as the early Precambrian,
weathering conditions were intense enough to form kaolinite. The formation of
kaolinite is usually indicative of high humidity and relatively intense leaching.
Pyrophyllite can be used as a climatic indicator in Precambrian rocks in the fashion
that kaolinite is used in younger rocks. In shales, kaolinite is usually a minor
component and is destroyed at burial temperatures in the 150" and 200°C tempera-
ture range. Therefore, the absence of kaolinite in old physilites or physilites
subjected to burial depths of 4 to 6 km does not indicate it was not present.
The physils present in the relatively unmetamorphosed and metamorphosed
Precambrian physilites suggests that the chemical conditions that control the
formation of physils were not significantly different from those that exist today. In
addition to 2M illite and chlorite, kaolinite, I/S, 1Md illite, Ch/S, chamosite, and
smectite are present in abundance in Precambrian rocks. The relative abundance of
illite and chlorite is apparently due to the high burial temperatures to which many
of these rocks have been subjected. It is likely that illitic material was formed both
by alteration of feldspar and acid volcanics, and diagenesis of montmorillonite.

EARLY PALEOZOIC

The distribution and mineralogy of physil suites is strongly influenced by


orogenic activity. Thus, it is important that we discuss the temporal distribution of
physils in the context of continental drift, plate collision, and volcanic activity. The
discussion is restricted to North America, South America, Europe, Africa, and the
Atlantic Ocean. The review of the available physil literature is far from complete
but I think it is thorough enough to illustrate the overall gross distributipn patterns
for the various periods.
572

CAMBRIAN

Fig. 9-7. Cambrian paleogeography. from Mintz, 1981.

Six major paleocontinents existed during Cambrian Time (Fig. 9-7). The three
supercontinents we are concerned with are Laurentia (North America, Greenland,
Scotland and The Chukotoski Peninsula of the Eastern USSR), Gondwana (South
America, Florida, Africa, Australia, Antarctica, India, Tibet, Iran, Saudi Arabia,
Turkey, and southern Europe) and Baltica (Northern Europe and Russia). The
following discussion is based largely on papers by Bambach et al. (1980) and Condie
(1982). In the late Cambrian all the continents except Baltica and China straddled
the equator. During the Ordovician and Silurian Gondwana swept across the south
pole. Laurentia and Baltica rotated counterclockwise and drew close together
separated by the Proto-Atlantic Ocean. They collided in the Devonian, beginning
573

the assembly of Pangaea. Further collisions during the Pennsylvanian through


Permian produced the supercontinent Pangaea.
The earliest mountain building event along the eastern margin of Laurentia
occurred in the Middle Ordovician (Taconic Orogeny) (Fig. 9-9). Later a borderland
arose along the continental margin in the same region and shed a large volume of
sediments westward (Queenston Delta) onto the former continental shelf. Earlier
shelf sediments were derived from the western cratonic interior. The continental
margin continued to be uplifted and underwent significant erosion in the late
Ordovician. Silurian sandstones and conglomerates coarsen eastward indicating
highlands were still present. Evaporites were deposited to the west. In the late
Silurian the geosynclinal strata in Britain and Scandinavia were strongly folded
(Caledonian orogeny). Volcanism occurred during the Ordovician and Silurian on
both sides of the Proto-Atlantic Ocean. Prior to the Silurian there were no land
plants and rocks were presumably eroded and transported as sediments as soon as
they were converted to a fine size. Thus, chemical weathering and the formation of
new clay minerals in the lughland source areas was probably much less than after
the development of land plants and a decrease in the rates of erosion.
Laurentia and Baltica collided in the Early Devonian (Arcadian Orogeny)
forming an elongate mountain belt (Fig. 9-11). The collision began in Late Silurian
with the Caledonian orogeny in north-western Baltica. Large volumes of sediments
were derived from the mountain belt created during the Arcadian orogeny and
spread to both the east and west. A thick sequence of redbeds was deposited, the
Catskill Formation (Upper Devonian) in North America and the Old Red Sand-
stone in Europe. These redbeds interfinger with shallow marine sediments. The
mountain belt was a source for large quantities of sediments throughout the
remainder of the Paleozoic.
Previous to the Acadian Orogeny (middle Devonian) shallow seas had dominated
the continental platforms. From the Devonian to the Triassic the continental plates
continued to emerge and shallow oceans were restricted to the continental margins.
Nonmarine rocks became increasingly abundant.
Throughout the ,remainder of the Paleozoic the separate continental blocks
continued to converge. In the Mississippian, South America approached North
America in the area of the Ouachita Geosyncline (Fig. 9-13). A volcanic arc
developed and supplied volcanic material to a large area of North America.
In general, the pre-Pennsylvanian Paleozoic physilites have a simpler physil suite
than younger physilites. Tlus is true of North America and Europe, the contrast is
perhaps less in South America and Africa. The lower Paleozoic physil suite is
composed largely of illite and chlorite. A significant portion, probably on the order
of 50%, of the illites contain 10% or more interlayers of smectite (K.I. > 3.0).
Relatively few illites in shales have no smectite layers (K.I. < 1.5). The illite in most
samples is a mixture of the 2M (high temperature) and 1Md (low temperature)
polytypes. The former is more abundant in the coarser fraction and the latter in the
finer fraction (Velde and Hower, 1963; Weaver and Associates, 1984). The 1Md
polytype (actually I/S) is the dominant illite in the c 1 p m fraction and commonly
in the < 2 pm fraction. The coarser ( > 1 < 10 pm) material commonly contains 30
574

M.Y. 0 10 20 30 4 ox
b
t I I I
M I S S .

... .
41

367 "
II @ * . * .
0. K-BENTONITE

41

~ E ~ O N I A N

416
,,
f * .
.
446
S I L U R I A N

11
..' . ..
*.
0 .

0 .

* .
0

0.
K-BENTONITE
o R n.

.
41

509

.
0.
0
2

C A M B R I A N

to 80% 2M illite. The average chlorite content is likely between 10 and 15%. The
kaolinite content is probably less than 5%. The chlorite has not been as extensively
studied as the illites but it appears that much of the chlorite that is not detrital
contains a small percentage of expandable layers.
Fig. 9-8 illustrates how the physil suite varies in a typical early Paleozoic
stratigraphic section from near the middle of the Appalachian Basin in central
Pennsylvania (Ridge and Valley Province). The physilite portions of twenty-two
575

formations from the Lower Cambrian to the Upper Mississippian were sampled.
Fig. 9-8 shows the percent of chlorite in the < 2 p fraction; the remaining material
is illite (mostly I/S with less-than 10% S). The Mississippian physilites contain a
minor amount of kaolinite. The kaolinite possibly heralded the major change that
occurred near the top of the Mississippian.
For much of the interval the physilites contain less than 10% chlorite. There are
three excursions where the chlorite content increases to 20 to 30%.The high illite,
low chlorite physilites are largely present in shallow shelf miogeosynclinal carbonates
and shales. The chlorite rich physilites in the Upper Devonian and Upper Ordovi-
cian are thick deltaic exogeosynclinal shales (Catskill-Chemung Formations and
Reedsville-Martinsburg formations respectively). The two delta are called the
Catskill and Queenston deltas. The high chlorite in the Upper Cambrian does not
seem to fit this pattern and is probably evaporitic. K-bentonite beds are present
immediately below the high chlorite Ordovician and Devonian sequences and Ch/S
is abundant in the Lower Ordovician and to a lesser extent, the Lower Devonian
carbonates (residue). The sequence is Ch/S, K-bentonite (1,’s) and chlorite. All
three are products of volcanic material either of different composition and/or
alteration under different environmental conditions.
During the Cambrian and Early Ordovician most of the illitic detritus supplied to
the Appalachan Basin was derived from the western craton. Mountain building
along the eastern margin of Laurentia began in the Middle Ordovician (Taconic
Orogeny) and the primary sediment source was from the east. The eastern moun-
tains became a major source of sediments in the Upper Ordovician and the
chlorite-rich deltaic sediments were deposited. As tectonic activity decreased
miogeosynclinal conditions returned, characterized by illite-rich physilites and
calcareous physilites. Laurentia and Baltica collided in the Early Devonian (Arcadian
orogeny) and by Late Devonian another thick clastic wedge of exogeosynclinal,
chlorite-rich sediments, derived from the east, were deposited. During the late stages
of delta formation, when sand-rich sediments were deposited, the chlorite content
apparently decreased. Both mountain building periods were preceded by the devel-
opment of volcanic island arcs.
The distribution of the physil minerals appears to be closely related to tectonic
activity. The Ch/S and K-bentonite beds are evidence that volcanic activity (island
arcs) occurred during the early stage of mountain building. The Ch/S in the
carbonate rocks apparently formed from scattered and dispersed fine volcanic
material. The first occurrence of volcanic ash beds approximately coincides with a
change in the type of feldspar supplied to the basin. K-feldspar is the dominant
feldspar in Cambrian and Lower Ordovician sediments. Near the top of the
Ordovician Na-feldspar becomes predominant. This is related to the change in
source from the west and north (granitic) to the east (metamorphic) (Weaver, 1961).
The high chlorite content of the deltaic exogeosynclinal shales indicates an
increase in Mg and Fe (mafic component) which suggests the eastern orogenic belt
had a higher content of basaltic material than the interior craton. This is presuma-
bly due to the incorporation of oceanic basalt and other volcanic material into the
eastern mobile belt.
576

The chlorite may be detrital, having formed during metamorphism and/or


hydrothermal activity in the orogenic belts, or the basaltic material may have altered
to montmorillonite which was transported to the Appalachan Basin. During burial
metamorphism the montmorillonite could have been converted to illite and chlorite.
The studies of diagenetic reactions in montmorillonite-rich physilites suggests that a
maximum of 15 to 20% chlorite is produced as montmorillonite is converted to illite
and chlorite. However, it is unlikely that basalts comprised enough of the source
area to provide a smectite dominated physil suite. It is likely that much of the
chlorite is detrital but some may have formed during burial metamorphism. In any
event the high chlorite content suggests the source rock had a relatively high content
of mafic material, which in turn must be related the orogenic activity in the source
area.
Another unlikely possibility that has to be considered is climate. Chlorite is easily
weathered under humid and even temperate conditions but tends to be preserved in
cold climates. Thus, it is possible that the eastern mountain ranges had attained
sufficient height during the late Ordovician and late Devonian that less chlorite was
destroyed by weathering than in the early stages of mountain building and after
they had been lowered by erosion. Of course, the climate could simply have been
cooler during the two time periods in which the chlorite-rich physilites were
deposited. As Axelrod (1981) notes, explosive volcanism can rapidly reduce temper-
ature. This discussion illustrates some of the problems involved in determining the
geologic significance of physils.

CAMBRIAN

North America

During Cambrian time much of North America was a low relief craton which was
transgressed by a shallow sea until approximately two-thirds of the continent was
covered by the end of the period.
There are relatively little data available on the composition of the Cambrian
physilites. The Appalachian Basin physilites contain from 0 to 40% chlorite,
averaging 11%.Kaolinite averages 3%. The other physils are illite and minor I/S.
Mg-rich chlorite, Ch/S, and biotite are abundant in some Upper Cambrian
carbonates. In the Ouachita Basin rocks chlorite ranges from 0 to 15%,averaging 7%
(Weaver). These sediments were derived from the craton (Transcontinental Arch).
West of the Arch in Montana the physilites have a similar composition containing
from 0 to 30% chlorite and averaging less than 10% with minor amount of I/S
(Lebaner, 1964; Leckie, 1962). Glauconite-rich rocks occur in Minnesota, Wisconsin
(Francona Formation) and Tennessee (Conasauga). In general they are not well
crystallized. The only physil in some Illinois shales is 1Md illite (Velde and Hower,
1963. The compositional range of the Upper Cambrian Conasauga Formation of the
571

S.E. United States is shown in Fig. 9-2. Illite, chlorite, kaolinite, and abundant I/S
are present in the shallow buried physilites (shales). Glauconite and biotite are also
relatively abundant. Middle and Upper Cambrian rock from Newfoundland are
composed largely of 2M illite, 10% chlorite and minor I/S. Some of the chlorite
contains expanded layers (Suchecki et al., 1977).
In North America during the Early Cambrian the land area was the Canadian
Shield and the Transcontinental arch and adjacent metamorphic and volcanic rocks
which extended southwest through the central United States. The scattered data
suggests there is nothing unique about the Cambrian physil suite. The Upper
Cambrian physilites (Fig. 9-2) from the western portion of the Appalachian Basin
(eastern flank of Transcontinental Arch) have a physil suite similar to that of the
Miocene Mississippian Delta physilites at a depth of 6000 m ( - 120°C). Both the
Mississippi Delta and Conasauga sediments were derived from the same craton but
500 m.y. apart. The composition of the source rocks may not have been too
different. K-Ar data (Chapter VII) indicate the coarse micas in the Conasauga were
derived from the Greenville metamorphic rocks to the west of the Appalachian
Basin. The K-Ar apparent age of the < 2 p m illite and I/S is less than the
stratigraphic age suggesting that an appreciable portion was originally montmoril-
lonite, presumably derived from volcanic rocks or montmorillonite-rich physilites.
The source area was, in general, similar to that in the present Mississippi River
Basin, a mixture of mica-chlorite metamorphic rocks (equivalent to the present
illite-chlorite Paleozoic physilites of the eastern portion of the drainage basin) and
volcanic and/or montmorillonitic sedimentary rocks (equivalent to the Cretaceous,
Tertiary, and Pleistocene montmorillonite-rich sedimentary rocks on the western
two-thirds of the Mississippi River Basin). The low concentration of kaolinite and
chlorite indicates weathering was moderate and the climate was presumably tem-
perate rather than tropical or arctic.

Europe, Africa

Cambrian sandstones from the central Sahara (Millot, 1970) and the Russian
Platform (Shutov et al., 1970) have a high kaolinite content. In the African
sandstone the kaolinite is believed to be secondary from mica and in the Russian
sandstone it is believed to be a primary constituent. Effusive - terrigenous material
from the Skamara Basin of the southern Urals contain illite and Ch/S (Rateev et
al., 1980). Marine Cambrian shales of Israel contain a physil suite of 90% illite and
10% kaolinite (Bentor et al., 1963). Cambrian rocks from the Central Massif of
France are composed largely of illite but contain both montmorillonite and corre-
nsite (Dunoyer de Segonzac, 1969). The Middle Cambrian shales of Swedan
(Billingen) and Norway (Oslo) contain illite and small amounts of chlorite; the
Upper Cambrian shales contain only illite (Bjorlykke, 1974).
578

ORDOVICIAN

North America

Conditions remained quiet during the Lower Ordovician as the sea transgressed
westward. Tidal flat carbonates were the major rock deposited along the southern
and eastern flanks of the Transcontinental arch and Canadian Shield. The physil
suite is quite distinctive. Varying amounts of illite and chlorite are present but the
carbonate rocks are characterized by the presence of mixed-layer Ch/S. Ch/S in
significant amounts is present in a belt extending from West Texas (El Paso, Alsate,
Marathon Formations), through Texas (Ellenburger), Oklahoma (Arbuckle), War-
rior Basin (Knox), then North through the Ridge and Valley Province (Georgia,
Tennessee, Virginia) (Weaver, 1961) and into Western Newfoundland (Cow Head
Breccia (Suchecki et al., 1977)), where it is also present in the lower portion of the
Middle Ordovician. Authgenic, idiomorphic quartz is commonly present in the
carbonate rocks containing Ch/S. Ch/S occurs in a few Upper Cambrian carbonates
and in Oklahoma it is present in Middle Ordovician sandstones. The information is
not complete enough to determine how far to the northwest the Ch/S extends. It
appears to give way to a predominantly illite suite with minor I/S in the Williston
Basin of Montana.
Ch/S has a variety of origins but it is commonly associated with relatively basic
volcanic rock. The Lower Ordovician Ch/S was probably derived from volcanic
material. During the early Ordovician the Proto-Atlantic Ocean or Iapetus Ocean
separated North America from Baltica (Fig. 9-9). As the Iapetus Ocean narrowed a
volcanic island arc developed along the eastern flank of North America (Wilson,
1966) and supplied volcanic material to the carbonate shelf. In Newfoundland much
of the volcanic material was apparently incorporated directly into the sediments in
volcanogenic sandstones. Farther south the relatively basic volcanic material was
apparently delivered to the carbonate shelf as volcanic ash, though there may have
been transport of fine detritus by marine currents. Some of the chloritic material
may have formed by weathering on the island arc.
Across the Iapetus Ocean, along the west coast of Baltica (Norway), a similar
sequence of events occurred. Illite is the only physil present in the Upper Cambrian
and early Lower Ordovician epicontinental shales and carbonates west of the Baltic
Shield. Chlorite (10 to 5 5 % , average 25 to 30%, calculated by Weaver) first occurs in
the late Lower Ordovician rocks and remains relatively high throughout the Ordovi-
cian section (Bjorlykke, 1974). The increase in chlorite closely coincides with the
time of the increase in Ch/S in Newfoundland. The first occurrence of Ch/S in the
United States has not been established but in some areas appears to be slightly
above the base of the Ordovician section.
In Norway the increase in chlorite is correlated with the uplift (island arc) and
erosion of basic lavas in the eugeosyncline west of the epicontinental sea. Thus, on
both sides of the Iapetus Ocean the increase in Mg-rich chloritic material is
associated with the development of volcanic island arcs.
579

Fig. 9-9. Ordovician paleogeography.From Mintz, 1981.

Lower and Middle Ordovician sediments in Wales are composed largely of illite
(anchizone grade), I/S and chlorite. Chlorite ranges from 2 to 60%, averaging
approximately 18% (calculated by Weaver). The source rocks were Precambrian
basement rocks and volcanic ash and lava fragments (Bjorlykke, 1971).
Whittington and Hughes (1972), on the basis of faunal provinces, suggested that,
in the early Ordovician, western Norway was not part of Baltica but was on the
western side of the Iapetus Ocean, north of Newfoundland. The physil data would
support this unless island arcs developed simultaneously on both sides of the ocean,
presumably possible.
Why Ch/S in one area and chlorite in the other? The sequence of formation of
chlorite from volcanic material is: smectite to Ch/S to chlorite, with increasing
5 80

temperature. X-ray patterns indicate the illitic material in the Newfoundland


material contains 5 to 10% expanded layers and probably has been exposed to a
maximum burial temperature of approximately 200°C. The Norwegian and Welsh
illites have much narrower illite peaks and have apparently been exposed to a higher
temperature, probably high enough to convert Ch/S to chlorite.
The physil sequence in the southern Urals is similar to that in North America
(Rateev et al., 1980). During the Upper Cambrian and Ordovician Fe-rich I/S (10
to 20% S) and a minor amount of chlorite and Ch/S was transported to the
Skamara Basin from the sialic craton to the northwest. During the Ordovician
volcanic tuffs and pyroclastic material (liparitic to andesitic) were supplied to the
basin from the east. The physil suite now present in the Skamara Basin consists of
detrital illite and abundant Ch/S and chlorite that originated from recrystallization
of the volcanic material during burial. The timing of events does not coincide with
those in North America but the sequence is similar.
Two major questions arise. Why is I/S (including 1Md illite) relatively abundant
in the Late Precambrian and Cambrian and Ch/S abundant in the Lower Ordovi-
cian? At what stage did these two physils form?
It is well established that I/S (10 to 20% S) can form from montmorillonite
subjected to burial temperatures of 150" to 250°C. It is well established, but less
appreciated, that K-feldspar alters to illite and/or I/S during weathering under a
variety of climatic conditions (see Chapter 111) and can form in some alkaline lakes
(see Chapter IV). K-feldspar is not now a major source of illite but it is quite likely
that in the late Precambrian and early Paleozoic, when land plants did not exist and
alkaline conditions prevailed on much of the land area, that K-feldspar was the
major source of 1Md illite and I/S. For example, The K-Ar apparent ages of the
0.2-2 pm fraction of the Conasauga I/S (Weaver and Associates, 1984) are the
same as the stratigraphic age, indicating they could have formed in the source area.
Some of the Ch/S may have formed in soils in the source area but most of the
occurrences suggest it crystallized from volcanic material or montmorillonite during
burial under conditions where Mg concentrations and the Mg/K ratio were high.
Thin section studies of the Russian material (Rateev et al., 1980) indicate the Ch/S
formed by recrystallization of a volcanic groundmass after burial. In volcanic areas
permeated by hydrothermal fluids, Ch/S does not develop until temperatures of
approximately lOO"C, but it can form at much lower temperatures under evapora-
tive conditions (see Chapters VI and IX). During evaporation of sea water, the
activity coefficient ratio of Mg2+/K+ increases markedly. Mildly hypersaline condi-
tions could well have existed on the broad Ordovician tidal flats. The presence of
authigenic quartz indicates hypersaline conditions may have existed. Grimm (1962)
reports that idiomorphic quartz crystals occur only in salt or saline sediments. The
Ch/S could be detrital but it is unlikely that Ch/S could be the dominant physil
produced from a source rock by weathering for millions of years.
The distribution of the I/S and Ch/S suggests that in general the material from
which the illite and I/S formed originated in the shield areas where granitic rocks
and rhyolitic volcanics, rich in K and Si, were relatively abundant. The Ch/S and
chlorite are related to more basic volcanic material in island arcs and the mobile
581

continental margins. The compositional differences are related to the evolution of


the continental plates.
Deformation and metamorphism (Taconic Orogeny) of the eugeosyncline fring-
ing the eastern flank of Laurentia began in the Middle Ordovician. This was
followed by uplift and erosion in the Late Ordovician (Fig. 9-9). A large volume of
detrital sediments spread westward onto the former continental shelf. In the central
Appalachian region the source was to the southeast (Pettijohn et al., 1973). Shallow
water carbonates continued to be deposited on much of the eastern craton.
The physils in the carbonates and interbedded thin shales of the North American
Craton and fringing basins (St. Lawrence Lowland, Eastern Appalachian Basin,
Warrior Basin, Ouachita Mountains, Oklahoma, west Texas, Ohio, Illinois, Iowa,
Minnesota, and Montana) are largely illites. Most of the illites contain some
smectite layers. Many samples contain illite as the only physil but chlorite in
amounts of 5 to 20% is commonly present. The average chlorite content is on the
order of 5 to 10%. Chlorite content generally decreases from east to west.
Some of the marine Decorah and Glenwood shales of Minnesota (Parham, 1966)
and Middle Ordovician limestones in the Arbuckle Mountains (Oklahoma) contain
appreciable kaolinite. The kaolinite is presumably detrital and was derived from
weathering of the granitic basement rocks of the Canadian Sheld and the Trans-
continental Arch. Though the extent of the kaolinite is not known, its presence
indicates that, at least locally, the conditions existed that favored its formation. The
climatic conditions were appropriate - warm and humid (Frakes, 1979).
The volume of clastic rocks and physils increases from west to east towards the
eastern highlands; clastic rocks are more abundant in the Upper Ordovician than
the Middle Ordovician section. The eastern hghland source contained more basic
volcanics than the western source. This is reflected by a general eastward increase in
chlorite.
The physil suite of the Upper Ordovician Cincinnatian Series shales of Ohio,
Indiana, and northern Kentucky has an average composition of 77% illite (primarily
2M), 7% I/S and 16% chlorite (Scotford, 1965). There is little lateral or vertical
variation. The physil suite is that of the total shale. The < 2 p m fraction should
have less chlorite and more I/S. Cincinnatian limestones from Hamilton County,
Ohio, contain 94% illite and 6% chlorite, possibly with minor vermiculite (Booth and
Osborne, 1971).
Farther east (Pennsylvania, Virginia, and Georgia), well into the miogeosyncline,
the limestones have a low chlorite content, the shales commonly have 15 to 30%
(average 16% for 15 formations).
The physils in the sandstones of the Upper Ordovician Juniata-Bald Eagle
red-bed sequence of central Pennsylvania are largely illite and chlorite (10 to 60%,
-
average 251) with minor chlorite-vermiculite (Thompson, 1970). The red color is
due to the presence of hematite. Thompson suggested that in the drab colored
sandstones the hematite was reduced and combined with degraded physils to form a
diagenetic Fe-rich chlorite. The drab sandstones have more chlorite, and they have a
higher Fe/Mg ratio than the red sandstones. In equivalent age red-green calcareous
siltstones from north Georgia, Manley et al. (1975) found a similar pattern: the
582

chlorite in the red rocks had a lower Fe content. They concluded Mg, from the
carbonate, replaced chlorite Fe which formed hematite.
Chlorite appears to increase towards the east but no systematic study has been
made. The average chlorite content is similar to that in the Welsh basins; the latter
had a mixed source area composed of metamorphic and volcanic rocks. In the St.
Lawrence Lowlands of eastern Canada (Dean, 1962) chlorite is scarce in the Lower
and Middle Ordovician and increases in abundance in the Upper Ordovician.
In the northeastern portion of the Appalachian Basin (Hudson River Valley, New
York) the Middle Ordovician physilites show a systematic increase in illite crystal-
linity, ranging from h g h diagenesis to epizone, from west to east (Rutstein et al.,
1983). The sequence is similar to that in the southern Appalachian Basin in the
Upper Cambrian physilites (Fig. 9-2).
With the activation of a mobile orogenic belt along the eastern coast of Laurentia
in the Early Ordovician, a two source system developed that strongly influenced,
probably controlled, the make-up of the physil suites in North America. The acid
igneous rocks of the Canadian Shield and Transcontinental Arch formed a northern
and western source whch supplied illite and I/S to the craton and miogeosyncline.
Locally, kaolinite was supplied. During the Early Ordovician a basic volcanic source
(island arc) supplied mafic material to the craton and miogeosyncline; the material
altered to chloritic physils. As uplift continued in the eastern orogenic belt the
relative amount of basic igneous material decreased but the source continued to
supply a suite relatively rich in chloritic material. The presence of metamorphic
rocks in both source areas tended to homogenize the two physil suites. Environmen-
tal effects were secondary.

Europe, Africa

Upper Ordovician mudstones from mid-Wales are reported to contain approxi-


mately 38% chlorite (Fe-rich) and 62% illite (Evans and Adams, 1975). The high
chlorite content is apparently related to the abundance of nearby volcanics. Chlorite
is the only physil present in basaltic pumice-tuffs of Snowdonia, North Wales (Ball,
1966).0rdovician samples from the Pyrenees contain illite and 20 to 60% chlorite.
The K.I. indicates they have been subjected to low-grade metamorphism (Kubler,
1967). In East Germany the Upper Ordovician sediments contain up to 30%
kaolinite. The kaolinite is believed to have been eroded from kaolinitic weathering
crusts developed on various high areas during the warm humid conditions that
existed during the Lower Ordovician (Storr, 1975).
In the southern Baltic Sea region (Lashkov et a/., 1970) the Lower Ordovician
contains kaolinite and illite. Kaolinite decreases and chlorite increases upward and
the Upper Ordovician contains only illite and chlorite. The Middle Ordovician
contains all these physils plus smectite with illite predominating. The authors
believe the terrigenous material delivered to the basin during the Lower Ordovician
was not altered under the shallow water conditions but as physical-chemical
conditions of the basin changed some of this material was altered to smectite and
583

chlorite. More likely the mineralogical changes are due to changes in source material
similar to those observed in other areas. Chloritic material is abundant in the
Ordovician geosynclinal rocks of the southern Urals. Illite is the predominant physil
in the Ordovician rocks of the Russian Platform. Kaolinite is present in the
continental rocks (Vinogradov and Ronov, 1956).
In northwest Africa, the physil suite of the Ordovician rocks is composed
predominantly of illite and minor I/S; chlorite averages 25% and kaolinite 10%but
is present in amounts up to 40%. Kaolinite is absent in Ordovician samples where
the borehole temperature is higher than 150°C. The earliest Paleozoic glaciation
occurred during the Ordovician in the Sahara region. It is unlikely that kaolinite
would form under Arctic conditions; however, in order to develop glaciers a high
humidity is required. It is likely that the humid conditions that led to the develop-
ment of the glaciers also led to the development of kaolinitic weathering to the
north of the ice sheet.
Significant amounts of kaolinite are present in north-central United States and
northern Europe. These areas are in the vicinity of the suggested location of the
paleoequator. The presence of the kaolinite confirms that warm, humid conditions
existed in these areas. As the K.I. values of the illite indicate many of the
Ordovician rocks were exposed to burial temperatures in excess of 150°C, which
would have caused the destruction of kaolinite, it is difficult to determine how
widespread kaolinite was. Its presence in rocks from north Africa, presumably not
near the equator, suggests it may have been fairly widespread.
The Middle Ordovician, Black River and Trenton of North America and Llan-
virnian and Llandeilian of Europe, was a time of extensive explosive volcanic
activity. The volcanic activity is presumably related to Taconian (North America)
and Caledonian (Europe) orogenic activity. Ash beds were deposited over large
areas of Laurentia and Baltica (Fig. 9-10). For the most part the ash beds are now
composed of I/S containing 10 to 30% S . The beds are called K-bentonites. Some
contain chlorite (mixed dioctahedral-trioctahedralvariety) and Ch/S.

Fig. 9-10. Distribution of Middle Ordovician bentonites.


584

In eastern North America at least 19 separate ash beds (few centimeters to 4m


thick) have been identified in Middle Ordovician carbonate rocks. The maximum
reported aggregate thickness is 8 m in northeast Tennessee. At least one bed is
present in the Lower Upper Ordovician (Martinsburg Shale of Tennessee). They
have been found in an area ranging from Quebec (Brun and Chagnon, 1979) to
Alabama (Nelson, 1922; Kay, 1935; Weaver, 1953) and occur as far west as Iowa
and Minnesota (Mossler and Hayes, 1966). Discrete ash beds do not appear to have
been preserved in the south-central region; however, the Middle Ordovician shales
and limestones of Oklahoma and west Texas have a high content of I/S.
Twenty-four K-bentonite beds have been found in the Middle Ordovician
sediments of Scandinavia (Hagemann and Spjeldned, 1955; Bystrom, 1956). Similar
beds are abundant in the same age sediments throughout the general Baltic region
(Rateyev and Gradusov, 1970). Thicknesses range from less than a centimeter to 2
meters. Mineralogic studies by the latter authors and Bystrom (1956) show that
these clays are similar to those in the United States though there is a wider range in
the illite-montmorillonite ratio. Ratios range from 3:7 to 8:2. Eleven thin (1-3 cm)
K-bentonite beds are present in the Lower Cincinnatian flagstones of North Wales
(Ball, 1968).
Whereas much of the chloritic material in the Ordovician was derived from basic
volcanics, the K-bentonites were derived from explosive acid volcanic material. The
volcanism was presumably related to orogenic activity (Taconian-Caldonian) along
both flanks of the Iapetus Ocean (seaward of the ash deposits). Where the ash fell
into marine waters it apparently altered to montmorillonite in a manner similar to
that of younger ash beds and later altered to I/S during burial diagenesis. As noted
in the review of DSDP data (p. 374), the ash may not have altered until millions of
years after deposiiton.
In general the illite content of the I/S, particularly for the thm beds, increases
with increasing depth of burial. Thus, in North America the I/S ratio generally
increases from - 3:2 in the west (shelf) to - 4:l in the east (geosyncline), with
numerous exceptions; in Baltica the relation is the same except the ratio increases
from east to west. In the Oslo region the ash has been converted completely to illite
and chlorite (Bystrom, 1956).
The distribution of chlorite is related to the I/S ratio. In general the more deeply
buried K-bentonites, with the higher I/S ratio, contain more chlorite. Thus, chlorite
is relatively abundant in the K-bentonites of the Appalachian region and is scarce or
absent in the mid-continent.
Superimposed on t h s general trend are numerous anomalies. Adjacent beds have
different I/S ratio. In Sweden thin beds have an I/S ratio of 7:3 and an adjacent
thick bed (2 m) has a ratio of 4:6. In Pennsylvania adjacent t h n beds have the same
I/S ratios, but some have chlorite and others do not. The K-bentonites on the
eastern flank of the Cincinnati Arch (Kentucky and Tennessee) have a high I/S
ratio even though they have not been deeply buried (Huff and Turkmenoglee, 1981).
There are three obvious factors that influence the composition of the K-bentonite
beds: differences in bed thickness, differences in composition of the volcanic ash,
differences in the composition of the enclosing rocks.
585

The K-Ar apparent age of a Swedish Ordovician K-bentonite sample is 25%


lower than the true age (Bystrom-Asklund et al., 1961). I obtained a similar value
for a sample from Pennsylvania. These data confirm that most of the K was
incorporated into the 2 : l layers after a considerable depth of burial.
The thin beds commonly have a higher I/S ratio than the thicker beds. Further,
some thick beds are zoned. The K and A1 content and the proportion of illite layers
increase from the center of the bed towards the edges. Thus, it appears that K and
perhaps A1 in the surrounding sediments have diffused into the bentonite beds
(Velde and Brusewitz, 1982). Possibly there has been some outward diffusion of Si
rather than an inward diffusion of Al. Where the K-bentonites occur between
limestone beds the K must have traveled long distrances parallel to the bedding or
else the K was derived from the original ash.
The chlorite that is present in the more highly metamorphosed samples ap-
parently is a by-product of “cleaning-up” the I/S. The fact that chlorite is present
in some K-bentonite beds and not adjacent beds suggests the original ash beds had
slightly different compositions or that there was a difference in the Mg content of
the adjacent rocks. This would indicate there was diffusion of Mg as well as K.

SILURIAN

North America
The eastern highlands along the flank of North America continued to be a major
source of detritus. The orogenic activity that began in the Ordovician spread
progressively northward into the Arctic (Franklin) and northern Greenland during
the Silurian. Volcanism was important in New England and Maritime Canada.
Sandstones and shales are the predominant rocks in the Appalachian Basin. Oolitic
iron beds are present in the southern part of the Basin. These grade westward into
carbonate rocks which contain an abundance of reefs. Carbonates are the predomi-
nant rocks in West Texas and the Williston Basin. In the late Silurian salt was
deposited in the Michigan Basin, northern Ohio Basin and northern Appalachian
Basin.
Illite is the dominant physil in nearly all the Silurian rock x-rayed. Varying
amounts of chlorite and kaolinite are present. Mixed-layer I/S as a discrete major
phase is scarce but illite commonly contains 5 to 10% smectite layers.
In the Appalachian Basin the basal Silurian Tuscarora deltaic sandstones of West
Virginia and Virginia contain up to 50% kaolinite. It is not known if the kaolinite is
detrital or diagenetic. To the north, in Pennsylvania, the thin physilite beds in the
Tuscarora contain mostly illite and commonly less than 5% chlorite. To the south, in
Georgia, parts of the Red Mountain Sandstone contain 10 to 20% kaolinite, other
parts contain 10 to 20% chlorite and no kaolinite. Equivalent age limestone
(Brassfield) in Ohio are composed predominantly of illite, with a significant amount
of I/S and minor amounts of chlorite and kaolinite. Kaolinite increases to the
southeast, towards the source of clastics (Ehlers and Hoover, 1961). Halloysite is
586

present in Brassfield samples from central Tennessee.


The lower Silurian coarse clastic wedge is overlain by a series of limestones and
shales indicating the eastern orogenic belt had become inactive. Rocks of the Middle
Silurian Clinton Group (shallow marine) occur in a belt extending from New York
to Alabama. They are characterized by the presence of thin but extensive beds of
oolitic ironstone. Illite is abundant but the oolites are composed largely of primary
chamosite partially replaced by diagenetic chlorite (Schoen, 1964). Illite and varying
amounts of chlorite are the only physils found in any abundance in the other
Silurian formations in the Appalachian Basin and farther west.
The Upper Silurian Wills Creek shales (largely lagoonal) of eastern West Virginia
contain from 14 to 63% chlorite (average 16%). The samples with a high chlorite
content are presumably from thin evaporitic units (Folk, personal communication).
K.I. values for the illite ranges from 2.5 to 3.0, lower anchizone, thus some of the
chlorite may have formed during low-grade metamorphism. In Pennsylvania and
New York equivalent age shales contain an average of only 7% chlorite; illite K.I.
values are commonly in the range of 3.0 to 4.5 (diagenesis) and mostly of the 1 Md
variety. It is not known whether the difference in chlorite content in the two areas is
due to the fact that they have been exposed to different degrees of metamorphism,
different enviornments, or different source rocks. Environment is a likely choice.
Farther west, along the eastern flank of the Cincinnati Arch (east-central
Kentucky) the shallow buried, Middle Silurian Crab Orchard shales are composed
predominantly of illite, apparently with appreciable 9 : l I/S. The chlorite content
averages approximately 3%. The northeasternmost samples contain an average of
14% kaolinite and no chlorite (Scott, 1973). As in the Lower Silurian the kaolinite
content increases towards the eastern highlands, though relatively little is present in
the eastern miogeosynclinal sediments. The pattern suggests that appreciable
kaolinite was probably present in sediments of the eastern Appalachian basin and
was destroyed during burial metamorphism.
Samples from the eastern part of the craton indicate the physils are largely illite
with minor chlorite. Silurian rocks are absent from most of the midcontinent region
but are present, primarily as carbonates, in west Texas and the Williston Basin of
Montana and Alberta. Illite, diagenetic grade, is the only physil present in over 95%
of the samples studied from these two carbonate provinces. The other 5% of samples
contain minor chlorite.
The Silurian pattern is similar to that in the Upper and Middle Ordivician. Illite,
with significant amounts of chlorite and some kaolinite, is present in the geosyncline
near the eastern highlands. Chlorite and kaolinite decrease from east to west onto
the craton area, with illite being essentially the only physil on the western flank of
the craton.
Again the question arises, why exclusively illite in the western sediments? This
may be due, in part, to the fact that the rocks are primarily carbonates. Paleozoic
carbonate rocks characteristically have a high illite content; however, carbonate
rock can have an abundance of any of the physils. It is unlikely that detrital physils
are destroyed in carbonate enviornments, unless it occurs during burial diagenesis.
In general carbonates are deposited at a considerable distance from mountainous
587

source areas which usually supply a relatively heterogeneous physil detritus. The
near-exclusive illite suites could be created during burial, high-grade diageneses or
anchimetamorphism. Kaolinite would be destroyed and smectite or I/S converted
to illite; however, chlorite should develop as a by-product of the reaction. It is
possible that the Mg and Fe released from the smectite could be incorporated in the
carbonate minerals rather than used to grow chlorite. It is also possible that
illite-I/S was neoformed in partially enclosed, shallow marine, relatively K-rich
(alkaline source rocks) environments.
The major physils in the Upper Silurian halite deposits of New York are illite
and chlorite. Some samples contain talc and serpentine (Boding and Standaert,
1977). The illite is well crystallized but has considerable compositional variation
suggesting it is recrystallized detrital illite. The chlorite is a hgh Mg variety with a
relatively uniform composition indicating it probably crystallized in the halite. Illite,
with minor I/S and chlorite is the predominant physil in the Michigan Basin halite
deposits; several varieties are present in most samples. Interbedded dolomites have
a similar physil suite. Mixed-layer vermiculite-montmorilloniteis present in a few
samples (Droste, 1963). The illite-I/S is presumably detrital and the Ch/S and
perhaps the V/S, authigenic or diagenetic. Most, if not all, of the physils formed in
evaporite deposits are Mg-rich. It is not clear whether the differences between the
New York and Michigan deposits are due to differences in burial temperature or in
water chemistry.

Europe, Africa

Scattered data show that the Silurian rocks of Norway contain primarily illite
and chlorite. Chlorite is relatively abundant, commonly comprising 25 to 40%of the
physil suite (Hagemann, 1957; Bjorlykke, 1974; Turner, 1974). Turner (1974) found
that at least part of the hematite in the fluviatile Ringerike Group red beds was
derived from Fe leached from chlorite after burial. In the southern Baltic Sea region
the physil suites are also primarily illite and chlorite but some contain kaolinite and
montmorillonite. The chlorite and montmorillonite is believed to have formed in the
depositional basin (Lashkov et al., 1970).
Silurian K-bentonite beds are present in Norway, Sweden, Estonia, Lithuania
and Poland (Snall, 1977). The distribution is similar to that of the Middle Ordovi-
cian K-bentonites and suggests renewed volcanic activity in the same orogenic belts.
Snall suggest the sources may be the Alno district of Sweden, the Scandinavian
Caledonides and the western border of the Ukranian shield. Over 200 layers
commonly less than 5 cm thick, are present in Gotland, Sweden. The Silurian
K-bentonites commonly have more expanded layers (35-42%) than the Ordovician
K-bentonites. Biotite is abundant in many beds. Much of it has been altered to
chlorite or kaolinite. Detrital kaolinite is present in many beds, indicating some of
the volcanic material was first deposited on land, weathered, and then transported
to the sea. The land source was believed to be the geosynclinal uplift of Poland
(Ratyev and Gradusov, 1970).
588

In Baltica, the Caledonian Orogeny is believed to have been active throughout


much of Ordovician and Silurian time (Ager, 1980), accounting for the abundance
of K-bentonite deposits. The apparent absence of Silurian K-bentonites in eastern
Laurentia suggests orogenic activity, or at least volcanic activity, was less than in
Baltica. However, submarine volcanic rocks are present in eastern New England and
Maritime Canada. Undetected Silurian K-bentonites are likely present in northeast-
ern North America.
The physils in the shales and limestones of Wales are largely illite with a lesser
amount of chlorite; minor kaolinite is present in most samples. I/S is the dominant
physil in clay beds from the Welsh Border (Perrin, 1971). These beds are apparently
volcanic ash deposits.
Illite is the predominant clay in the Lower Paleozoic sediments of the Russian
Platform. Kaolinite is present in the continental facies (Vinogradov and Renov,
1856). Eastward in the southern Urals the story is more complex (Rateev et (11..
1980). At the end of the Cambrian, faulting and separation of the Russian - Siberian
platform began. By the Silurian an open ocean (geosyncline) had developed. Sialic
clastics were derived from the Russian platform to the northwest. To the southeast,
island volcanoes arose. The physil suite in the Lower Silurian rocks in the western
part of the basin is composed of 80% illite and 20% Ch/S and chlorite. Farther east
the content of illite decreases to 50 to 60% and the amount of Ch/S increases. The
Ch/S developed, after burial, from volcanic clastics derived from the eastern
volcanoes. The illite was believed to have been derived from the Russian Platform to
the northwest. The situation is similar to that in eastern North America, though
metamorphism was apparently more intense in North America and much of the
volcanic material was converted to chlorite. The physil suite of the remainder o f the
Silurian in the western portion of the geosyncline is composed exclusively of illite
(apparently I/S, 9:l). As in North America, the source of the illite is not estab-
lished. There is no evidence of volcanic material. In the eastern portion of the
geosyncline, where volcanic material is abundant the physil suite contains only 40 to
50% illite. The remaining material is Ch/S derived from volcanic material.
The physil suite of the shales and sandstone of the Polignac Basin of northwest
Africa is composed of approximately 40-50% illite and approximately equal parts
I/S, kaolinite and chlorite, as are most of the Paleozoic rocks in the basin. Kaolinite
is present in amounts as high as 70% (Dunoyer de Segonzoc, 1969).
Kaolinite is apparently more abundant in Silurian age rocks than in rocks of the
preceding periods. Kaolinite is relatively abundant in northwestern Africa, Baltica
- Scandanavia and to a lesser extent in eastern North America. The presence of

reefs in these areas indicate the climate was warm. Land plants, in any abundance,
first appeared in the Silurian; however large forests did not develop until the
Devonian. It is possible that the increase in kaolinite was related to the spread of
land plants and the development of acid soils.
589

DEVONIAN

North America

In eastern North America, at the beginning of the Devonian, the highlands raised
in the Taconic orogeny were submerged or of low relief. Most of the northern
Appalachian Early Devonian geosynclinal sediments are limestones, shaly limes-
tones, and siltstones. During the Middle and Late Devonian Baltica is believed to
have collided with North America. Whether or not a collision occurred is in doubt,
but there was a period of mountain building. The collision and/or mountain
building (Acadian Orogeny) was restricted to northeastern North America (Acadia)
and had little effect on the central and most of the southern Appalachians. The
northern highlands were the source of the sediments that comprise the Catskill
clastic wedge which extends from eastern New York and Pennsylvania into Michi-
gan.
The Lower Devonian Helderberg thin bedded limestones and shales of New York
are composed primarily of 2M illite and 5 to 10% chlorite (Borst, 1966). The physil
suite in Pennsylvania is similar except appreciable Ch/S is present in some
limestone samples.
As in the Ordovician, the occurrence of volcanic ash beds heralded orogenic
activity (Acadian) in the eastern mobile belt. Mountain building extended from the
Middle Devonian through Early Mississippian. The Catskill clastic wedge was
deposited with sediments becoming coarser, more red, and more continental from
west to east. The eastern source rocks were largely only slightly older Paleozoic
rocks and intruding granitic plutons.
Tuff and ash beds are thickest and most abundant in north-central Virginia,
where 32 tuff beds have been identified (Dennison, 1983). At least three ash beds
(Center Hill, Belpre, and Tioga), altered to K-bentonite, occur throughout much of
the Appalachian Basin (Roen and Hosterman, 1982) and one, Tioga bed, is present
in the Illinois Basin (Droste and Vitaliano, 1973) and the Michigan Basin (Baltrusai-
tis, 1974). The beds commonly range from 0.5 to 30 cm thick. Biotite is characteristi-
cally present. Six beds, some up to 70 cm thick, are present in limestones of the
Gasp6 peninsula, Quebec, Canada (Smith, 1967).
The major physil is an I/S with 10 to 20% smectite layers (R > 1). Ten to twenty
percent kaolinite is present in most samples and many contain discrete illite. The
illite is probably a detrital component. The kaolinite apparently formed from the
feldspar or biotite in the ash. The I/S presumably formed by burial diagenesis.
Rb/Sr isotopic studies show that the I/S is about 70 m.y. younger than the
sedimentation age (Fullagar and Bottino, 1969).
In the Middle Devonian (Ludlowville) of central New York the chlorite content
of the eastern continental red bed facies averages approximately 10% to 15%.
Westward for 200 km, in the marine shales, the chlorite content remains relatively
constant at 10%.The only other physil present is illite (Towe and Grim, 1963). The
sharpness ratio of the illite peak systematically increases from west to east. The
590

change in peak width is probably due to burial diagenesis rather than environmen-
tally induced changes. The conodont color alteration index (CAI) increases from 4
(60-140°C) in the western portion of the area to 4 (190-300°C) in the eastern area.
Thus, burial depth presumably accounts for lateral changes in illite crystallinity.
Kaolinite could have been present in the eastern area, but the high temperatures
would have destroyed it. The Middle Devonian thin shale beds of Ohio (Droste and
Vitaliano, 1973) and Pennsylvania are composed primarily of illite, occasionally
with minor chlorite.
In Arkansas, Oklahoma, and Texas much of the Devonian is represented by a
chert facies (Arkansas and Caballos novaculite) and interbedded thin limestones
and shales. Illite is the only physil in most samples; kaolinite is present in minor
amounts in a few samples.
Illite comprises 90% of the physils in the Early and Middle Devonian rocks.
During the Early Devonian minor uplift, domes and arches were formed by mild
orogenic activity and a craton-wide unconformity developed. Much of the illite may
have been derived from the earlier Paleozoic sediments exposed in the cratonic
highs.
During the Upper Devonian a thck wedge of continental clastic sediments
(Catskill Formation) was deposited in eastern New York and Pennsylvania. This
tectonic delta complex contains a high content of coarse red sandstones and gravels
that were deposited marginal to a mountain front. A sequence of sandstones and
siltstones from the Catskill Mountains of New York (Friend, 1966) have a physil
suite composed of 85 to 90% illite and 10 to 15% chlorite plus kaolinite (recalculated
from 7 A/lOA values). The chlorite plus kaolinite is more abundant in the
sandstones than in the siltstones. The published values indicate this increase is
primarily due to an increase in kaolinite. This is particularly true of the red
sandstones, as opposed to the non-red sandstones. The chlorite/chlorite plus
kaolinite ratio for the non-red sandstones is greater than 0.6 and for red sandstones
less than 0.6. Though Friend concluded the red color, hematite, is unrelated to the
physils, the kaolinite-chlorite distribution could be interpreted to indicate there was
some post-depositional alteration of chlorite and/or mica to kaolinite, releasing Fe
to form hematite.
In a study of Catslull rocks from the same general area Liebling and Scherp
(1976) reported chlorite values ranging from 4 to 40% (recalculated). The chlorite
content was highest and had a higher Fe content in fine-grained sediments de-
posited in oxbow-lake and floodplain environments. Chlorite was less abundant and
more Mg-rich in the pointbar sands. They suggest the Fe chlorite in the sandstones
was destroyed, producing primarily hematite and silica. They further suggest that
the illite-chlorite content of the fine-grained sediments reflects the composition of
the original detrital suite. In central Pennsylvania the Catskill silty shales contain 0
to 20% chlorite. The chlorite content of the finer grained, more marine facies
(Chemung) ranges from 10 to 30%.
Westward and southwestward from the Catskill Delta a thin blanket of marine
organic-rich (up to 20%) and pyrite-rich (up to 10%) black shales was deposited.
These shales range in age from Middle Devonian to Lower Mississippian but most
591

Table 9-1
Average physil composition of Middle and Upper Devonian shales in the Appalachian Basin. After
Hosterman and Whitlow, 1983.
Age Location lllite l/S Chlorite Kaolinite I/Ch
Devonian
Late Upper E. Ohio 58 21 12 4 Tr
N.E. Kentucky
W.W. Virginia
Early Upper W. New York 58 20 22 Tr Tr
N.W. Pennsylvania
Middle C. Pennsylvania 65 15 20 Tr Tr
C. West Virginia
W. Virginia

of the shale is Late Devonian. The sediments were deposited in a broad epicontinen-
tal sea and covered much of the eastern United States (Chattanooga and New
Albany) and extend as far west as Texas (Woodford) and northwest into Canada.
In the eastern United States the shale decreases in thickness and increases in
organic carbon content from east (Virginia) to west (Illinois). The easternmost area
has a relatively high content of non-marine sediments. The source was to the east
and northeast. Analyses (Hosterman and Witlow, 1983) of more than 2,000 shale
samples from wells in the northern portion (Tennessee to central New York) of the
Appalachian Basin indicates the shales contain 50 to 75% physils, most 60 to 7556,
and 20 to 30% quartz. Minor pyrite and calcite are present in most samples. The
average physil composition of three age units is shown in Table 9-1.
In general the section is thinner and becomes younger from east to west. Illite
(2M) and chlorite are more abundant and I/S less abundant in the older shales
(from west to east); in addition, the sharpness of the 10 A illite peak increases to the
east. The conodont color alteration index (CAI) (Epstein et a]., 1976) suggests burial
temperatures increased from 50" to 140°C in the western part of the basin to 110"
to 200°C in the more deeply buried eastern part of the basin. Hosterman and
Whitlow believe the development of 2M illite and chlorite is due to diagenesis and
metamorphism of degraded illite and montmorillonite. The temperatures do not
appear to have been high enough for 2M illite to develop.
Analysis of 110 samples from three wells in the Appalachian Basin and two wells
farther west in the Illinois Basin are shown in Table 9-2 (Weaver). The I/S contains
approximately 5 to 10% smectite layers. The difference between my values and those
of Hosterman and Whitlow are due to difference in sample preparation and slightly
different correction factors.
Though there is a considerable range of values for individual samples the average
values for each well are essentially identical. The illite is mostly the 2M polymorph.
Most of the K.I. values average between 1.3 to 1.8 (high anchizone to low epizone)
which would indicate that they were exposed to burial temperatures in excess of
300°C. In the underlying Middle and Lower Devonian shales the K.I. values are in
592

Table 9-2
Composition of Upper Devonian Black Shales.
Location Mite I/S Chlorite Kaolini te
Virginia 73 19 8 0
Kentucky 68 22 7 3
Ohio 68 20 11 1
Indiana 70 20 10 0
Illinois 74 12 14 0

Range 50-90 0-42 3-26 0-14

the range of 3 to 4 (lower temperature). The average W.I. values, which take into
account small amounts of I/S material (shoulder on low angle side of 10 peak), A
range from 3.3 to 5 and generally increase to the east, reflecting the increasing depth
of burial. However, even though the W.I. values indicate a lower burial temperature
(lower anchizone) than the K.I. values, the indicated temperatures are higher (250 to
300°C) than those suggested by the CAI index ( < 50 to 200°C). Estimated burial
depth also indicates it is unlikely these physilites have ever been exposed to the high
temperatures, - 3OO0C, indicated by the K.I. and W.I. indices.
The sharpness of the 10 A peak, for such a shallow depth of burial, indicates the
illite and probably the I/S are detrital. For most of the Paleozoic section in this
area, the chlorite content increases from west to east, so it is likely that the eastward
increase in chlorite in the Devonian black shales reflect the same trend. However, as
Hosterman and Whitlow suggest, it is likely that in the easternmost area some I/S
has been converted to illite and minor chlorite.
The black shales of Texas, Woodford, Caballow, Lower Tesnus, and Barnett,
have a physil suite similar to the eastern shales, primarily illite, with minor I/S and
chlorite. The Besa River black shales of British Columbia are composed primarily of
illite and approximately 10% chlorite (Pelzer, 1966). Analysis of a few black shales
from Alaska show they are similar to the others, containing primarily illite and trace
amounts of chlorite.
I t is of interest to note the effect of organic material on the peak intensity of the
A A
10 x-ray peak. The intensity of the 10 peak decreases linearly with increasing
organic content though the amount of illite in the physil suite appears to remain
A
constant. When the black shales contain 5% organic carbon the 10 peak is only
1/3 as high as when 0.5% carbon is present. Adsorption of x-rays by organic
material and pyrite presumably accounts for the decrease in intensity.
Fig. 9-12 contains typical x-ray patterns of a number of Paleozoic black shales.
Most are from the Upper Devonian-Lower Mississippian. The high 5 A peak of the
Tesnus and Barnett samples is due to the relative abundance of I/S. The illites in
the organic rich black shales commonly have a 10 A/5 Apeak-height ratio near 4.
This is a relatively high value considering the sharpness of the 10 peak and A
suggests the illites are relatively Fe-rich. Some Fe may have entered the octahedral
layer under the anoxic conditions that existed in the depositional basins.
593

DEVONIAN

Fig. 9-11. Devonian paleogeography. From Mlntz, 1981.

The general absence of kaolinite in Devonian black shales is surprising. The


humid conditions necessary for the formation of abundant land plants normally
results in the formation of abundant kaolinite. The lack of kaolinite suggests much
of the organic material is marine rather than continental. However, the absence of
significant kaolinite in marine black shales of all ages suggests it may not be stable
in organic-rich marine environments.
The physils in the Upper Devonian Ouray Limestone of southwestern Colorado
(San Juan Basin) and Utah (Paradax Basin) are predominantly illite. Farther north
in southwestern Montana the physil suites contain 10 to 30%chlorite and 10% I/S
(Leckie, 1962). The normal marine shales of the Williston Basin, Montana, and the
Central Alberta Basin, Canada, contain up to 70% chlorite but typical values are 10
594

to 20%. In the Northwest Territory, Canada, the Melville Island Group physils are
predominantly kaolinite and IMd illite. Environments range from marine to non-
marine (Cauffman and Bayliss, 1973). Samples of the Upper Devonian evaporite
potash beds of western Canada contain either an illite, chlorite, and random
mixed-layer vermiculite/montmorillonite physil suite or corrensite with lesser
amounts of illite and chlorite (Droste, 1963). The evaporite physils and kaolinite in
the Melville Island Group suggest a warm climate existed in northwest Canada. The
area was apparently near the equator (Fig. 9-11).

Europe, Africa, South America

During the Devonian the Caledonian Orogeny continued along the northwest
margin of Europe. T h c k deposits of red alluvial conglomerates and sandstones were
deposited in Ireland and Scotland. The deposits, becoming finer to the east, covered
much of Great Britain, Scandinavia, and northwest Europe. The unit is called the
Old Red Sandstone and is essentially a mirror image of the Catskill red beds of
North America, except the former was deposited throughout much of the Devonian.
The major difference between the two deposits is the presence of abundant andesitic
lavas in the Old Red Sandstones. The presence of the volcanic rocks is reflected in
the physil suite.
The physils of the Old Red Sandstone of Scotland have been studied in some
detail by Wilson (1971). The rocks contain a wide variety of physils. An examina-
tion of the weathered volcanics showed they altered to Ch/V, saponite, I/S,
chlorite, illite, and kaolinite. The physils in the lower part of the Old Red Sandstone
(Downtonian) are primarily kaolinite and illite with minor amounts of I/S and
Ch/V. Montmorillonite is dominant in some areas (Perrin, 1971). These physils are
believed to have been derived primarily from weathered volcanics. The overlying
sediments (Dittonian) are composed primarily of Ch/V and Ch/S (dioctahedral)
with varying amounts of montmorillonite, illite, chlorite, kaolinite, and I/S. This
material was also presumably derived from volcanic material but under less intense
weathering conditions and more rapid erosion. Approximately 12,000 m of coarse
material was deposited in a short period of time.
The physils in the Middle Old Red Sandstones are predominantly well crystal-
lized 2M illite and chlorite. Minor amounts of montmorillonite and Ch/S are
present. Samples from some areas contain abundant kaolinite and variable amounts
of I/S. These physils are apparently detrital, but the source rocks contained fewer
volcanics. The physils in the Upper Old Red Sandstone are predominantly kaolinite
and I/S (15 to 30% smectite). Some of these physils are diagenetic, but most appear
to be detrital. Some of the kaolinite formed by alteration of micas. Weathering
intensity apparently increased during the Upper Devonian.
The red color is due to the presence of hematite which apparently formed after
deposition.
Analyses of a few marine samples show they are composed predominantly of
illite and minor chlorite; the composition is typical of those found in most of the
595

CHATTANOOGA-TENN TESNUS-TEXAS

WOODFORD-TEXAS BARNETT-TEXAS

CABALLOS-TEXAS HEEBNEIt-KANSAS

LODGPOLE-MONTANA POLK CREEK-OKLAHOMA

Fig. 9-12. X-ray patterns of Paleozoic marine black shales. Illite is the major physil in all samples.

Lower Paleozoic marine sediments. Either these few samples are atypical or there
have been major diagenetic changes in the marine sediments.
The complex physil suite of the Old Red Sandstone contrasts sharply with the
simple illite-chlorite suite of the Catskill Formation. The abundance of Ch/S,
Ch/V, I/S, saponite, and montmorillonite reflects the abundance of volcanic
material associated with the Old Red Sandstone. The relatively high kaolinite
content of the Old Red Sandstone indicates weathering conditions were more
intense, possibly indicating prevailing winds were from east to west with more rain
falling on the eastern slopes of the Caledonian Mountains. The difference in the
596

physil suites also suggests the two red bed sequences did not have a common source.
Sediments from the western flanks of the Caledonian Mountains were apparently
deposited in Greenland.
Kaolinite (40 to 60%), along with illite (hydromica), is abundant in the Middle
and Upper Devonian nearshore sediments of the Russian platform. Much of the
kaolinite was derived by weathering of hydromica. With increasing distance from
shore the proportion of illite increases and kaolinite decreases. Some chlorite and
smectite is present in the marine sediments (Rateev, 1964). A similar sequence
occurs in equivalent age rocks of the Kama-Vyatka Trough. The Middle Devonian
littoral deposits contain kaolinite and illite. Mite and chlorite increase upward as the
section becomes more marine. The kaolinite was believed to have been derived from
a kaolinite weathering crust of the Tartarian Arch (Vlodarskoya, 1962). These
kaolinite-rich sediments are located in the vicinity of the Devonian paleoequator
(Fig. 9-11).
Farther west, in the Ural eugeosyncline a variety of andesitic volcanic types are
abundant, as in the older Paleozoic rocks, in the continental “Tuffitic-siliceous
formation”. The physils are ordered I/S (7:3) and Ch/S which probably formed
during burial (Rateev et al., 1980).
The Ardennes. Belgium, are situated just south of the border between the Old
Red continent, the stable part of Europe, and the Devonian geosyncline to the
south. The sediments are primarily marine. The physils, which have been subjected
to burial metamorphism, are composed primarily of illite with minor chlorite and a
variety of other physils. Metamorphism, based on K.I., increases from diagenesis
grade in the north to epizone grade to the south. Minor amounts of Ch/V are
present in the anchizone samples and paragonite and mixed-layer
paragonite/phengite in the epizone (Dandois, 1981).
Illite, with minor chlorite, is the main physil in the Devonian rocks of Hungary:
however, Ch/V, Ch/V/M, and kaolinite are also present.
The contrast between the Devonian physils of Europe and North America
demonstrates, more so than in older sediments, that source material and climate
determine the nature of the physil suite. At least during Devonian time there was no
global factor that controlled the formation of physils. The physils indicate that in
Europe volcanic rocks (Ch/S, Ch/V, I/S) were more abundant than in North
America. The relative abundance of kaolinite in the European sediments indicates
weathering was more intense, presumably indicating the climate was more humid.
The Devonian marine shales and limestones in the Polignac Basin of northwest
Africa contain varying amounts of illite, I/S, chlorite, and kaolinite. Illite plus I/S
comprises only 25 to 50% of the physil suite. Burial depth is relatively shallow, and
the physils have only been exposed to diagenetic grade metamorphism. Kaolinite is
abundant, comprising as much as 70% of the physil suite. It increases in abundance
from south to north, presumably in the direction of the source area. Some sections
contain 40 to 50% chlorite. Chlorite increases to the southeast (Dunoyer de
Segonzac, 1969). The presence of abundant kaolinite and easily weathered chlorite
suggests either there was more than one source area or the kaolinite formed during
an early stage of weathering.
591

To the west, in the Tindouf Basin, pyrophyllite, up to 25% of the physil suite, is a
common component of the marine Devonian and Silurian sedimentary rocks (zone
of diagenesis). The pyrophyllite occurs in the deepest part of the basin and is
associated with carbonate rocks, halite, and gypsum. It is believed to have been
formed from mica, kaolinite, and quartz by the action of interstitial solutions after
burial (Chennaux and Dunoyer de Segonzac, 1968). They note that pyrophyllite is
present in equivalent age rocks in western Europe; however, most of the European
pyrophyllite appears to have a metamorphic origin.
Illite, I/S (1:l regular mixed-layer), and chlorite each comprise 20 to 30% of the
physil suite. Near the top of the Upper Devonian section pyrophyllite decreases and
kaolinite appears in amounts of 30 to 50%.The vertical decrease in pyrophyllite and
increase in kaolinite suggest the former was formed from the latter and at a
relatively low temperature. The source area was to the north and northwest
(Cavaroc et al., 1976). The source area was presumably close to the North American
continent, but there are no equivalent physil suites in North America, suggesting an
ocean barrier existed between the two continents.
Most of the Siluro-Devonian fluvial deltaic physilites of South Africa have been
subjected to anchimetamorphism. K.I. values decrease from north to south as the
depth of burial increases. The physil suite is composed of illite and chlorite. Most
samples contain between 30 and 60% chlorite, which is more than most Paleozoic
shales contain. In part the hgh chlorite is due to the relatively advanced stage of
metamorphism, but probably also indicates there was considerable basic volcanic
material in the source area. Porphyroblasts of chlorite, muscovite, and intergrowths
of the two are very common. The physilites are slates (Rowsell and De Swardt,
1976).
The Parano Basin of southern Brazil, Uruguay, Paraguay, and Argentina has a
well developed Devonian section. In Brazil the physil suite of the Lower Devonian
littoral to shallow sandstones is composed predominantly of illite and varying
amounts of I/S, chlorite, and kaolinite. About half the samples contain kaolinite.
Pyrophyllite is present in a few samples. The illite is well crystallized (low K.I., low
10 A/5 A ratio) suggesting most of the illite is detrital. The Upper Devonian dark
marine shales are also composed predominantly of illite with subordinate amounts
of chlorite and I/S (Ramos and Formoso, 1976). To the southeast in Uruguay the
physil suite of the marine sediments contains 70 to 90% kaolinite (Elizalde and
Steinberg, 1973).
The scattered data indicate kaolinite is relatively common in the Devonian rocks
of Europe, North Africa, and southern South America and rare in North America.
It is not clear from the various paleoclimatic reconstructions what the kaolinite
areas have in common. The various geographies proposed for the plates during
Devonian time show the plates containing kaolinite all lie east of the North
American plate (Fig. 9-11). The presence of the kaolinite suggests the climate was
more humid and perhaps warmer over the eastern plates.
598

CARBONIFEROUS

North America

Mississippian
The Early Mississippian eperogenic conditions were the prelude to the major
tectonic and sedimentological changes which occurred in the Late Mississippian and
Pennsylvanian when Africa and South America (Gondwana) collided with North
America and Europe (Laurussia) (Fig. 9-13). During the Lower and Middle Missis-

CARBONIFEROUS

Fig. 9-13. Carboniferous paleogeography. From Mintz, 1981


599

Fig. 9-14. Paleogeographic and lithofacies map of conterminous United States during the Mississippian.
Data unavailable in blank areas. From Gutschick and Sandberg (1983). Copyright 1983 SOC.Econ. Paleo.
Miner.

sippian epeiric seas covered much the same areas of the North American craton as
did the Upper Devonian seas. Conditions were less anoxic and the sediments were
largely carbonates rather than organic shales. The influence of the approaching
plate collisions is first detected in the Upper Mississippian Chester sediments. Fig.
9-14 shows the paleogeography of the United States during early Mississippian time.
The northeastern highlands were still in existence and relatively coarse alluvial
sandstones (Pocono Formation) were deposited in the Appalachian Basin. These
grade laterally and vertically into the MauChunk shale. The physils are predomi-
nantly illite, 5 to 15% chlorite and 5 to 15% kaolinite. Only a few samples were
analyzed.
The Lower Mississippian MacCrady Formation of southwest Virginia and West
Virginia (Blancher, 1973) consists of shallow marine to tidal carbonates and shales.
Illite is predominant; kaolinite is present in minor amounts; chlorite and
montmorillonite(?) are present in the red siltstones (supratidal) and vermiculite
(dioctahedral) and corrensite in the green siltstones (intertidal to subtidal). Evaporite
deposits are present in the same formation (Nelson, 1973). The detrital physils in
the clastic rocks are primarily illite (I/S, 9 : l ) and minor chlorite. The interbedded
evaporite rocks (halite and gypsum) contain well crystallized (sharp peaks) illite and
Mg-chlorite ( - to 25%). The detrital illite is believed to have been altered to chlorite
in the high Mg evaporitic waters. The K replaced by the Mg was incorporated in the
remaining illite, increasing the crystallinity of the illite. This interpretation is highly
600

conjectural. To the southeast in Tennessee and Georgia the only physil in the Lower
Mississippian phosphatic Maury shale is an Fe-rich ( - 7% Fe,03) poorly crystal-
lized illite. The thick Floyd shale of Georgia contains approximately 70% illite with
varying amounts of chlorite and kaolinite.
Farther west in the predominantly carbonate rocks of Indiana and Illinois the
physil suite is composed largely of illite with varying amounts of chlorite, 1/S and
minor kaolinite (Droste and Harrison, 1958). Corrensite is present in the Middle
Mississippian shallow water carbonate rocks of southern Illinois (Fraser et u/.,
1973). Gypsiferous rocks in Indiana and the Michigan Basin contain illite and
chlorite (Droste, 1963). The physil suites of the shallow buried Lower to Middle
Mississippian carbonates and mudstones of southeast Iowa reflect the depositional
environments. Illite (2M) and a lesser amount of I/S (4:l) are present throughout
the section. Minor chlorite is present in the lower marine sediments. Chlorite and
kaolinite increase upward as the sediments become more nonmarine. Kaolinite is
more abundant than chlorite in the most shoreward facies but does not comprise
more than 10% of the physil suites. The Lower Mississippian shales of southwest
Montana mostly contain 70 to 80% illite, 10 to 20% I/S and 10 to 20% chlorite. The
Upper and Middle Mississippian Besa River black shales of British Columbia
contain an average of approximately 85% illite and 15% Fe-rich chlorite and
septachlorite (berthierine). The source is believed to have been volcanic ash (Pelzer,
1966). Illite with minor amounts of chlorite comprises the physil suite of the Lower
Mississippian Leadville limestone of southwestern Colorado (Merrill and Winer,
1958).
The Lower and Middle Mississippian rocks (largely shales) across the southern
United States (Warrior Basin, McAlester Basin, Ardmore Basin, Anadarko Basin,
Bend Arch, Delaware Basin and Marathon Basin) are similar to those to the north.
Illite is dominant. Varying amounts of I/S (9:l to 7:3) and chlorite are present.
Chlorite is generally the minor component (Weaver, 1958; Weaver 1968).
Farther to the south in the rapidly subsiding Ouachita-Marathon geosyncline the
thick, shale-rich Stanley and Tesnus Formations were deposited. The sediments
were derived from the southern Appalachan region and a microcontinent to the
south (Morris, 1974). There is considerable doubt about the age of these two thick
rock sequences; the physil data support a late Mississippian age.
The black shales in the lower portion of the Stanle? have a physil suite composed
of 50 to 60% illite (2M) and 40 to 50% chlorite plus kaolinite. Kaolinite is relatively
abundant. In the northern part of the Ouachitas montmorillonite becomes the
predominant physil. Four tuff beds, up to 40 m thick, occur several hundred feet
above the base of the Stanley. The shales adjacent to the tuffs contain abundant
Ch/S presumably derived from volcanic material. The thicker tuff beds contain an
Fe-rich 1 M illite, possibly celadonite. The thin “tuff beds” contain a 2M illite and
possibly are not tuffs. The physil suite of the shales from the upper portion of the
Stanley consists of approximately 80% illite and 20% chlorite and Ch/S. To the west
(West Texas), the physils in the Tesnus shales are largely illite (2M) and Ch/S
(about 30%) in the black shales (marine bay) of the lower portion, illite and
vermiculite in the middle, and illite and I/S (with a high content of smectite) in the
601

upper portion. Both the Ch/S, vermiculite, and I/S were apparently derived from
volcanic material. The source area to the south apparently contained a mixture of
volcanic and metamorphic rocks.
The presence of the tuff beds and Ch/S in the geosynclinal Stanley and Tesnus
and the relatively abrupt appearance of montmorillonite, I/S, and Ch/S in the
Chester age rocks of the south central and midcontinent United States indicates
there was a volcanic province to the south. The period of volcanic activity is
presumably related to the northern drift of South America and Africa (Gondwana).
The initial plate collision occurred shortly thereafter, near the end of the Mississip-
pian and beginning of the Pennsylvanian. As during the earlier Taconic orogeny the
presence of physils (montmorillonite, I/S, Ch/S) derived from volcanic material
was a harbinger of the collision. The tuff beds thicken and coarsen from north to
south (Niem, 1977).
Montmorillonite and/or I/S (40 to 80%) are abundant in Chester age rocks of
the Forest City, Illinois, Warrior, Arkansas, Ardmore, Anadarko, Palo Duro, and
Delaware basins, Ozark uplift, Hueco Mountains, Williston Basin, western Col-
orado, Utah, and the Great Basin, Nevada (Fig. 9-15). I/S is relatively abundant
(about 20 to 40%) in many Meramecian age rocks from these areas but in a lower
amount than in the Chester rocks. The thick Springer shale of the Ardmore and
Anadarko Basins, Oklahoma, has a high content of montmorillonite - I/S similar to
the underlying calcareous Chester rocks. The age of the Springer has always been a
problem. It has been placed in the latest Mississippian and early Pennsylvanian and
in both. Mineralogically it belongs in the Upper Mississippian; if not Mississippian,
then the physils are reworked Mississippian. The low illite content (Fig. 9-15)
indicates it was deposited before the collision and mountain building activity that
marked the beginning of the Pennsylvanian (Weaver, 1958).
There is little doubt that the expanded clays of the Chester-Springer had a
volcanic origin. As in the Tertiary of the Gulf Coast the proportion of illite layers in
the I/S increases with depth of burial. The “montmorillonite” in the shallow buried
rocks contain about 20% illite layers. Those in the deeper portions of the Anadarko
basin contain 70% illite. The pattern is similar for both the Chester and the
Springer. Most of the Upper Mississippian sediments deposited in the south-central
geosyncline and fringing basins were derived from the south and southeast. Farther
north the sediments were derived largely from the igneous and metamorphic rocks
of the Canadian Shield and the northern Appalachian Mountains.
In the Illinois Basin (Illinois, Indiana and Kentucky) the Upper Mississippian
shales contain appreciably more I/S and kaolinite than the Lower Mississippian
shales. I/S is commonly the dominant physil (Grim et al., 1957; Droste, 1963;
Weaver, 1958). The situation is similar in southwest Montana (Leckie, 1962). The
relatively high I/S content probably indicates volcanic ash from the southern
volcanic chain was blown into the midcontinental region. Much of the midcontinent
craton was land and would have received large volumes of ash. Much of the ash was
probably retransported into the marine basins.
Along the western flank of the craton, in western Colorado, Wyoming, and
eastern Utah, a highly weathered, kaolinite-rich, red soil developed on the flat-lying
602

Devonian and Lower Mississippian limestones during the Upper Mississippian and
lower Pennsylvanian (Molas Formation). Eastward, in the area of the ancestral
Front Range much of the regolith formed by the weathering of granodiorites and
granite gneiss. The major physil in the regolith is I/S. Varying amounts of
montmorillonite, illite and kaolinite are present. Apparently, weathering was less
intense in the eastern portion of the area. A kaolinite-rich sequence extends from
the Upper Mississippian into the lower Pennsylvanian (Atoka) (Merrill and Winar,
1958; Power, 1969). During the Mississippian the equator extended diagonally
across the United States, passing through Colorado and slightly north of the Illinois
Basin. The proximity of these two areas to the equator could account for the
relatively high kaolinite in the sedimentary rocks of both areas.
On the southeastern edge of the craton (Cumberland Plateau) shallow marine to
tidal carbonates were deposited (Monteagle Formation). Mica-illite, I/S and chlo-
rite are present but Ch/S is the predominant physil. Much of the Ch/S is the
ordered variety, corrensite (Peterson, 1962). Vermiculite and dioctahedral chlorite
are also present. The physil suite is complex and changes over short intervals.
Montmorillonite is the major physil in some of the limestones; the chlorite (Mg-rich)
material occurs in both the dolomites and the limestones. The Mg-rich physils
apparently formed in a mildly evaporitic environment, probably from volcanic ash
from the southern volcanic belt. The x-ray patterns indicate some of the mica is
biotite (high 10A/5A ratio). On the other side of the craton corrensite is present in
the Brazer Limestone of Colorado (Bradley and Weaver, 1956).
The physils of the Upper Mississippian presaged the impending collision of
Gondwana and Laurussia. The relative abundance of montmorillonite, I/S and
Ch/S indicates volcanism was extensive. The absence of bentonite-like beds sug-
gests much of volcanic material may have been deposited on the emergent land
areas and later transported to the depositional basins. The abundant kaolinite that
formed in the western United States and the northern Appalachians indicates a
climatic change--increased temperature and humidity.

Pennsylvanian
The deformation and collision that occurred as Laurussia and Gondwana com-
bined to form the super continent Pangaea extended from the Upper Mississippian
into the Permian. Much of the action occurred during the Pennsylvanian. The
worldwide orogeny that melded the two continents is known as the Hercynian
Orogeny. The North American orogenies are known as the Ouachi ta-Marathon
Orogeny (Mississippian-Pennsylvanian) and the Allegheny Orogeny (Mississippian-
Permian) (Fig. 9-13).
At the end of the Mississippian Period, a major regression moved the sea entirely
from the craton, with the possible exception of a few basins (e.g., Anadarko). The
earliest Pennsylvanian marine rocks are restricted to the marginal mobile belts. The
craton was strongly warped and faulted and when the sea returned in mid-Pennsyl-
vanian time a wide variety of unconformities were created. Fig. 9-16 shows the
locations of the basins and elevated areas that will be discussed.
603

In the southern Appalachian region and south of the Ouachita-Marathon trend,


highlands rose during the Lower Pennsylvanian and a relatively thick flysch section
was deposited in the Ouachita-Marathon geosyncline and landward basins. The
physil suites of the shales in these deposits are relatively similar and in their
complexity differ from the physil suites of the older Paleozoic rocks. The suite
typically contains approximately equal amounts of illite (30 to 50%) and I/S (9:l to
7:3) and lesser amounts of chlorite and kaolinite (Fig. 9-17). All four physils are
commonly present. Fig. 9-15 shows, in a general way, the major physils in the Lower
Pennsylvanian of the major basins of the southern United States. The patterns are
meant to show the dominant physils.
Fig. 9-15 and 9-17 demonstrate the differences in the physil suites of the Upper
Mississippian and Lower Pennsylvanian (Morrow and Atoka). Some assign the
Springer to the Pennsylvanian but, if so, it is the only Lower Pennsylvanian
formation in which montmorillonite-I/S is the predominant physil, whereas most
Chester age formations in the region have a composition similar to the Springer.
The I/S in the Lower Pennsylvanian shales commonly appears as an inclined
shoulder on the low angle side of the 10 A peak. When samples are treated with
ethylene glycol, the I/S reflections tend to spread out over a wide interval and not
form a well developed peak. This material resembles a degraded (weathered)
illite-mica with a range of I/S mixtures rather than an I/S formed from montmoril-
lonite, though the latter type is presented and is the dominant physil in some
samples. South and southwest (Texas) of the Anadarko Basin and to the north
(Kansas) montmorillonite-I/S is commonly the dominant physil in the Lower
Pennsylvanian Morrow and Atoka. In the Ardmore and Anadarko basins
montmorillonite-I/S is abundant in a transition zone at the base of the Morrow and
is abundant in scattered intervals in both the Morrow and Atoka, particularly in the
limestones. Some of this montmorillonite-I/S was probably locally derived from
reworking of Springer and/or Chester rocks, but some was derived from volcanic
material. Frezon and Schultz (1961) found four K-bentonite beds in the Atoka
Formation. In the Fort Worth Basin, along the western flank of the Ouachitas, I/S
decreases from north to south. The Morrow and Atoka shales to the east and
northeast in the Warrior, Illinois and Appalachian Basins contain less I/S than the
I/S rich Oklahoma - North Texas material but do contain appreciable I/S with a
high illite content.
In the south-central region at least two general “source-types” supplied physils to
the Lower Pennsylvanian seas. In general montmorillonite-I/S increases northward
and westward from the Ouachitas. Illite and chlorite increase to the south and west.
As suggested by Graham et al. (1975) the major source was probably the Appa-
lachian orogenic belt to the east and southeast and perhaps an orogenic belt to the
south. Flawn et al. (1961) believe pre-Atoka low-grade regional metamorphism
occurred in the southern Ouachitas but most of the thrusting and metamorphism
was post-Atoka. As the South American margin of Gondwana gradually moved
northwestward, it caused the repeated piling up of the accretionary wedges to form
rapidly eroding highlands on the cratonic margins. In west Texas this southern
source was active from within Desmoinesian into early Late Permian time (Ross,
nllu nllta amd chlorlta Montmorllloalta andlor
mlxcd-l.rcr ch).

Fig. 9-15.Distribution of physil facies in the Upper Mississippian and Lower Pennsylvanian rock of the mid-continent region.
605

Uplift
Marine Structural Elemenh
Fig. 9-16. Tectonic features of the mid-continent during the Upper Pennsylvanian and Lower Permian.
From Cubitt (1979). Copyright 1979 Kansas Geol. Survey.
1. Arbuckle Mts. 13. Sedgwick basin
2. Wichita Mts. 14. Anadarko basin
3. Amarillo uplift 15. Plainview basin
4. Apishapa uplift 16. Fort Worth basin
5. Upcompahgre uplift 17. Llano uplift
6. Bighorn Mts. 18. Midland basin
7. Forest City Basin 19. Scurry Platform
8. Nemaha Anticline 20. Central basin platform
9. Salina Basin 21. Delaware basin
10. Central Kansas Uplift 22. Diablo platform
11. Hugoton embayment 23. Northwestern shelf
12. Dalhart basin 24. Central Colorado trough
606

ANADARKO McALESTER ILLINOIS OUACHITA

nALE JACKFORN

CH€Sr€R

SPRINGER FAYETTEVILLE U STANLEY

MERAMEC

Fig. 9-17. Typical X-ray patterns of Upper Mississippian and Lower Pennsylvanian shales. Stippled:
mixed-layer illite-montmorillonite. Black: montrnorillonite. Horizontal-lined: kaolinite. Open 10 A: illite.
Open 7 A and 14 A: chlorite. From Weaver, 1958. Reprinted by permission of American Association of
Petroleum Geologists.

1986). These source areas apparently had a high content of metamorphic rocks and
supplied illite-mica and chlorite. The northward and westward increase in
montmorillonite-I/S indicates there were other sources. Much of this material was
likely derived from local Upper Mississippian rocks that were exposed, or uplifted,
as the seas withdrew from the craton. In south-central Texas the shales contain a
higher kaolinite content (20 to 30%) than is normally found in the Lower Pennsyl-
vanian shales of the south-central area. This presumably reflects a local source.
The change from a montmorillonite-I/S-rich physil suite in the Upper Mississip-
pian to a more complex and relatively illite-rich physil suite in the Lower Pennsyl-
vanian reflects the orogenic activity in the southern Appalachians - Ouachita-
Marathon orogenic belt. The change transgresses time. In the Warrior Basin,
Alabama, the change occurs within the Upper Chester, in Oklahoma the change
occurs at the base of the Morrow (approximately at the Mississippian-Penn-
sylvanian boundary) and in West Texas the change occurs at the base of the Atoka.
Thus, orogenic activity, other than volcanism, started first in the southern Appa-
lachans and migrated west. This is in keeping with the general trend of orogenic
activity whch started earlier in the northern Appalachians and progressively moved
south. Apparently the collision of Gondwana with Laurussia was a pivotal move-
ment.
By Middle Pennsylvanian time both the Ouachita and Wichita-Amarillo Moun-
tains were uplifted and were the major source areas (largely sedimentary and
low-grade metamorphic rocks) in the south-central United States during Middle
Pennsylvanian to Permian time. The physil suites of the younger Pennsylvanian
sediments are similar to those in the Lower Pennsylvanian sediments. Most of the
suites are complex, containing illite-mica, I/S, chlorite, kaolinite and less commonly
Ch/V or Ch/S. Illite-mica commonly comprises 50 to 70% of the physil suite
( < 2 pm). On a regional basis the physils reflect the relative importance of the
various source areas; the Ouachita Mountains had a higher content of well crystal-
lized illite-mica than the other source areas.
The Pennsylvanian shale-rich rocks in south-central Texas (New Braunfels
trough), slightly north of the buried Ouachitas, have an exceptionally uniform physil
suite. In one well (40 samples), through 1200 m of section the 10 A (illite plus minor
I/S)/7 A (chlorite > kaolinite) ratio ranges from 1.4 to 1.8, which is equivalent to a
maximum range of variation of 18 to 22% for the 7 A components. In another well
the 10 A/7 A ratio ranges from 1.0 to 1.7 over an interval of 1700 m. The
chlorite/kaolinite ratio varies to some extent but both physils are present in all
samples.
Farther west in the Val Verda Basin, a series of wells 10 to 70 km in front of the
Ouachita Fold Belt have the same uniform physil suite. The 10 A/7 A ratio ranges
from approximately 1.3 to 2.0 over a distance of 240 km in sections up to 3350 m
thick (Speights and Brunton, 1961).
These physils must have been derived from the low-grade metamorphic rocks in
the Ouachta-Marathon orogenic belt and were deposited in a rapidly subsiding
basin. The uniformity of the physil suite contrasts sharply with the inhomogeneity
of the shales to the north where several source areas existed.
In north-central Texas, two relatively distinct physil suites intermingle, an
illite-mica-rich suite derived from the northeast (Ouachita Mountains) and a suite
relatively rich in kaolinite and I/S, apparently derived from a southeastern source
(Shover, 1964). In the Anadarko Basin, north of the Wichita Mountains, the
10 A/7 A ratio (illite and I/S / chlorite and kaolinite) systematically increases
from 1.0 (approximately 708 illite and I/S) near the Wichta Mountains to 2.5
(approximately 85% illite and I/S) at the northern edge of the basin. Illite increases
from the Lower (Morrow) to the Upper Pennsylvanian (Missouri). I/S is more
abundant in the Upper than in the Lower Pennsylvanian. The physils were presuma-
bly derived from the Wichita Mountains and some segregation occurred during
608

transport. Ch/S is abundant in the western part of the Anadarko Basin and was
apparently derived from the Ancestral Rockies to the west.
In the shelf area north of the Anadarko Basin (Kansas) the composition of the
Middle and Upper Pennsylvanian (Des Moines, Missouri and Virgil) shales also
changes from east to west. Shales from the eastern wells have the typical Pennsyl-
vanian four-component physil suite, though I/S is less abundant than in the Lower
Pennsylvanian shales. Shales from the western part of Kansas contain an apprecia-
ble amount of Ch/S, minor kaolinite and varying amounts of illite and I/S. As in
the Anadarko Basin, illite is more abundant in the younger shales and I/S in the
older. The occurrence of Ch/S in the DesMoines shales of the western portion of
the shelf and Anadarko Basin suggests the Ancestral Rockies which started to
develop at this time, contained basic volcanic material. There is a westward increase
in illite in the Lower Pennsylvanian, suggesting the Ancestral Rockies may have
been a source at this time.
The Pennsylvanian shales of the foreland areas of northern Texas, Arkansas and
Oklahoma have a relatively uniform composition. Illite commonly comprises 50% or
more of the physil suite, mixed layer illite-montmorillonite 20 to 40% and kaolinite
plus chlorite 10 to 30%. The suite is similar to that in the shales of the Midcontinent
and eastern and southeastern interior basins.
Environmental modifications cause some variations in the clay suite but the
overall uniformity is impressive and indicates most clays were detrital and were
derived from a relatively constant source (shales and low grade metamorphics
primarily from the Appalachian-Ouachita-Marathon orogenic belt and interior
uplift) and only mildly weathered. Chemical modification in the depostional en-
vironment was relatively minor except in some of the terrestrial environments. The
non-marine red beds, marine and deltaic shales of Upper Pennsylvanian age from
north-central Texas all have the typical detrital Pennsylvanian physil suites. How-
ever, the underclays commonly contain more than 70% kaolinite (Shover, 1964).
A well developed physil-environmental relation exists in the Upper Penn-
sylvanian (Virgil Series) marine-deltaic sequence on the east flank of the Bend
Arch, Texas. The environmental boundaries shown in Fig. 9-18 are based entirely on
the physil data and coincide almost exactly with boundaries determined by other
techniques. The physil suite is composed predominantly of illite with varying
amounts of I/S, chlorite and kaolinite. A few marine shales contain some
montmorillonite. The 10 A/7 A peak-ratio of the samples from each well is shown
in Fig. 9-18.
Most of the deltaic or transitional shales have 10 A/7 A ratios ranging between 1
and 2. Both chlorite and kaolinite contribute to the 7 A peak. Chlorite is more
abundant than kaolinite. The marine shales are composed largely of illite and
varying amounts of I/S (with a high illite content). The boundary, based on clay
mineral changes, between these two types of sediments is quite distinct. Systematic
variations in the 10 A/7 A ratio of the shales within these two environments reflect
more subtle environmental differences. The lowest average ratio values (1.49 and
1.39) occur in the center of the deltaic facies (wells No. 7 and No. 6). These values
systematically become larger to the east (1.65) and the west (2.33) towards the edge,
609

or more marine portion of the delta. In each well penetrating the delta, the
10 A/7 A ratio increases in the upper and lower portion of the deltaic sediments. In
the marine facies the chlorite and kaolinite (7 A peak) decrease with increasing
distance from the delta and increase from west to east (seaward to shoreward).
The “marsh” shales contain the same general clay suite as the delta, but peak
intensities are considerably less, the I/S contains a higher proportion of montmoril-
lonite layers and the organic content (as indicated by the x-ray patterns) is higher.
Organic acids apparently caused some minor degradation of the I/S.
The physil suite of the marine embayment shales is similar to that of the marsh
shales, though the x-ray peaks are sharper and the organic background is not
present. The suite differs from that of the deltaic shales in that the montmorillonite
content of the I/S is significantly larger and similar to that in the “marsh” shales.
The continental shales in the eastern part of the section have a lower 10 A/7 A
ratio than any of the other shales. In general these shales have higher
kaolinite/chlorite ratios than the shales from the other environments.
In this particular section changes in the physil suites closely relfect changes in the
depositional environment. The deltaic, marsh and marine embayment clays pre-
sumably were derived from the same (eastern) source and the minor differences are
due to environmental modifications. The contrast between the physils in the coastal
shales and the marine shales suggests the two physil suites were derived from
different source areas and transported by different river systems. The illite-rich
physil suite presumably entered the sea some distance from the delta and was
distributed by marine currents.
The detritus deposited in the eastern and midcontinent regions was derived
primarily from the northern Appalachian highlands (sedimentary and metamorphic
rocks) and to a lesser extent from the Canadian Shield (sedimentary and granitic
rocks). Sediments were deposited in shallow water, primarily in deltas, coal swamps
and shallow marine environments. This was a time of crustal, or sea level, instability
when coal bearing cyclothems were formed and the shoreline moved back and forth
hundreds of kilometers across the platform. Compressional uplift of the Appa-
lachian Mountains started in the Upper Mississippian and continued to the Per-
mian.
Though land plants were well developed by late Devonian, they reached their
maximum development in the Pennsylvanian and should have had a significant
effect on the development of soils and physils. Most of the coal bearing deposits
occur within 20 to 25 latitudinal degrees of the apparent equator indicating tropical
to subtropical, humid climatic conditions prevailed.
Wanless et al. (1970) and others have demonstrated fairly convincingly that many
of the Pennsylvanian rocks of the Illinois Basin, and probably of many other areas,
were deposited as sediments on a deltaic-alluvial plain. Low-gradient rivers trans-
ported the sediment and built deltaic platforms largely of mud, up to sea level.
Plants in the interdistributary portions of the rivers formed coal swamps. Limy
muds formed in the bays, lakes and lagoons. Slight raises in sea level and shifting of
distributary channels allowed the sea to invade the deltaic plain halting plant
growth and temporarily allowing marine sediments to accumulate until the
610

deltaic-alluvial plain again started to prograde. Aside from the swamps, such an
environmental complex does not allow much opportunity for physical and chemical
modification of the detrital physil suite, thus accounting for the relative homogene-
ity of the physil suites in these sediments.
Studies of cyclothem units from Tennessee (Weaver), Pennsylvania (Degans et
al., 1957), Illinois (Murray, 1954; Potter and Glass, 1958; Glass, 1958), and
Iowa-Missouri (Brown et al., 1977), ranging in age from Lower to Upper Pennsyl-
vanian, indicate the physil-environment relations are similar over this broad region.
The physil suite is typical Pennsylvanian, illite, I/S, kaolinite and chlorite (or
vermiculite). In all rocks other than the underclays, illite is the predominant (50 to
80%) physil. The other physils are present in varying amounts; chlorite is usually the
minor component. Kaolinite is more abundant in the nonmarine facies and illite in
the marine facies. In the eastern areas I/S is commonly more abundant in the
marine facies. In Iowa-Missouri, I/S is more abundant and has more smectite layers
in the nonmarine facies than in the marine facies.
Though the general relations are consistent, there is a considerable amount of
variation, as might be expected in a complex of closely related coastal environments.
The relatively high kaolinite in the more nonmarine environments is probably due
to weathering at or near the site of deposition.
Analyses of a complete Pennsylvanian section from the southern part of the
Illinois Basin (Potter and Glass, 1958) show that chlorite (14 to 24%) and kaolinite
(11 to 15%) increase and illite (61 to 40%) decreases from the older to younger
Pennsylvanian. I/S remains relatively constant (14 to 15%). Based on a study of the
sandstones, orthoquartzites in the Lower Pennsylvanian to subgraywackes in the
Upper Pennsylvanian, the authors concluded that the early sediments were derived
from pre-existing sediments, but erosion progressively unroofed metamorphic and/or
igneous rocks in the source area. In another study of sandstone and shale pairs from
the Lower Pennsylvanian, Glass et al. (1956) found that kaolinite was consistently
more abundant and illite and I/S less abundant in the sandstones. Thin-section
studies indicated the kaolinite in the sandstones had a post-depositional origin. At
what stage it crystallized was not established.
Extensive deposits of fire clay and diaspore developed in soils and swamps to the
north of the Ozark dome, Missouri (Keller, 1968).

Underclays
Underlying many of the coal beds are nonlaminated physil-rich beds commonly a
few feet thick but usually having an irregular thickness. Plant fragments or impres-
sions and slickensides are commonly present. Some beds show evidence of root
penetration downward from the coal. These beds are called underclays and early
were considered to be soils. Mineralogic studies indicate only a few have the vertical
variation indicative of a soil profile.
The principal physils in underclays are poorly crystallized kaolinite, I/S, illite
and to a lesser extent vermiculitic and chloritic materials. Schultz (1983) studied 10
underclays (400 samples) from Lower to Upper Pennsylvanian age from the
Appalachian, Illinois and Mid-continent Basins. H e found no consistent mineralogic
611
.-v1
d
e
h
M
ii
612

Table 9-3
Mineral Composition Variation in Underclay Profiles (Schultz, 1958).
TOP Same Bottom
Kaolinite 31 * 38 33
more abundant
Montmorillontie 29 60 6
more abundant
Feldspar 10 13 8
more abundant
Mica 4 9 39
more abundant
14 A mineral 6 12 31
more abundant
14 A mineral 5 21 4
more Vermiculite
* in 31 profiles the top was more kaolinitic than the bottom

trends within the underclay profiles to indicate there had been significant post-de-
positional leaching (Table 9-3). Chlorite, muscovite and quartz are commonly more
abundant and coarser in the lower portion of the underclays. This is believed to be a
depositional feature related to a decrease in current velocity with time. Schultz
concluded that the physils in underclays were largely detrital and derived from soils
formed in the source areas during periods of slow erosion which preceded almost
complete stagnation of the erosion-depositional cycle when peats were formed.
Other studies show some of the highly kaolinitic underclay may, in part, be the
result of some post-depositional leaching by organic acid water. Some of the I/S
was probably developed at the same time.
Shales, largely brackish to marine, from above the coals contain more illite and
chlorite, and less I/S and kaolinite than the underclay. Most of the chlorite and
kaolinite in the shales is better crystallized. The shales, sandstones, and calcareous
rocks more intimately associated with the underclays are mineralogcally similar to
the underclays.
Table 9-4 shows the average kaolinite content of underclays from the shelf,
intermediate and deep basin portions of the Appalachian and Mid-Continent
Basins. The sections are thicker in basin areas and the underclays are thinner and
more sandy. Kaolinite is generally more abundant and less well crystallized in the
thicker shelf underclays. The basin underclays contain less I/S and more chlorite.
The Upper Pennsylvanian underclays are less kaolinitic than the Lower Pennsyl-
vanian underclays. The contrast is more pronounced in the western area where I/S
is predominant and marine sediments are more abundant.
The variations in the composition of physil suites are apparently related to source
and climate. In the Appalachian Basin cross-bedding studies indicate the highly
kaolinitic shelf underclays were derived from the slightly positive Cincinnati Arch
and Canadian Shield where the weathering period was relatively long. The less
kaolinitic underclays in the deeper parts of the basin were derived from the
relatively high relief area on the northeast flank of the basin.
613

Table 9-4
Average Kaolinite Content of Underclays, in percent (Schultz, 1958).
Zone Mid-ContinentBasin Appalachian Basin
Shelf Int. Deep Basin Shelf Int. Deep Basin
10 19 20 29
9 9 33 20
8 3 34 33 53
I 5 29
6 21 22 50 38 17
5 42 58 30
4 48 30 42 27 29
3 51 40 52 38 25
2 38 40 49 48 26
1 51 42

The upward decrease in kaolinite indicates the climate changed from warm and
humid in the Lower Pennsylvanian to a cooler and/or less humid climate in the
Upper Pennsylvanian. The flora indicate a similar trend.
In a more detailed study of several underclays from Illinois and Ohio, Parham
(1963) established that the distribution of the various physils was related to
paleogeographic position in the basin of deposition. Kaolinite is most abundant in
the nearshore facies. Illite, chlorite, and I/S increase seaward with the latter
showing the more systematic increase. The physil distribution is similar to that
found in most coastal to shallow marine deposits and further substantiates the idea
that the physils in underclays are largely detrital. Differences in source and climate
would account for the variations in the relative amounts of physils in the under-
clays.
Though there is an apparently normal seaward change in the physil suites in
some underclays, there is no indication that any underclays are marine. The mineral
fractionation presumably occurred in coastal swamp environments. These swamp
areas were apparently hundreds of kilometers in extent and mineral fractionation,
similar to that which occurs in the present shallow marine environments, could have
occurred.
A detailed study of the I/S physils indicates that the underclay beneath the
Herrin (No. 6) coal in southwestern Illinois has undergone mild in situ alteration
(Rimmer and Eberl, 1982). Throughout an interval of less than 1.5 m the expanda-
ble layers in the I/S decrease from 70% for samples near the coal to about 20% for
those at the base of the section (Fig. 9-19). This is accompanied by an increase in
the amount of ordered I/S with respect to randomly interstratified I/S and an
increase in the amount of discrete illite, chlorite and calcite. Some of the expanded
layers are vermiculitic. The pH increases from < 3 immediately under the coal to
pH 5 at 60 cm depth; it increases abruptly to 8 at 70 cm, where calcite is first
encountered.
Post depositional acid leaching has overprinted the detrial mineralogy, largely
I/S. This does not negate the suggestion of others that some weathering may have
614

0 0
20 0
0

O O
-5
I

40 0
0
0

..
O.
0 8
=
2
Y
80 A
m A
AA A INTERLAYERING
A O R
A 0 R+IM

120
A8 A
A
IMtR
IM

140
I
f
A
I I 1 I
0 20 40 60 ao
PERCENT EXPANDABLE L A Y E R I
Fig. 9-19. Relationship between depth below coal, percentage of expandable layers, and type of
interlayering for illite/smectite from the < 2 p m size fraction of underclay samples. From Rimmer and
Eberl (1982). Copyright 1982 the Clay Miner. Soc.

occurred in the source area (e.g., illite to I/S). It is not clear when the leaching
occurred. The pH values suggest it could still be going on. The authors suggest the

-
sequence illite + I/S ordered + I/S random could be extended to include
kaolinite boehmite diaspore.
-+

Various studies indicate some underclays have suffered little or no postdeposi-


-+

tional leaching, some have been mildly leached (I/S) and others have been thor-
oughly leached (kaolinite). Local environmental conditions, both during and after
deposition, apparently determine the effectiveness of acid leaching.
Kaolinite-rich flint clays (hard, fine-grained, conchoidal fracture, lack of plastic-
ity and composed primarily of kaolinite), fire clays and ball clays are common in
Pennsylvanian underclays, particularly in the Lower Pennsylvanian. They occur as
irregular lense or channel-like deposits and generally have a brecciated texture. Flint
underclays have been described from Kentucky (Patterson and Hosterman, 1960).
Pennsylvania (Williams et al., 1968; Hosterman er d., 1968), Illinois, Indiana, and
the Ohio region (Smith and O’Brien, 1965; Hughes and White, 1969). The Missouri
fire clays, covering an area of 7700 sq. km, are similar to the underclay flint clays
but not associated with coal beds (Keller et al., 1954).
615

The flint clays are composed predominantly of well crystallized to poorly ordered
kaolinite. The well crystallized clay occurs as mosaics of interlocking books, which
accounts for the hardness of the clay. The more poorly crystallized kaolinite can be
plastic. Minor illite and I/S (up to 25%) are present in many of the plastic flint
clays. The flint clays originated in coastal swamps. The consensus opinion is that
illite, I/S and kaolinite, present in the surrounding sediments, were deposited in
these acid swamps and altered, in varying degrees to kaolinite and some instances
diaspore. Commonly, the relatively pure kaolinitic clays (flint clay) grade laterally
into facies containing progressively more illite and I/S and less kaolinite (semiflint
and plastic clays).
In addition to the low pH conditions created by organic acids it has been
suggested that bacteria could have played a major role in creating the acid
environment (see Hughes and White, 1969). Ivarson et ul. (1980) have shown that in
a little over a one year period iron-oxidizing bacterium dissolved 15 to 20% of a
montmorillonite and converted much of a muscovite to I/S. Keller et al. (1953)
suggested the Missouri flint clays formed by desilication and K removal from illite
via the Donnan effect. This process involves the diffusion of ions, such as Si and K,
from illite into the water and H and OH ions from the water into the illite until
equilibrium is obtained. When the Si and K saturated water is flushed from the
swamp by the influx of fresh water the reaction continues and eventually the Si/AI
ratio is such that kaolinite crystallizes.
Altered volcanic ash beds are common in the Carboniferous of Europe but only a
few have been reported in the United States. One bed, altered to flint clay, of
Middle Pennsylvanian age occurs throughout most of eastern Kentucky and in parts
of Tennessee, Virginia and West Virginia (Chestnut, 1984). Another bed, altered to
montmorillonite, is present in Virginia (Nelson, 1959). The volcanic source is
believed to have been eastern North Carolina. Volcanism was probably more
extensive but, because of the predominantly clastic deposition during the Pennsyl-
vanian, much of the ash was probably reworked and mixed with other sediments.
Low temperature ashing of coals from the Illinois Basin (Gluskoter, 1967; Rao
and Gluskoter, 1973) shows the ash contains on the order of 50% physils, illite, I/S,
kaolinite and minor chlorite. The physil suite and the lateral variations are similar to
those in the associated underclays. Much of the kaolinite is secondary and occurs in
desiccation cracks and small cavities. Though some kaolinite crystallized late in the

?
depositional hstory it appears that the detrital illite and I/S was not significa y
altered in the coal swamps. This suggests the detrital clays are entrapped, rapid1 , in
the peat and have relatively little contact with the open swamp waters. On the basis
of SEM studies of West Virginia, Grady (1984) concluded illite crystallized in the
peat at a relatively early stage. Renton et al. (1980) suggested physils crystallize in
the live plant cells (p. 269).

Discussion Eastern and Midcontinent Regions


A fairly distinctive, complex, but relatively consistent physil suite, is present in
Pennsylvania age sedimentary rocks: illite dominant; varying amounts I/S, kaolinite
and chlorite; Ch/S and/or vermiculite have a scattered occurrence. The physils
616

were derived largely from older sedimentary and metamorphic rocks located in
active highlands along the eastern and southern flanks of the North America craton.
Mountain building was a result of plate collision. Though the climate was humid,
erosion was relatively rapid in the mountains and only relatively minor amounts of
kaolinite and I/S were formed by weathering. Weathering was locally extensive in
coastal swamps.
The presence of kaolinite and I/S, the K.I. and W.I. and the CAI (color
alteration index of Conodonts) indicate most of the physils have not been materially
altered by burial metamorphism. Most of the Pennsylvanian rocks discussed have
been exposed to burial temperatures less than 150"C, probably less than 100°C.
Burial depths and temperatures were highest in the basins along the continental
margins. In areas extremely close to the orogenic activity, such as the Narragansett-
Norfolk Basin of Massachusetts and Rhode Island, which contains anthracite coal
and volcanics, Late Pennsylvanian physilites have been subjected to temperatures
high enough ( > 350°C) to produce epizone conditions and higher (Hepburn and
Rehemr, 1981).

Western North America


From the Middle Pennsylvanian to Permian the Ancestral Rocky Mountains
were created by tensional forces which produced a broad zone of basement fault
block mountains and associated basins along the western edge of the North
American craton. They were bounded on the west by the Cordilleran Geosyncline.
In some areas, basement was exposed and large volumes of arkosic sediments were
deposited in the fault-block basins.
During the Late Mississippian and Early Pennsylvanian a soil zone (Molas)
developed on the relatively flat-lying carbonate and granitic rock in the Rocky
Mountain region (Paradox Basin to the east flank of the Ancestral Rocky Moun-
tains) (Fig. 9-16). The physils are illite, I/S and kaolinite in varying proportions
(Power, 1969). The physil in the underlying unweathered rocks is largely illite.
Kaolinite comprises more than 50%,up to loo%, of the physils in most soil samples.
The conversion, illite to kaolinite, indicates weathering was relatively intense and
the climate probably warm and humid. Many of the soil samples have compositions
similar to those of the flint clay-semiflint-plastic clays. The regolith is overlain by a
thin marine unit. The physils in the Glen Eyrie, on the east flank of the Ancestral
Front Range, are largely montmorillonite and kaolinite; the physils on the western
flank, Belden Formation, are predominantly illite and I/S and lesser amounts of
chlorite and kaolinite (Raup, 1966).
Uplift of the Ancestral Front Range followed deposition of the Lower Pennsyl-
vanian marine formations. The mountains reached their maximum elevation during
the Middle Pennsylvanian. Large volumes of red continental to marine arkosic
sediments were deposited on both flanks of the Front Range, the Fountain Forma-
tion on the east flank and the Minturn Formation on the west flank. The physil
suites of the two formations are generally similar. The physil suites are complex and
quite variable in composition. Illite and I/S (2:l to 3:l) are most commonly the
dominant physils. Ch/V, Ch/S, chlorite, kaolinite and montmorillonite are present
617

in varying amounts: all except Ch/S occur as major components in some samples.
Some of the kaolinite is secondary. The Front Range was presumably the source of
the Ch/V (or Ch/S) and presumably the other physils that occur in the Pennsyl-
vanian shales of western Oklahoma and Kansas. Erosion and deposition were
relatively rapid and the physils were only mildly weathered. The physils suggest
climatic conditions were dryer than in the Early Pennsylvanian but not arid. The
presence of evaporites west of the Front Range indicates arid conditions prevailed
during part of the Middle Pennsylvanian.
Farther north, in southwestern Montana, the physils in the Lower to Middle
Pennsylvanian red shales and carbonate rock (Amsden Formation) are predomi-
nantly illite with, generally, a lesser amount of I/S. Minor chlorite and kaolinite are
present in about one-third of the samples (Leckie, 1962). The overlying Quadrant
orthoquartzites contain primarily secondary kaolinite (Weaver, 1955).
The next major Pennsylvanian uplift west of the Ancestral Front Range is the
Uncompahgre of Utah, Colorado and New Mexico. This uplift formed the northeast
boundary of the Paradox Basin. Displacement appeared to have been greatest
during the Des Moinesian, when the Paradox Basin was filled and a thick sequence
of cyclic evaporites were deposited (Paradox Formation). Late Pennsylvanian rocks
are largely arkosic clastics derived from the Uncompahgre Uplift. The cyclic
character of the marine-evaporite depositions is believed to be due to global eustatic
sea level changes coincident with glaciation in Gondwanaland. These same sea level
changes apparently accounted for the cyclic nature of the eastern coal deposits.
The physil suite in the “near normal” marine sediments in the lower part of the
Paradox and underlying formation is composed predominantly of illite and a
slightly lesser amount of chlorite (Bodine and Rueger, 1983) derived from the
Uncompahgre Uplift. These physils are also present in the evaporites (anhydrite,
halite, K-salts) but in addition a variety of Ch/S-types, corrensite, talc, Al-talc, and
serpentine are present. These latter physils are Mg-rich and were formed authigeni-
cally and diagenetically under hypersaline conditions in an arid climate (Padan and
Weaver, 1985). The physils in the overlying Upper Pennsylvanian shallow marine
sandstones and carbonates are largely I/S (10 to 40%) and chlorite. The lack of illite
and abundance of I/S suggest the climate became more humid and weathering
more intense. K/Ar data indicates most of the illite in the evaporite section is
detrital.
Farther west in the Cordilleran geosyncline a carbonate-rich sequence (Oquirrh
Formation) was deposited (Tooker, 1960). Illite is present in all rocks studied. It is
dominant in the limestones and less abundant in dolomitic limestones and quart-
zites. Corrensite is most common in rocks containing some dolomite. Chloritic
material, with some expandable layers, is the dominant physil in the dolomitic
quartzites. Kaolinite was found only in quartzite. The Mg content of the clay suite
increases as the relative proportion of dolomite increases, suggesting the chlorite
and corrensite are authigenetic or diagenetic. Corrensite has been found in the
Upper Pennsylvanian-Permian Supai Group in the Grand Canyon, Arizona (Hauff
and McKee, 1979). Volcanic rocks occur in the western portion of the geosyncline,
suggesting the chloritics and corrensite may have formed from volcanics.
618

General Discussion
Various studies of the organic material in coal beds indicate the climate condi-
tions changed from dry in the Upper Mississippian to warm and humid in the
Lower to Middle Pennsylvanian, when the best coal beds were formed. Starting in
Desmoinesian time rainfall decreased and the climate became dry, tropical (Schutter
and Heckel, 1983; Phllips et al., 1983; Cecil, 1983; Donaldson et al., 1983). The dry
climate continued on into the Permian. The climatic change is believed related to
the progressive drift of North America north of the equator (Fig. 9-11, Fig. 9-13).
The Pennsylvanian physil suites generally confirm the climates deduced from
other data. Kaolinite is relatively abundant in the Lower Pennsylvanian sedimentary
rocks of much of the present day United States. It is significantly more abundant
than in the older Paleozoic rocks. Kaolinite is the physil most commonly formed by
weathering in warm, humid climates, if the weathering period is long enough. Much
of the Lower Pennsylvanian kaolinite was formed in coastal swamps and low lying
soils.
One question might be, why kaolinite was not more abundant. The paleogeogra-
phy must have been somewhat similar to that of the present Amazon drainage basin
(p. 195). In the low lying tropical part of basin kaolinite is being formed. However,
the physil suite delivered to the ocean is dominated by the mildly weathered physils
derived from the Andes Mountains. The North American continent was ringed on
the east and south by a high mountain range. Though, kaolinite was forming in the
wet coastal areas, most of the physils supplied to the deltas were derived from the
high mountians where erosion was rapid and weathering moderate to minor.
As the climate became dryer in Middle Pennsylvanian time kaolinite became less
common. Conditions appear to have been dryer in the west than the east. Only 23%
of the Middle and Upper Pennsylvanian rocks flanking the Ancestral Front Range
contain kaolinite (Raup, 1966), whereas, in the midcontinent and eastern United
States probably more than 80% of the shales contain kaolinite.
The distribution of physils variously classed as corrensite, Ch/S, and Ch/V is
interesting. It is relatively abundant in the Upper Mississippian rocks of the
southeast and southcentral United States and the Pennsylvanian rocks of the
western United States. There is little doubt that most of this chloritic material that
occurs in evaporite and shallow water dolomitic rocks is authigenic and/or di-
agenetic. Though it forms readily from volcanic material, it can apparently form
from illite if the Si/Mg values are correct. On the other hand, the presence of this
chloritic material in the thick geosynclinal Stanley and Tesnus shales of the
Ouashita-Marathon Belt and in the continental red bed deposits flanking the
Ancestral Front Range suggests some of it was formed by the mild weathering of
mica or chlorite and/or by the more extensive weathering of volcanic material.
Chloritic material formed from the weathering of mica should have a high charge
and be a Ch/V, whereas, the chloritic phase formed from volcanic material is
usually Ch/S. The mixed-layer phases formed by weathering chlorite can probably
be either Ch/V or Ch/S. Unfortunately, in many instances the mixed-layer phase is
not accurately identified; commonly the classification is not clear cut.
Aside from evaporite deposits the distribution of the Ch/S, Ch/V, and corrensite
619

is associated with orogenic activity. Thus, its presence in the southern and eastern
United States is contemporaneous with the approaching plate collision and the
development of volcanic areas. As collision and thrusting occurred in the Early
Pennsylvanian volcanic rocks became less important in the source area and the
major source rocks were sedimentary and metamorphic. Active tectonic activity
shifted to the west, where block faulting was extensive. Presumably some volcanic
activity occurred. The mixed-layer chloritic facies also shifted to the west.
The association of mixed-layer chloritic material with tectonic activity could be a
result of high relief source areas and relatively rapid erosion. This is less likely to
explain the general distribution but probably accounts for much of the mixed-layer
chloritic material that occurs as a minor component in the shales of both eastern
and western United States.
Another word about the chloritic physils. Chlorite is the most easily weathered
physil under most weathering conditions. It is virtually impossible to collect a
surface sample of an illite-chlorite shale in which the chlorite does not contain some
expanded layers. For years I used Ch/V or Ch/S as an index of the freshness of
outcrop samples. If expanded layers were present I assumed the sample was
weathered. This is too often the case, dig deep. Reports of Ch/V, Ch/S in outcrop
samples have to be treated skeptically. However, such physils are common.

Europe, Africa, South America

Europe
The Ouachita-Appalachian orogenic belt, which was the source of most of the
sediments deposited in central and eastern United States, extended into northwest
Africa (Mauritanides) and northwest Europe (Hercynian). The Hercynian fold belt
is more complex and less well understood than the United States counterpart
(Windley, 1977). There were several periods of deformation during the formation of
the Hercynian belt. There were three main phases of deformation: Bretonian
(Devonian - Carboniferous boundary), Sudetian (Meramecian - Chesterian
boundary), and Asturian (Desmoinesian-Missourian boundary).
Following a regression at the end of the Devonian, the Carboniferous period
began with a transgression across the Old Red Sandstone, There was a major phase
of lava extrusion in the mid-Lower Carboniferous. The Sudetic orogenic activity,
mid-Carboniferous, was associated with granitic intrusion and acid-to-intermediate
explosive volcanism. This gave rise to intermontane basins in which continental
sediments and coal deposits were localized. The felsic volcanic and plutonic activity
continued in the Late Carboniferous, through the Asturic phase of deformation, into
the early Permian. There is a considerable difference of opinion on the nature of the
plate motions that produced the Hercynian fold belt. In general, it is related to the
interaction between the Laurasia and Gondwana plates.
The most common physil suite in the Lower Carboniferous of Great Britain is
composed primarily of kaolinite and illite in various combinations, either may be
predominant. Some chlorite and, to a lesser extent, I/S are present in a few samples
620

(Perrin, 1971). In the Craven Lowlands, Yorkshlre kaolinite is commonly more


abundant than illite in the marine shales (Purton and Yowell, 1969).
The physil suite of the Upper Carboniferous of Ireland and England is more
complex and shows the effects of local source areas. In addition to kaolinite and
illite, I/S is commonly a significant component. In some areas illite is predominant
and varying amounts of chlorite and Ch/V are present. I/S and illite are the major
physils in some samples from Wales. Most of the physil suites in the Coal Measures
(Middle and Upper Carboniferous) contain only kaolinite and illite. I/S and
chlorite are present in a small percentage of samples. Kaolinite fireclays, with minor
mica, are abundant (Perrin, 1971).
In Scotland, illite (largely I/S with minor smectite) is the most abundant physil
in the lower Carboniferous (largely lacrustrine and shallow marine sediments).
Kaolinite increases in abundance towards the top of the Lower Carboniferous rocks
and in the Upper Carboniferous (largely deltaic) is the dominant physil with illite
being subordinate. Kaolinite is dominant in both marine limestones and shales, and
in non-marine shales and underclays is more poorly ordered than that in the other
rocks, suggesting it was further weathered after deposition (Wilson et al., 1972).
In South Wales coalfields, the physils are primarily illite (58 to 77%) and I/S (18
to 35%) with variable amounts of kaolinite (3 to 40%) and chlorite (3 to 12%).These
rocks have been exposed to Late Hercynian metamorphism and I/S decreases to the
west as the crystallinity of the illite increases. The maximium grade of metamor-
phism, in the athracite area, is anchimetamorphism (Gill et af., 1977).
Wilson (1965) found underclays (seat rock) to be associated with all the hundreds
of coal seams he examined in South Wales. They are fine grained, contain fossil
rootlets and contain randomly oriented slickensided surfaces. Many contain iron-
stone nodules. Carbonaceous matter and rootlets are concentrated in the upper
portion, slickensides in the central portion and ironstone nodules in the lower part.
The lower part grades into unweathered shales or sandstone from whch the
underclay was formed. The profile is similar to that of a soil.
The physils in the underclays (76) from the Lower and Middle Coal Measures
consist largely of kaolinite and illite. Crystallinity of the kaolinite is moderate.
Many of the underclays from the Upper Coal Measures are characterized by an
abundance of chlorite, sharp illite x-ray peaks and little or no kaolinite. There is
little vertical variation in the clay mineral suite except for a slight broadening of the
001 illite peak for illites at the top of the underclay. Wilson believes the clays are
detrital, formed outside the swamp area and have suffered little post-depositional
alteration. The amount and presumably size of quartz increases with depth in the
profile. He believes the underclays are not true soils but were deposited as sediments
in an oxidizing non-swamp environment. Vegetation started to grow retarding the
velocity and carrying power of the water. As vegetation became more concentrated
slack water conditions developed and only fine clay was deposited. At the same time
the environment became reducing and the organic matter was preseved.
The presence of easily weathered detrital chlorite in the Upper Coal Measures
suggests the climate was drier and weathering less intense than during Lower and
Middle Coal Measure time. A change in plant species suggest a similar change in
621

climate. These changes are similar to those that occurred at approximately the same
time in the central and eastern United States.
Illite, with I/S, is the predominant and commonly the only physil in the Lower
Carboniferous limestone of Belgum (Dandois, 1981; Thorez and Bourgouignon,
1973).
The physil suite of the shallow buried Upper to Middle Carboniferous shales in
the Upper Silesian Coal Basin of Poland is composed primarily of illite and I/S
(25% smectite) with minor chlorite and kaolinite. With depth, as coal rank increases,
the kaolinite disappears, being converted to illite, the smectite layers in the I/S
decrease to 10% and chlorite develops (Srodoh, 1978).
In the Alsace region of France the Upper Carboniferous physils contain an
average of approximately 40% kaolinite and dickite and approximately equal
amounts of illite (1Md) and I/S. Kaolinite averages 60% in the coal bearing
interval. Dickite occurs in fractures and is believed to have a hydrothermal origin
(Dunoyer de Segonzae, 1969).
In addition to the kaolin deposits associated with underclays (seat earths) and
tonsteins, kaolinite-rich weathering crusts were developed on many of the Hercynian
or Variscan uplifts (Massifs) during the humid tropical to subtropical Carbonifer-
ous, primarily Upper Carboniferous. Kaolinite weathering crusts developed on the
Bohemian Massif (Germany, Poland and Czechoslovakia), Baltic Shield, Massif
Central of France and others. Relict crusts are still present on some of the uplifts
but in many instances the presence of the crust is inferred from the presence of
kaolinite-rich sediments flanking the uplifts. The shift of climate to semiarid at the
beginning of the Permian terminated the kaolinite-type weathering (Storr, 1975).
Throughout the Upper Paleozoic the Asian continent continued to converge with
Laurasia although there was a connection between the Urals Trough and the Tethys
to the south. A mountain chain did not develop until the Permian. During the
Lower Carboniferous limestones were deposited on much of the Russian Platform
and the western Urals. The limestone deposits were interrupted by the development
of coal deposits, along with lava and tuffs, during the late Lower Pennsylvanian
(Visean). In the Middle Carboniferous newly rising mountains started encroaching
on the Ural Trough and coarse clastics, red beds and gypsum migrated westward.
Coal beds were deposited intermittantly from the Lower Carboniferous to the
Lower Permian. Russian geologists recognize a major boundary somewhere in the
middle of the Russian Carboniferous. The boundary approximately coincides with
the major change in the physil suite that occurs in the Upper Mississippian of North
America.
The physil suites in the Carboniferous of Russia are in general similar to those in
North America. The main difference is that the development of coal swamps, along
with kaolinite and I/S, began much earlier (Visean, Mississippian). Another major
difference is the presence of appreciable palygorskite in the Upper Carboniferous of
the Russian platform.
The western part of the Moscow Basin in Visean time was a low swampy coastal
plain, periodically submerged by a shallow epicontinental sea. Well developed
kaolinite weathering crusts formed in the surrounding mountain areas during
622

periods of quiescence and when coal swamps were well developed (Strakhov, 1969).
Detrital kaolinite developed on Devonian illitic rocks to the west and was delivered
to all depositional environments. Pure kaolinite occurs in the coastal lacustrine and
coal swamp sediments. The original detrital suite consisted largely of kaolinite and
illite (hydromica or weathered mica). The kaolinite purity was further improved by
the alteration of the illite by humic acids and crystallization of kaolinite in the
swamp environments. The physils in the littoral-marine and submarine deltaic
sediments are also largely kaolinite and illite but I/S, montmorillonite and beidellite
are also present. These latter three clays are believed to have formed from kaolinite
in the marine environment; more likely they are detrital (Vikulova, 1964).
Environmentally controlled physil facies are also well developed in the Lower
Carboniferous of the southwestern part of the Moscow Basin (Zkhus, 1961).
Continental sediments are characterized by kaolinite with lacustrine deposits having
hexagonal flakes and alluvial and paludal sediments having fragmented kaolinite
flakes. The marine sediments contain beidellite and montmorillonite. The physil
suites are not pure but contain two or more clay mineral types. Karpova (1969)
reports that in Carboniferous terrigenous clay-type rocks at the southwestern border
of the Russian Platform (Bolshoy Donbas) kaolinite comprises only about 10 to 15%
of the clay suite and is absent in the Upper Carboniferous. The major clay mineral
is 2M illite (hydromica) in various degrees of hydration. Additional kaolinite,
formed from the illite, occurs in the paludo-lacustrine environments (refractory
clays, underclays, and tonsteins). Montmorillonite is present in the marine sedi-
ments. In the Karaganda Basin a series of about 20 beds (several centimeters to a
few meters thick) of acidic pyroclastic material are distributed through the 4000 m
thick coal-bearing section (Lower to Upper Carboniferous) (Shutov et al., 1969).
Palygorskite and, to a lesser extent sepiolite, are relatively abundant in the
Devonian, Carboniferous, and Permain rocks of the Russian Platform. Russia has
the most extensive deposits of Paleozoic chain structure clays. Many of the reported
occurrences are vein deposits of hydrothermal or weathering origin (Rateev and
Kotelnikes, 1956); however, extensive deposits occur along the flanks of the Moscow
Basin in the Middle Carboniferous terrigenous and shallow marine carbonate
(largely dolomites) rocks (Rateev, 1964). Sepiolite and palygorskite have been found
in more than a hundred horizons in the Kashira Basin of the Russian Platform
(Rateev and Timofeev, 1929). These deposits range in age from late Lower
Carboniferous (Chesterian) to Middle Carboniferous. The clay beds occur primarily
in shallow water marine to lagoonal carbonates. Similar deposits are present in the
Upper Carboniferous deposits of the Moscow syncline. Palygorskite was deposited
during initial aridization (mark and limestones) and sepiolite during intense aridiza-
tion (dolomite and gypsum). The Mg, Al, and Si were supplied to the shallow-water
platform basins from adjacent areas (southwest) where kaolin weathering crusts
were forming. The A1 was probably supplied as montmorillonite.

Africa
The Carboniferous rocks of the Polignac Basin, northwest Africa, contain a
relatively uniform, “typical”, Carboniferous physil suite of 50 to 70% illite and I/S
623

and approximately equal amounts of kaolinite and chlorite (Dunoyer de Segonzac,


1969). In contrast to Europe and North America kaolinite is less abundant than in
the older Paleozoic rocks.

South America
The Middle Carboniferous of the Paranh Basin of southern Brazil contains
sedimentary rocks deposited in continental glacial, fluvioglacial, glacialmarine to
marine environments. The source area was the Brazilian Shield to the north though
glacial deposits were derived from the south. Analyses of samples from 42 wells
indicate illite, relatively well crystallized, is the predominant physil, followed in
abundance by chlorite. Montmorillonite is locally abundant. Kaolinite and I/S are
rare. The physils and other data indicate a cold climate and a high-relief source
area.
During the Upper Carboniferous, deltaic, tidal flat, coal swamp and shallow
marine environments became abundant. The source area was to the east (Africa)
and the climate was more humid and warm than in the Middle Carboniferous. The
climatic change is reflected in the physils. I/S is the dominant physil. Varying
amounts of illite are present. Kaolinite, some secondary, is abundant in the beds
associated with the coal beds. Detrital kaolinite increases to the east and is most
abundant in the nearshore and terrestial deposits (Ramos and Formosa, 1976).
To the southeast, in Uruguay, the Upper Carboniferous sediments are largely
tillites. Kaolinite comprises 70 to 90% of the physil suite; minor illite and I/S are
present (Elizalde and Steinberg, 1973). In northwest Argentina (La f i o j a Province),
kaolinite is the dominant (958) physil in the Upper Carboniferous coal fluvial and
swampy lagoonal deposits. The sands are arkosic and the kaolinite was apparently
produced by the weathering of basement granites. Many of the clay beds are
commercial (Iniquez and Zalba, 1979).
In the Paranii Basin of east-central South America the physils, largely detrital,
reflect an increase in weathering intensity (kaolinite, I/S, degraded illite) in the
Upper Carboniferous. Kaolinite increases to the east and southeast, towards the
shore and towards the Carboniferous ice sheet. Though some of the physils were
derived directly from older Paleozoic sedimentary rocks, the change in the nature of
the physil from the Middle to Upper Carboniferous indicates there was appreciable
weathering in the source area flanking the ice sheet. Coals are relatively abundant in
the flanking deposits. The coals in Europe and North America were derived from
lycopsid flora that grew in tropical to subtropical climates; however, a different
flora, the Glossopteris flora, formed the glacial coals. These latter flora are believed
to be deciduous and to have grown in a temperate climate. Thus, even though the
climate was temperate, appreciable organic material was present and in the intergla-
cial periods the environment was apparently humid enough that weathering pro-
vided physils that are normally formed in warmer climates. In the Paleozoic
extensive kaolinite formation appears to be more closely related to the conditions
that favored the formations of coals, high humidity and abundance of organic
material, rather than directly to temperature.
624

Tonsteins
Tonsteins are mudstone beds in coal-bearing sequences, commonly within the
coal beds themselves. They range from less than a few centimeters to a meter thick.
Most are composed of well ordered kaolinite but I/S is common. Most tonsteins
apparently formed from volcanic ash. In most instances the volcanic ash fell into
coal swamps and under the acid conditions that existed altered to kaolinite;
characteristically books of vermicular kaolinite are present. In some instances the
volcanic ash altered to montmorillonite and I/S and is similar to K-bentonites. The
I/S tonsteins more commonly occur in the rocks associated with the coals rather
than in the coal beds.
In Scotland, Francis (1961), and in England, Price and Duff (1969), and
numerous others have identified altered tuff beds (up to 40 in Scotland) in the
Upper Carboniferous. The deposits in coal beds (coal tonsteins) are altered prim-
arily to poor to well crystallized kaolinite. Many of those in the non-coal beds
(non-coal tonsteins) altered to I/S similar to that in the Ordovician K-bentonites.
The former beds were deposited in acid swamp environments, whereas the latter
were deposited in shallow marine environemnts. Similar deposits, ranging in age
from Lower to Upper Carboniferous, are present throughout much of Europe
(Bouroz, 1966; Burger, 1964; Thorez and Pirlet, 1978; Williamson, 1961).
Based on a study of the mineral composition and trace elements, Spears and
Kanaris-Sotiriou (1978) established that most of the British tonsteins were derived
from locally derived basic volcanic ash, commonly mixed with detrital non-volcanic
sediment. They concluded that a few of the British tonsteins and many of the
European tonsteins were derived from acid volcanics, presumably related to
Hercynian orogenic activity.
A detailed study of a 1 m thick tuff bed in the Westphalian A/B of Upper
Silesian Coal Basin (Parachoniak and Srodon, 1973) demonstrates the nature of the
I/S tonsteins. The tuff can be traced over an area of 1000 km2 and occurs within or
close to a coal seam. The physils in the tonstein are kaolinite and I/S ( > 45%
smectite). As the tuff was deposited in an acid peat bog it is necessary to explain
why it was not altered entirely to kaolinite. The authors suggest that during a short
period following deposition of the tuff chemical equilibrium was not established and
the weathering tuff alkalized the adjacent water, allowing montmorillonite to form.
With time, equilibrium conditions returned, the pH decreased, and kaolinite formed.
The biotite altered to vermicular kaolinite and the feldspars and fine-grained matrix
to a finer kaolinite. In the 1 m tuff bed the amount of kaolinite increases as the
proportion of illite layers in the I/S increases. Parachoniak and Srodon suggest that
the I/S was formed from montmorillonite by the absorption of K released during
the kaolinitization of biotite and K-feldspar. The alteration to I/S occurred during
early diagenesis at near surface pressure and temperature. Apparently under these
same conditions Si was lost from the montmorillonite in order that the layer charge
could be increased (Al). The wet-dry mechanism (p. 180) may be the means by
which the illite layers formed.
Altered tuff beds are common in the Upper Carboniferous of Europe, but only a
few beds have been described in equivalent age rocks of North America. In the
625

United States volcanism was relatively active in the Upper Mississippian (upper
Lower Carboniferous) and was related to the collision of North and South America.
In Europe there was a major phase of tholeiitic lava extrusion in the mid-Lower
Carboniferous, approximately Upper Mississippian. This was followed in the mid to
late Carboniferous by granitic intrusions and with acid-to-intermediate explosive
volcanism (Windley, 1977). It was these latter volcanics that were altered to
tonsteins in the coal measures.

PERMIAN

Hercynian orogenic activity continued into the Permian. The assemblage of


Pangaea was completed in the Middle Permian when Siberia and Europe collided
(Uralian Orogeny) to produce the Ural Mountains (Fig. 9-20). Three major factors
affected a change in the biota (Axelrod, 1981) and also physil formation. As
Pangaea enlarged, the subtropical high-pressure cells strengthened and drought
increased during the warm season over middle to lower latitudes. Extensive redbeds
and evaporites developed throughout the world. The spread of redbeds coincided
with the rapid decrease of the typical late Carboniferous flora (spore-producing
trees) which were replaced by taxa with more-enclosed seeds and other adaptations
to periods of drought. Glaciation in Gondwana lowered sea level and the area of the
continental shelves was reduced. With the introduction of cold water into the world
ocean, climates became more zoned. The strengthened high-pressure cells increased
the areas of upwelling and cold water should have spread along some coastal areas.
Volcanism, accompanying plate collision and subduction as Pangaea was assem-
bled, was extensive. Episodes of explosive volcanism gave rise to sharp episodes of
chilling and further increased continentality (reduced rainfall, lower temperature,
greater extremes) and led to further ocean chilling. Faunal evidence indicates ocean
temperatures were lowered in the Permian.
Axelrod quotes Case (1926), “the environment determines what path life shall
take”. Likewise, the environment determines what path the physils shall take. The
well-documented change in climatic conditions between the Carboniferous and the
Permian is reflected in the physils.

North America

During the Permo-Triassic the eastern portion of North America continued to tilt
upward, as it had from late Paleozoic time. Terrigenous detritus encroached
westward and marine sediments were restricted to the western half of the United
States and an even smaller portion of Canada. Carbonates, evaporites, phosphorites
and red beds are the most common sediments. Volcanoes were abundant along the
western flank of North America and in northern Mexico. In general the region was
arid, but a few coal beds were deposited in the eastern part of the area.
The red pigment in most red sandstones is diagenetic hematite. Following
626

Fig. 9-20. Permian paleogeography. From Mintz, 1981.

depasition of non-red sands, detrital physils mechanically infiltrate the sands via
surface waters. The migrating ground waters alter the physils and other Fe-bearing
minerals (olivine, augite, biotite, etc.). These latter minerals frequently alter to an
Fe-bearing smectite. Fe is released and in the oxidizing-alkaline waters precipitates
as hematite, or as a precursor oxide which converts to hematite upon aging. The
infiltrated physils characteristically redden first. Continued alteration of the silicates
releases additional Fe, whch results in the formation of additional hematite, and
additional reddening (Walker, 1976).
In the United States there were three general types of source rocks supplying
physils to the western Permian basins: the uplifted Appalachia-Ouachita-Marathon
fold belt to the east and south containing largely sedimentary and low-grade
627

metamorphic rocks; interior uplifts containing granitic and high-rank metamorphic


rocks; and volcanic rocks along the western and southern flanks.
During the Early Permian thin interbedded limestones and shales were deposited
along the eastern flank of the epicontinental sea (Nebraska, Kansas, Oklahoma).
The physils are largely illite with lesser amounts of chlorite and kaolinite (Cubitt,
1979). The source was believed to be the Ouachta Mountains to the south. Slightly
to the west, north-south studies of a thin shale bed (Lee and Chaudhuri, 1979) and a
limestone bed (Twiss and Stindl, 1975), extending from Nebraska to Oklahoma,
indicate the physil suite is more complex. In general, illite and I/S are present
throughout the section. Kaolinite is more abundant to the south, nearer the source,
and Ch/V to the north in the more marine environment. The physils are primarily
detrital, but the Ch/V may have formed authigenically from volcanic detritus,
which increases to the west. Illite is the dominant physil in the shelf carbonates of
southeast New Mexico; chlorite increases westward and was presumably derived
from the Front Ranges.
In central Kansas the Lower to Middle Permian sediments are largely arkosic red
beds with thin evaporite deposits. The physils are largely illite and chlorite (some
may be Ch/S). Sediments were derived largely from the Front Range to the west.
Montmorillonite is abundant in the Middle Permian (Guadalupian), presumably
derived from volcanic activity in northern Mexico (Swineford, 1955). Farther west
evaporite deposits increase in abundance and Ch/V and illite are the major physils
(Holdoway, 1978). Some corrensite is present (Kopp and Fallis, 1974). The physils
in the Permian evaporites are discussed in detail in a few pages.
Farther west the physils in the red arkosic sediments along both flanks of the
Ancestral Rockies were apparently similar to those in the thick Pennsylvanian
section, largely I/S and illite (Raup, 1966). In the continental red beds in northeast-
ern New Mexico, derived from the granitic rocks of the Sierra Grande Uplift, illite
and montmorillonite are the dominant physils; minor I/S is present (Palmer, 1981).
The physils in these red beds indicate weathering conditions were mild and the
climate arid to perhaps temperate.
In the Cutler Formation continental red beds of Utah (Paradox Basin), illite,
with minor smectite, is the dominant physil; chlorite and kaolinite are relatively
abundant (Bodine and Rueger, 1982). North, along the western edge of the craton
(Montana), phosphate and chert rocks were deposited, presumably related to
upwelling currents. Kaolinite and illite are the dominant physils in the siltstones;
both appear to be secondary (Weaver, 1955).
In southwest Texas (Val Verde and Midland Basin) the physil suite is similar to
that of the underlying Pennsylvanian rocks, illite predominants with varying amounts
of chlorite, kaolinite and I/S (Weaver; Speights and Brunton, 1961). The source was
presumably the Ouachita-Marathon foldbelt to the south. Illite, montmorillonite
and I/S all occur as predominant physils in the Delaware Basin limestones of West
Texas and New Mexico. Bentonite beds are abundant in the limestone member of
the Cherry Canyon Formation (Guadalupian) of the Delaware Basin. These beds
are interesting in their variety. The predominant physils in the various beds are I/S
(4:1), illite (lM), Mg chlorite, and illite and montmorillonite. These bentonite beds
628

occur over an interval of 50 m. In the Yates Formation (Guadalupian), corrensite is


commonly the predominant physil, particularly in the sandstones (Dodd et a/.,
1954; Earley et al., 1956). These studies not only indicate that there was abundant
volcanic activity in this area during the Permian, but further demonstrate that
volcanic ash responded to the chemistry of the depositional environment and altered
to a wide variety of physils.
The Permian Sea gradually regressed southwestward and by mid-Permian time
was largely confined to West Texas and southern New Mexico, where it occupied a
complex of basins (Dalhart, Palo Duro, Midland and Delaware Basins) separated by
highs. Evaporation exceeded precipitation and a thick series of evaporites was
deposited. The evaporites are primarily anhydrite and halite, though potash salts are
present in southeastern New Mexico. The evaporites were deposited largely in
shallow water sabkha-like environments.
A wide variety of physils occurs in the evaporites and associated rocks of the
Palo Duro Basin. The source of detritus was likely granitic rocks and soils to the
east and northwest and some volcanic ash. The predominant physil in the dolomites
and dolomitic mudstones is largely detrital illite (K-Ar). Minor chlorite is present.
The more evaporitic rocks, anhydrite and halite, contain primarily Mg-rich miner-
als, Ch/S, Ch/V, corrensite, Ch/swelling Ch, saponite, chlorite along with illite and
minor kaolinite. There appears to be some relation between physil suites and
environments but there is much overlap. Detrital, Al-rich physils, illite, I/S,
montmorillonite and kaolinte, were apparently partially altered to Mg-rich chloritic
physils in the hypersaline environments. The extent of conversion is probably
related to the length of exposure in the brine and/or the salinity of the brine
(Palmer, 1981; Hall and Weaver, 1985).
In the New Mexico portion of the Permian Basin, where potash salts are present,
the Upper Permian evaporites contain, in addition to the chloritic physils and illite,
serpentine, talc and mixed-layer talc-saponite. The latter two physils are restricted
to the potash zone (Bodine, 1978).
A word of caution: Though it is commonly stated that illite-mica is altered to
Ch/S or Ch/V in hypersaline brines, the 10 A peak is usually relatively sharp
regardless of the relative amounts of illite-mica and Ch/S, Ch/V. There is no
obvious evidence of degradation, peak broadening, formation of I/S or I/Ch, of the
illite-mica; however, EDX analyses of the chloritic material indicate it commonly
contains appreciable K, presumably present in interstratified layers of mica. This
would suggest the chloritic precursor was an illite or mica.
The physil suites in the Permian rocks of North America characteristically have a
low kaolinite content, which is in keeping with the generally arid climate that
prevailed. It is of interest to note that kaolinite, generally indicative of humid
conditions, is more abundant in the northern portion of the western United States
where the Permian equator was located.
629

Europe, Africa, South America

Except for the Late Permian when the hypersaline Zechstein Sea covered most of
the northern margin, most of western Europe was the site of terrestrial, largely
desert, redbed deposition. The Russian part of the craton remained flooded by an
epeiric sea and carbonates and evaporites were deposited. At the close of the
Permian the influx of clastics from the rising Ural Mountains forced the sea from
the craton. Lower Permian environments include alluvial fans, dunes, sabkha and
salt lakes. Locally, thin coal deposits are present. Volcanics are relatively abundant.
The climate was primarily warm and arid and, as in North America, relatively little
kaolinite was produced, though it is abundant as a postdepositional product in
many of the desert sandstones.
A few analyses from England indicate the major physil in the mudstones and
limestones is illite; minor chlorite is present. Kaolinite is the predominant physil in
most of the sandstones and conglomerates; illite, with or without I/S, is also present
(Perrin, 1971; Fisher and Jeans, 1982).
The physils in the North Sea mudstones are apparently largely detrital illite. The
sandstones contain a variety of diagenetic physils. Illite is most abundant and
kaolinite usually the second most abundant. I/S and chlorite are less common
(Fisher and Jeans, 1982; Lee et al., 1984). K-Ar apparent ages of the illite fall in the
range of 100 to 174 m.y. This suggests the illite formed during the Kimmerian
movement whch caused the development of the graben system in the North Sea
area (Lee et al., 1984).
In the Paris Basin the primary physil is illite (70-80%), with a lesser amount of
chlorite. K.I. values indicate the illites are of anchimetamorphism grade but because
of the shallow depth of burial the K.I. values probably indicate the illite is detrital.
In the Sarre et de Lorraine illite, with minor I/S, is predominant (diagenesis grade).
Kaolinite is relatively abundant near the Permo-Carboniferous boundary, where
coal beds are present (Dunoyer de Segonzac, 1969).
The Upper Permian Zechstein evaporite deposits of Germany contain a typical
diagenetic, Mg-rich physil suite (Braitsch, 1971). Talc is the predominant physil in
the center of the basin. Chlorite, corrensite and muscovite (2M) tend to be most
abundant in rocks from the edge of the basin. Detrital muscovite-illite predominates
in the carbonate rocks and shales; chlorite is relatively abundant in the latter. Talc,
illite, chlorite, montmorillonite, Ch/S, I/S, V/S and Ch/V are present in varying
amounts in the anhydrite rocks. Illite and chlorite, with minor montmorillonite, are
the predominant physils in the salt rocks (Fuchtbauer and Goldschmidt, 1959). In
Zechstein zone Z corrensite is abundant in the halite beds and chlorite and
muscovite in the potash beds (Braitsch, 1960).
Illite, montmorillonite and kaolinite are all relatively abundant in the rocks of the
Russian Platform (Vinogradov and Ronov, 1956; Vlodarskaya, 1962). Much of this
material was presumably derived from the Urals. The presence of volcanic frag-
ments suggests some of the montmorillonite was derived from volcanic ash.
A study of rocks from the Preduralsk foredeep, deposited under contrasting
climatic conditions (Kossovskaya, 1969), demonstrates the role of climate in the
630

formation of physils. The major physil in the red bed-evaporite physilite deposits is
corrensite and Ch/S (8:2). In the coal measures, deposited under humid conditions,
the major physils are montmorillonite and chlorite; kaolinite is abundant in the
organic rich sediments. With depth montmorillonite and kaolinite are converted to
I/S (20% S).
An interesting aspect of this study is that the physils in the physilites are the
same as those in the associated sandstones (graywackes). The physils in the
graywackes are authigenic, which suggests those in the physilites are also authigenic,
rather than detrital. This is reasonable. The source rocks were volcanic rocks of the
greenstone complexes in the Ural Mountains. Where the source rock is largely
volcanic material, it might be expected that a major portion of the physils in the
depositional environments, both in sandstones and in physilites, would form authi-
genically from the detrital volcanics.
The Permo-Triassic of northwest Africa is discussed in the Triassic section. The
physils in the Permian marine geosynclinal sediments in the Karroo Basin of South
Africa show the effects of burial diagenesis (Rowsell and De Swardt, 1976). Illite
and chlorite (ratio 3 : l to 1:l) are generally predominant. I/S and kaolinite are
present in varying amounts. Kaolinite is abundant in some of the shales associated
with coal seams. To the north, in the zone of diagenesis, illite and chlorite comprise
< 70% of the physil suite; I/S and /or kaolinite are present. To the south, in the
anchizone, the illite and chlorite generally comprises 100%of the physil suite.
In the southeastern portion of the Parani Basin, Brazil, the physil suite changed
from predominantly kaolinite (70 to 90%) in the Upper Carboniferous to predomi-
nantly smectite in the Permian, approximately 80%; minor illite, kaolinite and
chlorite are present. The change represents a shift from a cool, humid climate to a
warmer, and presumably less humid, climate (Elizalde and Steinberg, 1973). Kaolinite
is abundant in the coal bearing Rio Bonito Formation (Bossi and Lenzi, 1981).
To the north, the physil suites in the shallow marine black to gray shales are
complex and variable. Illite, I/S, montmorillonite, chlorite and corrensite occur as
predominant minerals. Kaolinite is conspicuous by its absence (Ranios and For-
moso, 1976). The extreme variability indicates the physils are largely detrital, with
the possible exception of corrensite, and weathering in the source area was mod-
erate. In the continental deposits of Argentina illite and kaolinite are predominant
in the channel facies and smectite in the flood plain shales. The smectite was
apparently neoformed from pyroclastic material. Vermiculite and chloritized
montmorillonite occur in paleosoils (Ovejero and Bossi, 1974).

TRIASSIC

Sedimentation during much of the Triassic was similar to that in the Permian;
many rock sections, particularly the continental deposits, are referred to as Permo-
Triassic. Pangaea was fully assembled. The sea continued to withdraw from the
supercontinent, and an increasing proportion of sediments was deposited in arid
631

TRIASSIC

Fig. 9-21. Triassic paleogeography. From Mintz, 1981.

continental environments. Volcanism occurred in both western and eastern North


America, northeastern Africa and, to a lesser extent, northern Europe. Red beds
were abundant and the area covered by evaporites was at a maximum. Some coals
(humid climate) were deposited at high latitudes. The climate of the mid-latitudes
was warm but seasonably wet (Parrish el al., 1982).
During the Late Triassic, rift faulting and basic igneous activity occurred in
eastern North America and adjoining northwest Africa. Rift faulting occurred in
western Europe but volcanic activity was minor. Fault basins developed in eastern
North America and were filled with arkoses and basalts. The development of this
rift zone was the prelude to the opening of the North Atlantic, which occurred in
the Jurassic.
632

North America

Schultz (1963) analyzed 1200 samples of Triassic rocks of the Colorado Plateau,
Utah, Colorado, Arizona and New Mexico. The Lower Triassic Moenkopi Forma-
tion is composed predominantly of reddish siltstones, with a lesser amount of
sandstones and limestones, which were deposited primarily in flood-plain and tidal
flat environments. The source was to the east (Ancestral Rockies), and marine rocks
increase to the west.
Illite (with 10% smectite layers) and I/S (20 to 30% S) are the dominant clays in
the Moenkopi rocks. Minor amounts of chlorite (dioctahedral) or kaolinite are
commonly present; the distribution of the two is mutually exclusive. Chlorite is
present in the eastern fluviatile rocks, with kaolinite being present in the tidal flat
and marine rocks. Apparently the chlorite weathered to kaolinite, probably in the
tidal flat environment; however, it is possible the kaolinite altered to the Al-rich,
dioctahedral chlorite in the presence of alkaline pore waters. Dioctahedral chlorites
are relatively rare in the geologic section. Ch/S (some is corrensite) occurs almost
exclusively in sandstones and appears to have grown there, probably altering from
detrital physils.
The remaining Triassic section is represented by the Upper Triassic, Chinle
Formation, which is separated from the Moenkopi by a major unconformity
(Middle Triassic). The rocks, mostly siltstones and sandstones, are primarily reddish
and entirely continental in origin. Kaolinite and, to a lesser extent, I/S are
abundant in the lower portion of the Chinle. Some of the kaolinite is detrital and
some formed by weathering at the site of depositions. Apparently for a short period
of time conditions were relatively humid.
Montmorillonite and I/S (with a high content of smectite) increase upward in the
section and from north to south as kaolinite decreases. Dioctahedral chlorite
increases to the east. Much of the montmorillonite formed from water-transported
tuff particles (from the south) at the site of deposition. The montmorillonite occurs
in the sandstones and I/S in the claystones. Schultz suggested that the I/S formed
from volcanic material in the source area and was transported as clay-sized material
and concentrated in the claystones. Two types of volcanic tuffs are present: latitic
pebbles (K-poor) which altered to montmorillonite and rhyolite pebbles (K-rich)
which altered to I/S ( - 7:3). This strongly suggests that K-rich volcanic tuff can
alter directly to I/S without first altering to montmorillonite, as is commonly
assumed. The presence and retention of K in the environment of alteration,
relatively dry, may be the critical factor. The concentration of I/S in the claystones
may indicate that rhyolite tuff or ash, which should be finer-grained than the less
viscous latite, was concentrated in the claystones.
Near the middle of the Chinle Formation palygorskite is present. The palygors-
kite formed in alkaline desert lakes, presumably from tuffs or montmorillonite.
In the upper portion of the Triassic section illite and I/S are the dominant
physils in the northern part of the plateau. These illitic rocks grade and intertongue
southward with rocks rich in montmorillonite. In the northern area minor chlorite is
present to the east and kaolinite to the west, as in the lower Triassic.
633

A more detailed x-ray analysis of 61 Upper Triassic Chinle Formation samples


from northeast Arizona and southeast Utah (Schrock, 1975) shows that, in addition
to discrete illite, most samples contain a major amount of I/S. The I/S physils
contain from 15 to 80% illite, with most samples containing 50 to 80% illite layers.
The I/S ratio varies widely within individual outcrops, indicating it is detrital or
derived from montmorillonite and/or volcanic material which responded to minor
differences in the chemical environment. Minor kaolinite is present in most samples.
Three samples contain pure montmorillonite and four from one outcrop contain
ordered Ch/S, with 55% chlorite layers. Dolomite is present in the samples
containing Ch/S and is absent in samples from the same outcrop that contain I/S.
Evaporitic conditions favor the formation of Mg-silicates.
Throughout much of the Triassic sediments and physils were supplied from two
main source areas. A southern source supplied volcanic debris, which, under arid
conditions, altered to montmorillonite and I/S, probably both at the source and in
the depositional environment. An eastern source in the northern part of the area
supplied illite, I/S and dioctahedral chlorite, presumably from granitic and gneissic
rocks. Locally, secondary Ch/S formed in porous sands and dolomitic rocks,
apparently from chlorite, I/S, or volcanic material. Palygorskite is present in some
alkaline lake deposits and is associated with calcite. The climate was generally arid
and weathering moderate; however, during the mid-Triassic hiatus the climate was
more humid and considerable kaolinite was formed, both in the source areas and at
the depositional sites.
Schultz also x-rayed paleosoil samples developed on Precambrian crystalline
rocks underlying the Chinle Formation. Though the rocks have a widely diverse
composition (granite, monzonite, gneiss, schist, and amphibolite) the physil suite is
relatively uniform. The dominant physil is I/S (20-30% S); minor amounts of
poorly-crystallized kaolinite and illite are present. It is of interest to note that under
these arid climatic conditions both the acid crystalline rocks and rhyolites altered to
I/S. This is further evidence that much of the illite and I/S in the Precambrian
Lower Paleozoic rocks may have had a similar origin (pedogenic).
In the continental Pop0 Agie Member of the Chugwater Formation (Upper
Triassic) of west-central Wyoming, illite, presumably detrital, is the dominant physil
is the lacustrine carbonate rocks. Montmorillonite, I/S, and analcime, formed from
volcanic detritus, are abundant in lacustrine deposits (siltstones) with a low carbonate
content (High and Picard, 1965).
Detrital illite (1Md) is the dominant physil in the Upper Triassic of central New
Jersey and adjacent Pennsylvania. Chlorite is present in minor amounts and
palygorskite, I/S, and talc in trace amounts (Van Houten, 1965).
A thick section of Triassic sediments was deposited in the Arctic Sverdrup Basin.
Analyses of Upper Triassic samples from four wells show a relatively uniform physil
suite composed of 70% I/S (- 4:1), 20% illite, 5% chlorite and 5% kaolinite. The
physil suite reflects the moderate climate that presumably existed in this area, a
considerable distance north of the equator.
Preceding the rifting that separated North America from Africa, doming and
tensional forces produced a series of elongate Late Triassic graben basins along the
634

east coast of North America. This was accompanied by abundant basaltic igneous
activity. The rocks are primarily red continental rocks, ranging from mudstones to
conglomerates. Some thin dark shales and coals occur in the center of the basins.
Sections are from 1000 to 7000 meters thick. Source rocks are primarily gneisses,
schists and granites. Basaltic dikes and sills are abundant and in part served as
source rocks.
The physil suites of the rocks in the southern basins are complex and variable.
Illite and montmorillonite are the most abundant physils. Either can be the
dominant physil. Varying amounts of chlorite, kaolinite and vermiculite are present
(Hooks and Ingram, 1955; Weaver). The abundance of illite and montmorillonite
reflects the presence of the metamorphic and basaltic rocks on the flanks of the
grabens. A 900 m core from the Dunbarton Basin, South Carolina, contains a
relatively uniform physil suite consisting of illite (lMd), I/S (7:3-4:1), chlorite, and
corrensite (Lewis, 1974).
The physils in the red mudstones of the Late Triassic New Haven Formation of
the Connecticut Valley, Connecticut, are I/S (10-15% S), illite and montmorillonite
in various proportions. A study (Vergo and April, 1982) of the physils in contact
aureoles adjacent to basaltic dikes indicates that hydrothermal fluids rich in Fe and
Mg converted the I/S to a regular Ch/S ( - corrensite) (Fig. 9-22). The Fe,O,
content of the bulk rock adjacent to the dikes is approximately twice ( 5 vs. 10%)
that in the unaltered rock. Mg increases by a factor of approximately 5 (0.4 vs.
2.0%). The altered zone is on the order of 1 to 2 m wide. Other studies indicate
Ch/S is the most common physil produced when dikes intrude physilites.
Mg-rich physils, Ch/S, chlorite, sepiolite and palygorskite are apparently not
abundant, or present, in the Triassic red beds of eastern North America, whereas
they are common in the equivalent beds of Europe and North Africa, which were
only “slightly” to the east of the North American deposits. This is in keeping with
the general lack of evaporite rocks in the Triassic rift basins. Though the sediments
are red, the climate was apparently not particularly arid and the waters were not
highly alkaline. Rainfall must have been moderate but not enough to allow
appreciable kaolinite to form. In addition, in the narrow troughs the water depths
were probably greater than in the shallow lakes and epicontinental environments in
which the “evaporitic physils” formed.

Europe

By the end of the Permian much of western Europe and North Africa were land
areas. A multi-directional rift system developed that, except for North Africa, was
accompanied by little volcanism. The intensive erosion of the Permian and Lower
Triassic greatly decreased the relief and produced a semi-peneplane surface over
much of the area. The shallow high salinity Zechstein Sea occupied much of
Germany and northern Europe. To the southeast, a deep ocean (Tethys) occupied
the eastern part of the present-day Mediterranean. The two seas were not connected
635

CS
147

1
2 10 20

28 DEGREES
Fig. 9-22. X-ray powder diffraction patters of the oriented < 2 p m fraction of red mudstone samples
taken in traverse across a contact aureole. The clay-mineral assemblage consists of interstratified
chlorite/smectite + mixed-layer illite/smectite + chlorite. The unaltered red mudstone sample contains
mixed-layer illite/smectite + illite + chlorite (CuKa radiation). From Vergo and April, 1982. Copyright
1982 The Clay Miner. SOC.

and both periodically inundated the platform areas of western Europe, producing
sediments that are referred to as Germanic and Alpine facies (Ager, 1980).
The climate was warm to hot and apparently had alternating wet and dry
seasons. In the source areas micas and chlorite were degraded but seldom converted
to kaolinite. Feldspars may have altered to illite and I/S. Evaporation exceeded
636

precipitation and in the epicontinental platform hypersaline conditions prevailed;


gypsum and salt were precipitated in varying amounts. The geography was such that
evaporitic conditions prevailed much of the time but the influx of detritus was more
than in a “normal” closed salt basin. The ion concentration, relative to detrital
physils, was not always enough to cause the complete alteration of physils to the Mg
silicates that occurs in end-phase evaporite deposits.
The Lower Triassic, Buntsandstein, is characterized by the abundance of coarse
clastics along the edge of the basins and red mudstones farther from shore;
environments ranged from continental to hypersaline. During the Middle Triassic,
Muschelkalk, the Tethys invaded, much of the platform area and carbonates were
deposited. Westward and northward evaporites and dolomitic mads were deposited.
During the Upper Triassic, Keuper, a regression caused the termination of carbonate
sedimentation and a clastic-evaporite sequence was deposited. Shallow seas and
delta complexes, immense tidal flats, lagoons and sabkhas occupied much of
western Europe. Detrital physils were strongly altered and authigenic (neoformed)
physils were abundant.
Throughout the Triassic of the Euro-African area illite is commonly a major
component. It is predominant in the lower clastic facies and is replaced upward, in
the more chemical facies by varying amounts of Ch/S, corrensite, swelling chlorite,
chlorite, palygorskite and sepiolite.

England
In South Devon, England, the Lower and Middle Triassic rocks (New Red
Sandstone) are primarily basin edge fluvial conglomerates, sandstones and mud-
stones. Illite is by far the most abundant physil. Minor kaolinite and chlorite are
present in most samples. I/S, swelling chlorite and montmorillonite are present in a
few samples. Kaolinite is predominant in a pebble bed but most of it is believed to
be secondary (Henson, 1973).
The Sherwood Sandstone Group consists largely of red, fluviatile sands. Feldspar
and igneous and metamorphic rock fragments are abundant. During shallow burial,
some of the unstable silicates were dissolved and I/S crystallized; in the more
deeply buried sections the I/S was converted to illite. Following uplift acidic
groundwaters altered some of the feldspar to kaolinite (Burley, 1984).
The fluviatile deposits grade upward into a mixed marine and hypersaline facies
(Keuper). For much of western Europe two main facies are recognized. The Alpine
Facies (Tethys) contains open marine sediments and the physils are primarily
detrital. The red bed Germanic Facies contains continental and hypersaline sedi-
ments (mudstones, carbonates and evaporites) and an abundance of neoformed or
transformed Mg-rich physils. Fig. 9-23 is a schematic model of the depositional
facies-physil suites for these two facies. The Keuper Marl is primarily a lacustrine
deposit (Klein, 1962).
In south Devon the Keuper Marl contains evidence of three transgressions of the
Alpine Facies into the hypersaline Germanic Facies (Jeans, 1978; Fisher and Jeans,
1982). Two physil suites of different origins are present. The detrital suite is
composed predominantly of illite and minor chlorite. Superimposed upon this
631

GERMANIC FACIES red bed - ALPINE


FACIES
Continental Marginal Hypersaline open marine
hypersallne I

Wtrital kaolinlte j+halinlte I, I

7 zssso1uiloli7 -
unggmdoticn+ Detritalmica ond chlorite
I
I
I Detrital clays
;of d e p d e d mica I
I I
' Glaucmite
I
I
I
I
I
I

I
I
-
Sepiolite
I

Smxtite,smectite-mw I
-'1
I I I
I I I Neotormed
I II Smectite-chlorite I clays
I
II
I I Corrensite
I I
I I
I
I I Chlorite
I
I
I I I
I I I
I
I
I
I - I
Calcite
I I I
I I
1 - Dolomite 'J
4
I
I I
Evaporite
I
I I minerals
-um.anhydrit- I
1
I
! I
I
I
I
I
,
I
< Halite -> I
I - 1
Fig. 9-23. Schematic model of the depositional facies-clay assemblage relations for the Germanic red bed
Facies, Great Britain, and the adjacent open marine Alpine Facies of the Western European Trans. From
Fisher and Jeans, 1982. Copyright 1982 Miner. Soc.London.

matrix suite is a variety of authigenic or neoformed physils: sepiolite, palygorskite,


smectite, I/S, Ch/S, corrensite and chlorite. There are many exceptions to the
idealized physil-rock type associations. In general, the lateral variation in physils
parallels the lateral zonation of evaporite minerals. Neoformed sepiolite, palygors-
kite and I/S occur in calcite and dolomite, Ch/S occurs in gypsum, and corrensite
and chlorite in halite (not present in Devon section). The relation is better
demonstrated when the regional distribution is considered. For example, though
there are few neoformed physils in the upper part of the Dover section, farther to
the north (Midlands), where the salinity was presumably higher, swelling chlorite is
present.
Jeans (1978) believes the various neoformed physils resulted from the chemical
reactions occurring between normal sea water and hypersaline waters and the rapid
removal of Mg from solution by the precipitation of carbonate minerals. He believes
the physils precipitated directly from solution, with no physil precursor. This origin
may be possible for sepiolite (Mg, Si), but all the other neoformed minerals should
contain appreciable Al. It is unlikely that enough Al, in solution, was supplied to the
Triassic basins to allow the formation of an appreciable amount of neoformed
physils.
638

Scattered analyses of Keuper Marl samples (mostly marl and mudstone) from the
Midlands (England) indicate illite is the most abundant physil in all samples. Minor
chlorite is present in most samples. Most samples contain either corrensite, Ch/S,
or swelling chlorite (commonly about half abundant as illite). Sepiolite and paly-
gorskite are present in a few samples (Dumbleton and West, 1966; Davis, 1967;
MacNeill, 1978).
The Keuper in an offshore well 500 km southwest of the Devon coast consists
largely of red mudstones with minor sand, anhydrite and halite. As in England, illite
is the most abundant physil; in many samples illite is the only physil. Detrital
chlorite (relatively Fe-rich) is present in minor amounts. Neoformed corrensite and
Mg chlorite are present in the upper part of the section (Fisher and Jeans, 1982).
The latter sequence was apparently deposited under more saline conditions than the
Devon sediments.

Germany
As in western Europe in general, illite, commonly degraded, is abundant
throughout the Triassic section. Abundant corrensite is present in the red shales in
the upper part of the Lower Triassic (Lippmann, 1956) and in the Upper Triassic
Keuper (Echle, 1961; Schlenker, 1971; Lippmann, 1976). The physils in the more
marine Middle Triassic sediments are largely illite and minor kaolinite. Kaolinite is
abundant in the Upper Triassic of northern Germany, apparently derived from the
Bohemian Massif (Storr, 1975).

France
In France, the most complete Triassic section occurs in Lorraine, near the
western border of the Germanic Sea. With time the sea transgressed to the west,
gradually increasing the size of the Paris Basin (Lucas, 1962). The lower part of the
Triassic section contains fluvial conglomerates and sandstones; these grade upward
into a marine-evaporate sequence containing limestones, dolomites, marls, shales,
anhydrite and halite.
The physils consist largely of degraded illite (I/& I/V, I/Ch) and Ch/S.
Degraded or “open” illite is defined by Lucas as material which has a broad 10 A
peak and is a mixed-layer physil with at least two types of interlayers (Fig. 9-25).
The majority ( - 80-95%) of the layers are 10 A micaceous layers. The minor
component consists of 14 A layers, smectite, vermiculite or chlorite (identified on
the basis of their reaction to ethylene glycol and heating). There is a gradual change
in the physil suite in all sections. Degraded illite is the predominant, and commonly
the only, physil in the lower sandstones; minor degraded chlorite and kaolinite are
present in some intervals. The proportion of 14 A chloritic minerals increases
upward in the section (Fig. 9-24), comprising as much as 80% of the physil suite,
though in the uppermost Keuper degraded illite is the only physil. The 14 A phase is
primarily Ch/S. Near the center of the basin corrensite and Mg-chlorite are present
in the upper part of the section.
The Upper Triassic and Low Jurassic transgressive sequence on the northeastern
border of the Paris Basin contains a detrital physil suite consisting predominantly of
639

Fig. 9-24. Triassic section of the Pont-i-Mousson borehole, Meurthe-et-Moselle, France, showing upward
increase in chloritic physils. From Lucas, 1962. Copyright 1962 MBm. Sew. Carte GCol. Alsace Lorraine.

illite (both well crystallized and degraded) with varying amounts of kaolinite,
chlorite, smectite, and I/S (Muller et al., 1974).
In the Jura Basin, eastern France, the Triassic thickens from 100 m in the west to
over 1400 m in the east. Along the western edge of the basin the Lower and Middle
Triassic sediments are largely detrital sands derived from the Massif Central
(granites, gneisses and schists); the Keuper consists of marl and anhydrite. East-
ward, towards the center of the basin, carbonates and evaporites (halite) increase in
abundance (Lucas, 1962; Lucas and Ataman, 1968). The eastern edge of the shallow
basin was open to the Tethys Ocean. The abundance of evaporites indicates a
640

1 0 14 A
Fig. 9-25. X-ray diffraction pattern of detrital Triassic clays from the Paris Basin. Samples contain
degraded illite and poorly crystallized chlorite. From Lucas, 1962. Copyright 1962 MCm. Serv. Carte
Gkol. Alsace Lorraine.

barrier separated the shallow basin from the open ocean.


Detrital, degraded illite (Fig. 9-25) is the predominant physil in the eastern
sediments, deposited near the source area. Minor, poorly-crystallized kaolini te and
chlorite are present. To the north, in the Vosges Mountains, illite is commonly the
only physil present in the Lower Triassic sandstones. Basinward from the western
shore, the sediments become less sandy and the character of the physils gradually
changes. Progressively, the 10 A x-ray peak broadens towards the low angle. A
diffuse reflection appears between 10 and 14 A, and then a distinct peak near 14 A
(Ch/S) develops. With increasing distance the mixed-layer Ch/S becomes more
regular, and near the center of the basin corrensite is present. This material in turn
is transformed into a mixed-layer chlorite-swelling chlorite and eventually a well-

-
crystallized Mg-chlorite. The sequence is: degraded illite and chlorite + random
Ch/S + corrensite + random Ch/swelling Ch chlorite. Paralleling this sequence,
the broad 10 A peak becomes sharper and resembles that of a well-crystallized illite.
Corrensite is most common in the evaporites and chlorite in the clay beds. Locally,
attapulgite is present. Degraded illite is commonly the only physil in the uppermost
Triassic.
These samples have not been deeply buried, and Lucas believes the changes are
due to transformations of physils in the depositional environment rather than
diagenesis. Weathered illites may have been stripped of their K and retained a
relatively high layer charge or the chemical alterations could have affected the basic
lattice and reduced the layer charge to that of a montmorillonite. During evapora-
tion the Mg concentration of the shallow marine waters increased and Mg hydrox-
ide islands and layers grew in the open illite layers. At the same time the highly
charged open layers “fixed” K and were reformed into illite. The seaward increase
641

in the perfection of the chlorite layers is apparently related, in part, to a decreasing


amount of physil material. As the proportion of physils in the sediments decreases,
relatively more Mg is available per physil particle and the transformation can be
more complete. Though the MgO content remains relatively constant from the edge
to the center of the basin, the MgO content of the clay fraction increases from 5 to
14%. Lucas noted that in dolomites, where much of the Mg is tied up in the
carbonate minerals, chloritization is retarded. Chlorite is best developed in physilite
beds, where it did not have to compete with the carbonate minerals for Mg.
It is likely that many of the chloritic physils in the Triassic evaporite rocks of
Europe and North Africa originated in this manner; however, keep in mind that
chloritic physils can form from a variety of starting materials, e.g., montmorillonite,
volcanic glass, chlorite, biotite.

Switzerland
In the Molasse Basin of Switzerland, immediately east of the Juras, the Upper
Triassic, Keuper consists of a more shoreward facies consisting largely of red
alluvial fan deposits (clays, marls, sandstones and dolomite). The physils in the
lower part section are almost entirely illite (1Md) and in the upper part I/S ( - 7:3)
(Frey, 1970). Farther east, in the Glarus Alps, the Triassic has been subjected to
anchimetamorphism and the illite and I/S altered to phengite and a lesser amount
of Al-rich (30 to 35% A1,0,) chlorite. The chlorite is most abundant in the upper
section (originally I/S) and was formed because not enough K was available for the
conversion of all the I/S to phengite. Both 2M and 1Md phengites (illite?) are
present. In the more metamorphosed epizone only the 2M polymorph is present.

Other Areas
Kaolinite is abundant in the coastal sandstone facies along the northern boundary
(southern Sweden) and eastern boundary (western Poland (Krumm, 1969) of the
marine basins, and as far south as the Szeged Basin, Hungary. In the Lower Triassic
red sandstones of Hungary the kaolinite is due to the post-depositional alteration of
feldspars. The marine sandstones contain secondary chlorite (BCrczi, 1972).
Well-crystallized illite and chlorite are the dominant physils in the gypsum and marl
beds of the Lower and Middle Triassic of northern Hungary. To the south the physil
suite is composed of illite, I/S and montmorillonite. The Upper Triassic coal beds
contain illite, kaolinite, chlorite and chamosite (Viczihn, 1975).

Spain
The lithology and mineralogy of the Triassic rocks of Spain are very similar to
those in France and England. The lower portion consists of fluviatile pebble beds
and sandstones. The middle section, Muschelkalk, contains limestones and dolomites.
The Upper Triassic, Keuper, contains thin limestones, marls, shales, gypsum and
some salt.
In the Asturian Basin (northwest Spain) the Lower Triassic is composed largely
of illite (60 to 100%).Minor kaolinite and degraded chlorite are present, particularly
near the edges of the basin. Upward, in the Lower Keuper, Ch/S is abundant (60 to
642

80%) and in the Upper Keuper well-crystallized Mg-chlorite is abundant (70 to


90%). Montmorillonite and sepiolite are locally abundant in the upper part of the
section (Caballero and Martin Vivaldi, 1973).
In the Triassic belt along the southern slope of the Pyrenees, illite, slightly
degraded, is the only or major physil. Minor chlorite is present in some samples.
Upward, in the carbonate-evaporite facies, Ch/S is present. It increases in abun-
dance upward, and in the Keuper, chlorite with a few swelling layers is the
dominant physil (Lucas, 1962). In the Iberian Ridge area, in east-central Spain, the
physil sequence is similar to that in the northern areas, though chlorite is less
abundant and Ch/S generally more abundant (Caballero and Martin Vivaldi, 1973;
Palacio et al., 1977).
Along the northeast coast of Spain, Catalan Coastal Chain, the distribution of
the various physils is generally more erratic than in the inland basins. lllite is
commonly the most abundant physil. The 14 A phase consists of Ch/S (frequently
with smectite predominant) and montmorillonite. No chlorite was observed. In the
shoreward side of the Betic Range (southeastern coast), illite is abundant, com-
monly 70 to 80%,throughout the Triassic; chlorite ranges from 5 to 20%. Ch/S and
montmorillonite are present in the Middle Triassic (Muschelkalk) and increase in
abundance in the Keuper (40 to 50% and 10 to 4076, respectively). Basinward
(southeast), in the Keuper, montmorillonite decreases and Mg-chlorite (20 to 50%)
and Ch/swelling chlorite (10 to 20%) become relatively abundant. Illite ranges from
70 to 80%. Minor sepiolite is found in some localities (Caballero and Martin
Vivaldi, 1973).
The sequence is similar to that in France and England. In general, the Lower
(Buntsandstein) to Lower Middle (Lower Muschelkalk) Triassic is characterized by
detrital illite and poorly crystallized chlorite, the upper Muschelkalk and Lower
Keuper by transformed mixed-layer chloritic physils and the Middle and Upper
Keuper by well-crystallized chlorite and minor sepiolite. Caballero and Martin
Vivaldi (1973) found that the physil sequence does not necessarily represent a
basin-edge to basin-center sequence as suggested by Lucas and Ataman (1968) and
Fisher and Jeans (1982) (Fig. 9-23). Ch/S can be abundant in the Lower Triassic
coastal, detrital sediments. However, these Ch/S physils need not be transformed. It
seems quite likely that in the nearshore areas some detrital Ch/S (weathered
chlorite) would be present. They note that montmorillonite can also occur in the
nearshore area and is probably detrital rather than authigenic. Sepiolite develops in
areas of high alkalinity and ionic concentrations. In shallow basins with localized
high and low areas such environmental conditions can develop in areas other than
the center of the basin.
This appears reasonable. Mg-rich physils, either transformed or authigenic
(neoformed), would not be expected to show a consistent trend from the edge to the
center of a complex basin but, instead, to reflect local chemical conditions in the
basin, which need not be related to distance from shore, except in a very general
manner.
643

Africa

The Upper Triassic, dolomitic black shales of Sicily contain a complex physil
suite. I/S (2:l) is commonly most abundant (20 to 50%); significant amounts of
montmorillonite, illite and kaolinite are present; chlorite is a minor component. This
appears to be primarily a detrital physil suite (Long et al., 1964).
The physil suites in the Triassic of Morocco (Lucas, 1962) are similar to those in
western Europe, though the fibrous physils, palygorskite and sepiolite, are more
common. The High Atlas, in southern and southwestern Morocco, contain a rift
basin similar to those in the eastern United States. Red sandstones, marls, clays and
thin beds of gypsum and salt are present. To the west, in the Moroccan Meseta, a
thinner lagoonal-marine sequence was deposited. The base of the section is com-
monly sandstone which grades upward into red clays, containing beds of salt,
gypsum and carbonate rocks. Numerous basalt flows are present in both areas.
Well-crystallized illite and chlorite are the major physils in the High Atlas. At
least some of the chlorite is Mg-rich. Illite commonly comprises 50 to 100%of the
physil suite. The chlorite content is more variable (trace to loo%), but it commonly
comprises 50 to 70% of the physil suite. Ch/S, with varying ratios, is present in
about 20% of the samples. The illite and chlorite are believed to be detrital and to
have been transformed into well-crystallized material in the depositional environ-
ment. The high chlorite content may be related to the presence of basalt flows and
the availability of Mg leached from the basalt. In the eastern High Atlas and to the
northeast, montmorillonite is relatively abundant.
To the west and northwest, in the coastal Mesetas, the thin Triassic section is
characterized by the presence of fibrous physils. Gypsum and halite are relatively
abundant. The illite content of the physil suites is usually in the range of 30 to
1008, being more abundant in the lower part of the Triassic. The chloritic material,
10 to 508, is commonly some variety of Ch/S and/or swelling chlorite. Montmoril-
lonite is locally abundant. The section is divided into a lower and an upper section
separated by a basalt flow. Sepiolite is present in the lower section and palygorskite
in the upper, commonly in amounts of 50 to 80%. The distribution of these two
authigenic physils suggests there was a subtle change in the water chemistry. The
occurrence of palygorskite above the basalt suggests it could have formed from
montmorillonite; sepiolite is more likely to grow directly from solution (rich in Mg
and Si).
Fiberous physils have not been found in northeastern North America, whch was
adjacent to northwest Africa during Triassic time. It is likely such deposits are
buried under the Continental Shelf.
South of the High Atlas, in the Polignac Basin, the physils in the predominantly
shale section are primarily illite and I/S, present in approximately equal amounts,
with 10 to 30% chlorite (Dunoyer de Segonzac, 1969). Approximately 20% kaolinite
is present in a well close to the High Atlas. In contrast to the physils in Morocco,
the physil suite appears to be detrital and to have been relatively unmodified in the
depositional environment. However, in the El Gassi region of the Sahara, degraded
illite, predominant in the Lower Triassic, is progressively chloritized upward, with
644

chlorite being the predominant physil in the upper part of the section where salt is
abundant (StCvaux, 1961).
The major physils in the Triassic of South Africa are illite (commonly 50 to 90%)
and chlorite. These rocks are from the anchizone. Some I/S is present in samples
from the less metamorphosed areas.

Israel and Sinai

The Triassic sediments of southern Israel and Sinai were deposited near the
southeast margin of the Tethys Sea, with the continent extending to the southeast,
into the Arabo-Nubian massif. The lower portion of the section consists of con-
tinental sandstones. These grade upward into a shallow marine-lagoonal sequence
containing limestones, mark, dolomites and gypsum. The section contains several
basic sills (Heller-Kallai et a[., 1973).
Kaolinite is predominant in the continental sandstones in the lower continental
rocks and in some of the hgher, shallow marine deposits. Detrital illite (degraded)
and I/S (7:3) are the predominant physils in the lagoonal-shallow marine sequence.
Chlorite and Ch/V are abundant in some of the littoral and lagoonal sediments.
They are present in permeable sandstones and limestones and occur as authigenic
overgrowths on quartz grains and replacing muscovite, illite and feldspar. Both the
chlorite and Ch/V are Fe-rich, in contrast to the Mg-rich chloritic physils in the
Eurafrican deposits. The localization of the chloritic material in the porous rocks
and the composition (Fe-rich) suggest it was developed after burial, in contrast to
the Mg-chlorites which appear to be somewhat contemporaneous with deposition.
The gypsum beds contain primarily degraded illite and apparently no authigenic or
transformed minerals. Even though evaporation reached the stage where gypsum
was precipitated, no Mg-rich physils crystallized. Basically this indicates a lack of
Mg. Heller-Kallai et al. (1973) note that there is no chlorite in the Paleozoic and
Precambrian source rocks. This suggests that the evaporation of normal sea water
does not necessarily produce the Mg concentrations necessary to allow the forma-
tion of Mg-physils. Perhaps, in order for Mg-physils to grow, it is necessary that the
relative concentration of Mg be increased by the influx of fresh waters from source
areas containing reasonably soluble Mg (chlorite, biotite, basic volcanics, dolomite).

South America

During the Triassic, arid conditions existed in the area of the Paranh Basin and a
thick section of eolian sands were deposited. Large masses of basic lavas were
extruded. Montmorillonite is the dominant physil, presumably derived from the
lavas (Ramos and Formoso, 1976). In the red fluviatile, gypsiferous sediments of
west-central Argentina, chloritized montmorillonite or swelling chlorite (Mg-rich) is
abundant. Conditions were alkaline and the climate alternating dry to humid.
During the humid periods, weathering occurred and Al, Mg and Fe were mobilized.
645

During the dry seasons the Fe precipitated as Fe oxides and the A1 and Mg as
hydroxides between the montmorillonite layers (Bossi, 1972).

Comment

Illite (and degraded illite) is by far the most abundant physil in Triassic
sediments. The illite is primarily detrital. The abundance of slightly weathered illites
and chlorites and the scarcity of kaolinite and detrital vermiculite confirm other
lines of evidence that the Triassic was predominantly a dry period, but generally
with some seasonal rainfall. A high rate of evaporation, as indicated by the
abundance of evaporates, is confirmed by the abundance of Mg-rich silicates. This
situation was best developed in western Europe and northwest Africa where a
shallow, epicontinental sea transgressed over an irregular topography. Evaporation
of sea water produced the conditions that allowed the formation of chloritic
minerals from detrital illite, chlorite and montmorillonite. Palygorskite and sepiolite
are believed to have formed as a result of evaporation of brackish or lacustrine
water. In some areas the formation of abundant Mg silicates was facilitated by the
presence of basic volcanics.

JURASSIC

Beginning in the Triassic tensional forces along the plate margins of Laurasia and
Gondwana produced numerous rift valleys. One of these rift valleys was torn open
in the Early Jurassic to produce the incipient Atlantic Ocean (Fig. 9-26). Initially the
Atlantic Ocean rifting occurred along a nearly east-west line. Basaltic magma welled
up through fractures formed by rifting and intruded the Triassic rift valleys. Waters
from the Tethyan Sea flowed into the narrow North Atlantic where it evaporated
and formed thick deposits of evaporites. Similar evaporites formed in what is now
the Gulf of Mexico from waters that encroached from the Pacific Ocean. Though
the Jurassic climate was relatively warm, as attested to by the abundance of
evaporites, it was cooler than during the Triassic. Following the breakup of Pangaea
a general cooling trend started that has persisted to the present.

North America

During the Late Triassic and Early Jurassic a large area of central western United
States was covered with a thick deposit of quartz-rich sandstone (Navajo). These are
considered to be either dune or shallow marine and beach deposits. Beginning in
Middle Jurassic time, the sea advanced~fler the western craton from the Arctic and
northern Pacific regions. In this epikontinental sea a wide variety of carbonates,
shale and sandstones were deposited; evaporites were deposited in the lower part of
646

the section. The western portion of the craton, California, was an area of major
volcanic activity.
One of the most extensive formations in the central western United States is the
Morrison Formation. The detritus was derived from an arcuate source extending
from the southeast to the west and northwest. Granites and rhyolitic tuffs were
abundant in the source area. The depositional environments were largely alluvial
and paludal-lacustrine (Cadigan, 1967). On the Colorado Plateau, the predominant
physil in the lower part of the formation is illite (poorly crystalline, some smectite
layers), which is believed to be detrital (Keller, 1962). The illite facies interfingers
with facies rich in montmorillonite and I/S. Minor kaolinite is present in a few
samples. Chlorite and I/Ch are widely distributed in the sandstones. The upper
641

portion of the Morrison is characterized by a predominance of montmorillonite in


the northwestern part of the Colorado Plateau. Illite, chlorite and I/S increase
toward the south. Keller believed the illite and I/S were detrital and the montmoril-
lonite was formed from volcanic ash.
The lower portion (18 m) of a core of the Morrison Formation near Dillion,
Colorado, consists of a marl that contains illite as the only physil (Wahlstrom,
1966). The physils in the overlying 50 m of mudstone and sandstone, both calcare-
ous and noncalcareous, consist of minor kaolinite and abundant I/S. The per-
centage of illite layers in the I/S is quite variable, ranging from 50 to > 90%. In
general, the proportion of illite layers increases upward in the section. Altered
rhyolitic ash is abundant in the upper half of the section, and the mudstones are
composed primarily of altered volcanic ash fragments ranging from 0.03 mm to clay
size. Volcanic ash was not observed in the lower part of the section, but the
mudstones contain many rounded sand-sized physil particles. These grains are
apparently derived from volcanic ash that altered to physil material before deposi-
tion in the mudstone. As these physil grains are in the lower part of the section they
are presumably composed of illite. Wahlstrom observed that, in the upper section,
veins and fracture fillings of secondary, well-crystallized illite (“sericite-like”) occur
in the rhyolitic mudstones. In the published thin section pictures, the illite appears
to occur in clay skins lining soil peds. That these mudstones may have been “soils”
is further suggested by the fact that the mudstones break into angular and blocky
fragments rather than shaly chips.
In the Black Hills region of Wyoming and South Dakota, illite is the predomi-
nant physil. Kaolinite is abundant, and montmorillonite rare. All three physils are
considered to be authigenic but may not be (Tank, 1956).
Thin bentonite beds, derived from rhyolitic ash, near the base of the Carmel
Formation (Middle Jurassic) in southwest Utah, contain varying amounts of 1 M
illite, I/S (some of the smectite is chloritic) and kaolinite. The physils in the
adjacent sediments are similar, except for the presence of 1Md illite. The detrital
sediments include a large proportion of rhyolitic material, and it is likely that,
though the 1Md illite is detrital, the other physils were formed from the volcanic
material, as were the physils in the bentonite beds (Schultz and Wright, 1963). The
Carmel sediments were deposited in a tidal flat environment and the more shore-
ward bentonites, to the east, contain more kaolinite. The variation in composition of
the bentonites, illites, montmorillonites and chlorites, is not obviously related to
geographic position on the tidal flat. The differences may be due to a difference in
environments within the tidal flat or a difference in the chemistry of the ground or
connate waters.
The Jurassic rocks of the Colorado Plateau present convincing evidence that
rhyolitic material can alter to illite and I/S in surface environments. The environ-
ments in the Colorado Plateau, in which rhyolite altered to illitic material. ap-
parently ranged from alluvial to paludal-lacustrine to tidal flat to soils. Though it is
generally assumed that volcanic material must first alter to montmorillonite and
then the montmorillonite alter to I/S and illite, this does not appear to be a
necessary sequence. If K can be retained in the environment volcanic glass can
648

apparently alter directly to I/S-illite. A semiarid environment, alternate wet and


dry, with a high content of easily soluble, K-rich volcanic material, would appear to
be the ideal environment for I/S-illites to crystallize. The AI/K ratio of rhyolites is
within the range of most illites.
At the northern edge of the midcontinent basin, Saskatchewan, montmorillonite
and I/S, with a high smectite content, are the most abundant physil in the Middle
and Upper Jurassic. Kaolinite and illite increase to the north, shoreward (Weaver).
The Jurassic in several wells from the Sverdrup Basin, Arctic, contains a physil
suite composed predominantly ( - 60%) of randomly mixed-layered 1/S ( - 7:3), 20
to 30% illite and roughly equal amounts of kaolinite and chlorite. There is nothing
distinctive about the physil suite, which is probably detrital.
During the Early and Middle Jurassic, evaporites (Louann Salt) were deposited
in the basin of the incipient Gulf of Mexico. I have no information about the physils
in the bedded deposits; however, samples from salt domes (Louisiana), derived from
the Louann Salt, contain well crystallized 1M illite (K.I. = 1.5), pure talc, and
mixtures of Mg-chlorite and illite. Illite is the primary physil in the overlying
Smackover and Buckner limestones; chlorite ranges from a trace to 30%.
Along the northeast coast of North America rift valleys continued to develop.
The sediments are primarily red fluvial and black to gray lacustrine sediments
interlayered with basalt flows. The East Berlin Formation (145 to 450 m) of the
Connecticut Valley is sandwiched between two lava flows. Illite is the dominant
physil in the fluvial red mudstones, along with minor chlorite (April, 1980). The
physil suites in the lacustrine mudstones are more complex (April, 1981). The
various lacustrine units consist of alternating gray mudstones and black shales. The
black shales contain saponite and illite, along with varying amounts of I/S and
minor chlorite. The physils in the gray mudstones are predominantly 1Md illite and
Mg-rich chlorite, with varying amounts of saponite, except in the gray mudstones
immediately overlying the black shales. These latter mudstones contain a well-crys-
tallized, regularly interstratified Ch/S (corrensite), along with subordinate amounts
of 1Md illite and chlorite.
The black shales have a high content of MgO (13.7%), which is present in
dolomite, magnesite and saponite. It is likely that the saponite was formed authi-
genically in reasonably fresh lake waters. The sedimentary structures and the
presence of dolomite, analcime, gypsum and halite molds in the upper gray
mudstones indicate it was deposited under evaporitic conditions in a shallow
alkaline lake or pond. The accompanying high pH conditions apparently favored
the development of Ch/S rather than saponite. April suggests that, though Ch/S
started to form in the lacustrine muds, the complete transformation to corrensite
probably occurred during burial, when Mg-rich pore waters were expressed from the
underlying black shale during compaction.
The oldest sediments along the western edge of the Atlantic Ocean are Middle to
Late Jurassic age (Hollister et al., 1972). Site 100 was drilled at the northern edge of
the Bahama Platform. The hole bottomed in basalt. The basalt is overlain by 43 m
of greenish-gray limestone containing plant fragments. This grades upward into a
red, clayey limestone approximately 32 m thick. The depositional environment is
649

considered to be upper bathyal. The physils (Weaver) in the section immediately


overlying the basalt consist of I/S with a ratio of 1 : l . The latter phase produces a
broad 10 A peak and no 5 A peak and is presumably Fe-rich. It may be a form of
glauconite or celadonite. The upper portion of the greenish-gray limestone contains
I/S (2:3), relatively well-crystallized illite, kaolinite and palygorskite (near the top).
The red limestone contains I/S, with ratios ranging from 1:4 to 3:2, illite and
palygorskite, with traces of kaolinite. Palygorskite is the dominant physil in some
samples. Though there is no real evidence, except for the presence of kaolinite, the
physil suite is probably detrital.
The Jurassic was also cored in Hole 105, due east of Cape Hatteras on the
continental rise. The physil suite consists largely of illite and montmorillonite. The
former is predominant in the lower part of the section (Oxfordian) and the latter in
the upper part of the section. There is no information on the nature of the
montmorillonite (I/S?) (Zemmels et al., 1972). The physils would appear to be
detrital and reflect a source area in a more temperate environment than the source
area for the southern well (Hole 100).

Europe, Africa

During the Early Jurassic, rising sea levels caused a transgression of the Tethys
and Arctic seas over much of Europe. During the Middle Jurassic tectonic activity
increased in the Atlantic and Western Tethys rift systems, which resulted in crustal
separations and the onset of sea-floor spreading. This led to the lowering of sea
level.
Analyses (summarized by Perrin, 1971) of the Early Jurassic (Lias) of England
indicate illite (in some instances including I/S) is the predominant physil, com-
monly more than 70%.Kaolinite is commonly the second most abundant physil and
locally is predominant. Kaolinite, with a lesser amount of illite, is the dominant
physil in the lower part of the Middle Jurassic section. This grades upward into a
section dominated by montmorillonite. The latter interval contains commercial
fuller’s earth deposits (montmorillonite). Montmorillonite is present at scattered
localities, commonly in amounts of less than 15%,but in some intervals as much as
30% is present (Bradshaw, 1975).
There has been some discussion about the origin of the fuller’s earth deposits, but
the more recent studies indicate they were formed from volcanic ash (bentonites)
(Hallam and Sellwood, 1968; Bradshaw, 1975). Bradshaw noted the presence of
basalt in the North Sea and suggested this area was the source of the volcanics.
In Dorset the Lower Jurassic mudstones are composed of subequal amounts of
kaolinite and illite, with minor chlorite; sandstones contain an abundance of Fe-rich
chlorite. The physil suite was deposited in a nearshore environment. Illite is
dominant in the Middle Jurassic but montmorillonite is abundant in the upper
fuller’s earth interval (Knox, 1982). This sequence of physils appears to persist over
much of England. In southern England montmorillonite (I/S with 20 to 40% illite
layers) comprises 40 to 90% of the physil suite of the lower Upper Jurassic
650

(Oxfordian). Illite is the other physil present. The presence of zeolite and biotite
indicates the montmorillonite was derived from volcanic ash. The source is presuma-
bly the same as for the fuller’s earth montmorillonite (Chowdhury, 1982).
The Lower Jurassic shallow water limestones of northwest Scotland have a physil
suite containing approximately 70% montmorillonite and I/S which may have been
derived from basalts generated during the initial rifting in the North Atlantic. The
overlying, open marine sediments have an illite-rich (70 to 100%) physil suite,
indicating a change in source (Amiri-Garroussi, 1977). In northeast Scotland the
physil sequence is similar to that in the northwest. Montmorillonite and I/S (40 to
60% illite) are abundant (30 to 60%) in the lower part of the Jurassic and decrease
upward. Mite is dominant in the Upper Jurassic. Kaolinite, 20 to 30%. is present
throughout the section (Hurst, 1981). Thus, in northern Scotland montmorillonitic
physils are abundant in the Lower Jurassic, and in southern England in the Middle
Jurassic.
The Middle and Upper Jurassic physil suite in the North Sea is composed largely
of I/S (Pearson et a/., 1982). The I/S of the < 0.2 pm fraction in the shallower
buried Upper Jurassic sediments has a ratio of approximately 1:4. With depth,
burial diagenesis increases the I/S ratio to 4:l. The North Sea sandstones contain a
wide variety of physils, largely secondary: illite, kaolinite, dickite, chlorite and
Ch/V (Kantorowicz, 1984). For more information on North Sea sandstone, see
Chapter VIII.
On the Danish island of Bornholm the Jurassic consists of coal-bearing deltaic
and swamp deposits. The physil suite is composed of 10 to 30% kaolinite, 10 to 30%
I/S, 20 to 50% illite, about 10% vermiculite, and < 10% chlorite. In general the
amount of kaolinite and I/S increases upward towards the coal layers, suggesting
acid leaching (Graff-Peterson, 1961).
The Jurassic of the Paris Basin has a physil suite similar to that in England,
largely illite and kaolinite (Millot, 1949). In the north of France the kaolinite/illite
ratio decreases upward in the section. Montmorillonite comprised 50% of the physil
suite in the uppermost portion of the Middle Jurassic (Callovian) and the lower
portion of the Upper Jurassic (Lower Oxfordian) and remains relatively abundant
throughout the Upper Jurassic. The montmorillonite occurs in slightly younger
sediments than the fuller’s earth of England. Ziegler (1982) suggested that the
volcanism was probably related to wrench-faulting in the Bristol Channel area.
The carbonate rocks of central Spain contain a complex physil suite consisting of
varying amounts of biotite, montmorillonite (and goethite), with minor amounts of
kaolinite and sepiolite (Martin et al., 1976). Palygorskite is present in the Middle
Jurassic carbonate rocks of the Iberian and Betic Ranges. In some zones the pH
conditions became low enough, after deposition, that the palygorskite was altered to
smectite, and dolomite and chert were formed from the released Mg and Si (Bustillo
and Iglesia, 1978).
In the area of the southwest French Alps the Upper Jurassic (“terres noires”) is
progressively metamorphosed from west to east. In the western portion of the area
the physil suite is composed primarily of illite (1Md) and I/S, with minor amounts
of kaolinite and chlorite. Over an interval of approximately 120 km, eastward
65 1

towards the Alps, the kaolinite disappears, and the I/S is converted to illite and
chlorite. The final product is an anchizone slate in which the physil suite is
approximately 70% illite (largely 2M) and 30% chlorite. The same sequence of
mineralogic changes occurs with depth in individual wells, in the interval ranging
from Upper to Lower Jurassic (Dunoyer de Segonzac, 1969).
A similar sequence of diagenetic-metamorphic mineral change has been described
for the Lower Jurassic of Switzerland (Frey, 1970, 1974). The physil suite in the
northern portion of Switzerland, Tabular Jura and Molasse Basin, consists largely of
I/S (20 to 40%smectite) with lesser amounts of illite, chlorite and kaolinite (zone of
diagenesis). In the Glarus Alps the equivalent rocks have been subjected to
anchimetamorphism, epimetamorphism and higher. The physil suite has changed to
illite, chlorite, pyrophyllite (from kaolinite), mixed-layer paragonite/phengite and
paragonite. The epizone suite consists of phengite, paragonite and chlorite. The
paragonite is believed to have formed from the I/S.
The Jurassic of the Molasse Basin, between the Jura and the Alps of Switzerland,
consists largely of carbonate rock and a lesser amount of shale. X-ray analyses of
nearly 2000 samples from nine wells (Persoz, 1982) indicate the presence of a
complex variety of physil suites. Illite is the dominant physil but both kaolinite and
montmorillonite are present in amounts up to 80%.Significant amounts of chlorite,
I/S and I/Ch are present in many samples. Based primarily on kaolinite, illite and
illite peak sharpness, Persoz was able to identify 20 correlatable zones. The physils
are detrital and show no evidence of burial metamorphism even though the
overlying molassic series does. Persoz interprets this to indicate burial metamor-
phism of physils is inhibited in carbonate rocks. Maximum depth of burial ranges
from 1400 to 4172 m.
Though kaolinite is apparently a minor component of the physilites in southwest-
ern Europe, it is relatively abundant in north-central and eastern Europe. Kaolinitic
weathering crusts developed on the Bohemian Massif and areas of the Baltic Shield.
Evidence for the weathering crust is the abundance of kaolinite, up to 70%, in
sediments adjacent to the land areas. Kaolinite is most abundant in the Upper
Middle Jurassic sediments. In this area of Europe the climate was presumably warm
and humid during the Lower and Middle Jurassic and became progressively more
arid in the Upper Jurassic (Storr, 1975). A similar thick kaolinitic weathering crust
developed in northern Norway (Dypvik, 1979), in the Ural Mountains and in the
Ukraine (Petrov, 1958). The presence of abundant coal deposits in the Urals and
western Siberian platform indicates the climate in the region was moderate and
humid. Kaolinite deposits, with accompanying illite, are abundant in the Lower
Jurassic deposits of Romania. As in Russia, they are associated with a complex of
coal and sandstone beds. In the Schela-Viezuroiu region kaolinite has been meta-
morphosed to pyrophyllite (Neacp and Neacp, 1980).
Analyses of several hundred Upper Jurassic limestone samples from all over
Europe showed the presence of a number of provinces (Bausch, 1977):
1. Kaolinite-illite in most parts of the epicontinental shelf (France, Germany,
Dobrugea)
2. Chlorite-albite in some parts of Tethys (French pre-Alps and eastern Baltic
chains
652

3. Chlorite in northern Germany


4. Muscovite in the Iberian Peninsula
5. Kaolinite-albite in the eastern Alps
6. I/S in most parts of the Tethys (southeastern Europe)
The presence of albite is believed to indicate advanced diagenesis. The chlorite in
northern Germany indicates high salinity conditions. Muscovite suggests low
weathering intensity and kaolinite intense weathering. I/S may represent moderate
weathering.
In the Mecsek Mountains of Hungary, the Lower Jurassic coal-bearing sequence
contains a physil suite containing illite, I/S (high illite content), kaolinite and
chlorite. This interval is overlain by mads containing kaolinite, illite and montmoril-
lonite. In the Transdanubian Central Mountains and in the general region covered
by the Tethys, illite and montmorillonite are the predominant physils in the Upper
Jurassic (VicziAn, 1975).
In southern Israel the continental, littoral and lagoonal sediments contain various
mixtures of illite and kaolinite; offshore marine sediments contain smectite and
kaolinite in roughly equal amounts. The illite was derived from the Paleozoic
Nubian desert deposits. The kaolinite apparently altered from illite in fresh water
desert lakes, which had abundant vegetation (Bentor et al., 1963). The smectite in
the marine sediments may have been transported from a considerable distance by
marine currents.
In the Lower Jurassic limestones and shales of the Moroccan Basin de Guereif,
illite is the principal physil. Chlorite is abundant to the north and west and kaolinite
to the southeast (Porthault and Tixier, 1975). The Central High Atlas contains a
classical burial metamorphic sequence. The illite and chlorite show a systematic
decrease in peak width, which reflects a burial metamorphism intensity ranging
from diagenesis to epimetamorphism (Studer, 1980).
The Upper Jurassic claystones near LCopoldville and Brazzaville, The Congo,
have a physil suite composed primarily of montmorillonite, with minor amounts of
kaolinite and illite (Bartholome et al., 1963).

Comment

In North America and, particularly, in Europe there are significant differences


between the Triassic and Jurassic physil suites. The authigenic Mg-rich physils.
chlorite, Ch/S, palygorskite and sepiolite. which are relatively abundant in the
Triassic sediments, are relatively scarce, with the exception of Ch/S, in the Jurassic.
The physils suggest the climate was slightly cooler and more humid in the Jurassic
than in the Triassic.
In northeastern and eastern Europe the presence of abundant kaolinite indicates
the climate was more humid, and probably warmer, than to the west where kaolinite
is relatively scarce. This climatic difference may reflect the presence of the Tethys
Ocean fringing Eurasia (Fig. 9-26).
653

CRETACEOUS

Rifting continued throughout the Cretaceous. In the Early Cretaceous South


America and Africa began to separate, creating another narrow evaporite basin (Fig.
9-27). Eastern Laurasia and Africa converged and the Tethys Sea was restricted. By
the end of the Cretaceous, the South Atlantic had become continuous with the
North Atlantic, as the North American continents drifted westward. North America
had broken away from the remainder of Laurasia along a rift west of Greenland.
The eustatic sea level rise that started in mid-Jurassic time continued through
Cretaceous time, culminating in the Late Cretaceous with major encroachment of

CRETACEOUS

Fig. 9-27. Cretaceous paleogeography. From Mintz. 1981


654

the continents. In North America the sea advanced from both the Cordilleran region
and the Gulf of Mexico. The two seas joined in the Late Cretaceous, flooding the
entire Great Plains and Gulf Coastal Plain and the eastern edge of the continent to
southern New England.
Along the western coast of North and South America, Mesozoic Pacific litho-
spheric plates were subducted under the continents. Selective melting in the under-
thrust belt produced andesitic magma, which rose to the surface to form a volcanic
arc along the edge of the continents. Subduction began in Permian or Triassic time.
By Jurassic time the arc had been driven into the continent and a period of major
mountain building commenced, culminating in the Cordilleran orogeny which
spanned Late Jurassic through Early Cenozoic time. Uplift began near the Pacific
coast and moved eastward. Cretaceous sediments in the eastern Cordilleran coar-
sened and thickened westward, forming a large clastic wedge. Within this wedge the
present-day Rocky Mountains were initially uplifted in the Late Cretaceous (Lara-
mide orogeny). The flood of sediment from this uplift filled the seas, to the east, as
they retreated from the area. By Early Cenozoic time a series of low-angle thrust
faults carried immense slabs of rock eastward along a zone extending from Mexico
to northwestern Canada.
This orogenic activity was accompanied by extensive volcanism, which extended
from northern Mexico to Canada. Volcanism increased during the late Early
Cretaceous and reached a peak in the Late Albian. Volcanism continued into the
Tertiary and moved eastward as the orogenic activity extended eastward (Axelrod,
1981). In the western United States volcanic centers were concentrated in the
northern Rockies (western Utah, Montana, Canada and Idaho). Centers also
occurred in western Colorado, New Mexico, and Arizona (Rice and Gautier, 1983).
To the east volcanic activity occurred in the Mississippian embayment (Ross et al.,
1928). the Caribbean (Donnelly, 1973). offshore eastern Canada (Jansa and Pe-Piper.
1985) and the Mid-Atlantic Ridge. Though some volcanism occurred in the Early
Cretaceous, extensive volcanism, in both North America and Europe, started in the
late Early Cretaceous and continued into the Late Cretaceous. The late Early
Cretaceous volcanism was apparently related to an increase in spreading rates and
extensive marine transgression. The Upper Cretaceous sedimentary rocks of North
America and Europe are generally dominated by smectites derived from the
volcanic material. Lower Cretaceous rocks tend to have a relative high content of
illite and kaolinite.
By Late Jurassic time, the Appalachian mobile belt had been eroded to a
low-lying surface, and the present Atlantic and Gulf of Mexico basins had begun to
form. Sandstones and shales are the predominant sediments along much of the
Atlantic margin. Carbonate rocks are abundant in the southeast and the Gulf of
Mexico region.
The Cretaceous was a time of great warmth over the globe. During the warmest
time, tropical to subtropical conditions extended to at least 45"N and 70"s latitude.
The lower-latitude zone featured high aridity, except in western Europe. Tempera-
tures increased through about the first half of the period and cooled afterwards
(Frakes, 1979).
655

North America

Western North America


The westernmost area for which I have information is the Cretaceous physils in
the San Joaquin Valley, California. The clastic Upper Cretaceous was deposited on
basement rocks. A “granite wash” interval, immediately on top of the basement
rocks, has a physil suite composed predominantly of kaolinite, with varying amounts
of chlorite, illite and montmorillonite. The physil suites are extremely variable, but
various zones were identified in the Upper Cretaceous section (maximum thickness
- 1700 m). Montmorillonite ranges from 10 to 55%, illite from 0 to 50%, kaolinite
from 0 to 90% and chlorite from 0 to 25%. In addition, Ch/V is present in two
intervals. The vertical variations in the composition of physil suites probably reflect
changes in the intensity of weathering in the source area.
In Oregon, the physil suite of the Hudspeth Formation mudstones consists of
beidellite, vermiculite, mixed-layer illite-beidellite, illite, kaolinite and minor chlorite
(Jarman, 1973). Burial was deep enough to have converted beidellite to I/S and
illite in the more deeply buried parts of the section. Chlorite is relatively abundant
near intrusive bodies. The composition of the sandstones indicates the source area
to the north contained volcanic, low-rank metamorphic and sedimentary rocks. The
beidellite was presumably formed from the volcanic material, indicating relatively
mild climatic conditions.
Because of the presence of hydrocarbons and commercial kaolinite deposits, the
Lower Cretaceous Dakota group (which represents most of the Lower Cretaceous
time) has been more thoroughly studied than other Lower Cretaceous rocks. The
Dakota rocks extend from Utah to Kansas and from the southern United States to
well into Canada. Depositional environments are primarily shallow marine and
deltaic. Sediments were supplied from both the western and eastern flanks of the
basin.
In the foothills of the Front Range of Colorado, the Dakota sediments are largely
deltaic in south-central Colorado (Denver area), grading into marine sediments to
the north. The physils associated with the deltaic complex are largely kaolinite and
those in the marine sediments largely illite. Commercial kaolinite beds occur in
swamp deposits associated with the deltaic complex. The presence of illite in the
kaolinite suggests much of the kaolinite altered from illite. This is further suggested
by the presence of bentonite beds which are composed of kaolinite in the deltaic
complex and smectite and I/S in the marine sediments (Waagt, 1961).
East of the Front Range, in the Denver Basin, the distribution of the various
physils appears to be erratic, as might be expected in a shallow water basin filled
with clastic rocks. Illite, I/S (wide range of ratios) and kaolinite are the predomi-
nant physils; minor chlorite and vermiculite are present. The presence of I/S with a
high proportion of illite layers in shallow buried samples suggests much of it is
detrital. Episodic but frequent admixtures of I/S with a high proportion of smectite
layers indicates some bentonitic and/or volcanic material was supplied to the basin.
The effects of burial diagenesis is indicated by a statistical increase in the propor-
tion of illite layers in the I/S (Rettke, 1981).
656

Analyses of Dakota core samples from eight oil fields in Colorado and western
Nebraska show the physil suite is complex but relatively uniform. Illite, I/S (80 to
90% I) and kaolinite are relatively abundant. Varying amounts, usually minor, of
montmorillonite and chlorite are present in some samples. The only consistent
association is the relative abundance of illite in the continental shales and I/S in the
marine shales (Weaver). In western Kansas, the equivalent rocks contain abundant
montmorillonite (probably I/S) and illite. Shoreward, in central Kansas (Franks,
1966) and east-central Nebraska, kaolinite is the predominant physil. The kaolinite-
rich sediments were derived from a hghly weathered eastern source; the montmoril-
lonite, I/S and illite were probably derived from the western source.
North and west of the Denver Basin, in Wyoming, the siliceous Mowry Shale was
deposited, more or less comtemporaneously, with the Dakota Group sediments. I t
was deposited in a sea that advanced from the north, whereas the sediments to the
south were deposited in an extension of the Gulf of Mexico. Montmorillonite is the
dominant physil in the underlying Shell Creek Shale and the overlying Belle
Fourche Shale; however, I/S (minor and variable amount of illite layers) is the
dominant physil in the Mowry Shale. Minor illite is a common component (Davis,
1970).
The 1/S is detrital and was derived from a northwestern source. The presence of
abundant plagioclase feldspar in the shales suggests tuffaceous material was abun-
dant in the source area. Kaolinite is relatively abundant in the eastern part of the
basin and was presumably supplied by an eastern source. Early Cretaceous kaolinite
deposits, developed from the Precambrian crystalline rocks, are abundant in western
Minnesota (Parham and Hogberg, 1964). Non-siliceous shale beds in the Mowry
Shale commonly contain well-crystallized montmorillonite, suggesting they are
bentonitic. These shales commonly contain zeolites, as d o the overlying and underly-
ing montmorillonitic shales, tending to confirm the presence of a tuffaceous
material. Some of the siliceous cememt in the Mowry Shale was probably obtained
from the alteration of tuffaceous material, but most was derived from the dissolu-
tion of radiolarian tests.
I/S, with a high content of smectite layers, is apparently, the predominant physil
in the Early Cretaceous sediments of western Wyoming (Blatter et al., 1973). In the
overthrust belt of western Montana, Precambrian and Paleozoic rocks have been
thrust eastward of the Cretaceous. The smectites in shales and bentonites underlying
the thrusts have been exposed to burial temperatures ranging from 100" to 200°C
and converted to I/S with 60 to 100% illite layers (Hoffman and Hower, 1979).
Many of the bentonites in Montana, and throughout the Rocky Mountain region.
contain minor amounts of kaolinite. In some instances i t is pseudomorphic after
feldspar. The kaolinite presumably formed during the initial leaching period when
the ash was converted to smectite. Apparently Si was leached from the feldspars, or
other material, more rapidly than it was removed from the volcanic glass.
Montmorillonitic bentonite beds are common in the Mowry Formation (upper
Lower Cretaceous) and the lower part of the Frontier Formation (Upper Creta-
ceous) of Wyoming and Montana. A detailed study of the thicker beds in north-
657

central Wyoming (Slaughter and Earley, 1965) showed they have an elongate
east-west distribution. They thin from west to east, commonly from about 10 m to
< 1 m, indicating a western source. Most of the original volcanic ash was wind
transported; however, the presence of very thin intercalated shale laminations and
small scale flowage structures at the top of the beds indicates some water transport
during the final stages of deposition.
The ash was deposited primarily in marine and brackish environments. Zeolite is
commonly associated with the montmorillonite. The presence of fine-grained
montmorillonitic chert beds (porcelanite) on top of the bentonite beds and inter-
layered with thin bentonite layers indicates the excess silica released from the ash
migrated upward. It is not clear at what stage the silica was released from the ash.
Analyses of the sand fraction (phenocrysts) indicate volcanics ranged from rhyolites
to andesites, with dacites and latites being the most common. The volcanism is
believed to have been associated with the emplacement of the Idaho batholith.
Authigenic kaolinite is the predominant physil in the Cretaceous continental
sandstones in the Rocky Mountain foothills of Alberta and in the eastern plains.
Authigenic chlorite, illite and less commonly montmorillonite are present in sand-
stones containing volcanic detritus. Farther north in northeastern British Columbia
the 1400 m thick Bucking Horse Formation has a classic diagenetic sequence of
physils. The upper part of the section has a physil suite containing illite, chlorite,
kaolinite and I/S (1:4). With increasing depth of burial the kaolinite and chlorite
disappear and in the deepest part of the section only illite and I/S (1:l) are present
(Foscolos and Kodama, 1974). Kaolinite, illite and chlorite are present in the Lower
Cretaceous of northeastern Alberta (Carrigy and Kramers, 1973). In south-central
Saskatchewan the physils in the Cantuar Formation are closely related to deposi-
tional environments. Kaolinite is predominant in the deltaic sediments, montmoril-
lonite in the lagoonal and marine sediments and illite in the coastal beach sediments
(Weaver). Farther north, in the Mackenzie River Delta (northwestern area of
Northwest Territories) the Lower Cretaceous marine-deltaic rocks in one well
contain a physil suite composed of approximately equal parts kaolinite and illite
and a trace of chlorite (Bayliss and Levinson, 1970). In the Sverdrup Basin (Arctic
Ocean) the Lower Cretaceous contains approximately 55% montmorillonite-I/S,
20% illite, 20% kaolinite and 5 % chlorite (Weaver).
The Cretaceous climate over the total earth was milder and more uniform than
today. The climate in North America was largely subtropical. This is confirmed by
the relative abundance of kaolinite in the western and midcontent regions of North
America. The abundance of montmorillonite and I/S in some areas reflects the
local presence of volcanic material. The volcanic material either altered after
deposition in a marine environment, or in low-lying areas where weathering was
restricted.
The Mancos (west) and Niobrara (east) shales were deposited during the early
stage of a major Upper Cretaceous transgression. Bentonite (tonsteins) beds are
present in the Mancos coal-bearing deltaic deposits of central Utah. Most of the
bentonite beds are composed of well-crystallized kaolinite; some contain apprecia-
ble amounts of well-crystallized montmorillonite (Ryer et al., 1980). The Mancos
658

shales and bentonites of the Colorado Plateau are composed primarily of


montmorillonite and I/S (Nadeau and Reynalds, 1981). The I/S ratio varies from
1:9 to 9:l and generally increases with increasing burial depth and geothermal
gradient. I/S shales generally contain more illite layers than do associated bentonite
beds. This may be related to the availability of K or to the nature of the starting
material, glass in the ash beds and smectite in the shale. In some samples the
presence of calcite inhibited burial metamorphism. Thus, in a small outcrop
calcareous shales may contain pure smectite and noncalcareous shales I/S with
> 50% illite layers. K-Ar apparent age values of some marine I/S samples are
similar to the stratigraphic age indicating that the I/S probably formed from
volcanic material in the soil or sea shortly after deposition and not by burial
diagenesis (Nadeau and Reynolds, 1981). In the San Juan Basin of New Mexico
kaolinite is the major physil in the back-beach facies and I/S in the shales seaward
of the beach trend.
In the Washakie Basin of south-central Wyoming, as the last westward transgres-
sion of the Cretaceous sea was taking place (Lewis shale), the basin was being filled
with thick deltaic deposits from the west, northwest and southwest. Fig. 9-28 shows
the distribution of physils in a Washakie Basin well (typical of analysis of several
dozen wells). The Blair, Rock Springs and Erickson Formations (Mesaverde Group)
were deposited during a major eastern progradation. The Almond (beach-lagoon)
and Lewis (marine) Formations were deposited during a westward transgression of
the Upper Cretaceous sea. The Lance or Fox Hills Formation (beach-lagoon) was
deposited during a final regressive phase as the sea withdrew from the area.
There is no drastic change in the physil suite through the interval. I/S and illite
comprise the bulk of the physil suite. I/S and montmorillonite decrease with depth,
whereas chlorite and kaolinite (less systematic) decrease. These changes reflect a
change in source material. The source of detritus gradually shfted from west to
northwest. Detailed studies of the upper three formations showed the physils are
closely related to environments. Kaolinite is at a minimum in the marine shales.
Kaolinite increases shoreward and is at a maximum in the shales of the lagoonal-
marsh environment. Montmorillonite is most abundant in the marine shales, which
contain several bentonite beds. The boundary between the marine shales and the
deltaic wedge to the north is marked by the loss of montmorillonite. The montmoril-
lonite was derived from the western source, though some may have formed from ash
deposited in the marine environment (Weaver).
In Montana and the Dakotas the Upper Cretaceous Pierre Shale and equivalent
rocks were deposited in a regressive-transgressive-regressive sequence similar to the
equivalent age rocks in the Washakie Basin (Schultz et al., 1980). Near the western
source area (western Montana) the Pierre equivalent rocks consist of continental
sandstone and shale deposits, some with volcanic debris. These grade seaward into
marine sandstones and farther east into shales and marlstones. Despite the range of
environments the physil suites are relatively uniform (1220 samples).
I/S is the dominant physil (average 71%) in most samples. It contains three types
of layers: illite, montmorilllonite and beidellite. The proportion of beidillite remains
relatively constant at about 35 + 10%. The proportion of illite and montmorillonite
659

JACK K N I F E S P R I N G S
SWEETWATER C O . WYOMING
KAOLINITE MONT ILLITE
50 ;
LANCE
FEET

1 ,0 00
7

LEWIS

2.000

ALMOND

3.000
ERICKSC

4,000

ROCK
SPRINGS

5,000

BLAIR

6,000

Fig. 9-28. Relative percentages of the various clay minerals in the < 2 micron fraction of cutting samples
from a typical late Cretaceous section, Washakie Basin, Wyoming. From Weaver, 1961. Copyright 1961
Wyo. Geol. SOC.

layers varies widely with illite usually dominant. Bentonite beds consist of more
than 90% smectite, which is composed primarily of montmorillonite interlayered
with a smaller amount of beidellite and minor illite. The montmorillonite in some of
the bentonites contains scattered islands of gibbsite and brucite (incomplete col-
lapse when heated at 300°C) (Schultz, 1963). Zeolites are present in a few of the
bentonites.
As shown in Fig. 9-29, the marine and nonmarine rocks have similar physil
compositions. The vast majority of the detritus came from the west. Most of the
660

BOFN NOFB FNOB


i II ill

t
2
w
u
K
w
k 0 50 100
Illitstype Beidellite-type Montmorillonite-type
LAYERS AS PERCENT OF MIXED-LAYER CLAY

Mixed layer Illite Chlorite Kaolinite

Fig. 9-29. Abundance of mineral sin the Pierre Shale and equivalent rocks related to environment of
deposition. Cumulative curves are for 836 marine and 1268 non-marine shale and siltstone samples. Solid
lines, marine samples; dotted lines, non-marine samples. Short vertical lines above curves indicate the
arithmetic mean content of 496 offshore-marine (0). 340 nearshore-marine (N). 39 brackish-water (B).
and 229 freshwater (F) samples. From Schultz et ab, 1980.

illite, chlorite and kaolinite came from the erosion of older rocks. The marine and
nonmarine I/S are similar suggesting that they formed by weathering of the
volcanic debris on land and were little changed in the marine environment or after
burial. In the overthrust belt (westernmost Montana) burial metamorphism in-
creased the proportion of illite layers to 60 to 8048, interlayered mostly with
beidelli te.
Chlorite is abundant in both the marine and nonmarine volcanoclastic siltstones
of western Wyoming. In the volcanoclastic (andesitic) sandstones from the same
area, neoformed corrensite is the only physil present in the fresh water deltaic
distributary sandstones, and montmorillonite is the only physil in the marine
sandstones. Both physils have a high Fe content, -
20% and 10% Fe,O, respec-
tively. A study of the fluid inclusions in calcite indicates a temperature of formation
of - 50°C. Thermodynamic calculations suggest the corrensite formed under hypo-
saline conditions and the Mg/Ca ratio (10 to 50) was higher than in sea water
(Almon et al., 1976).
Along the eastern flank of the sea, in western Kansas, the Sharon Springs
Member of the lower portion of the Pierre Shale is composed mostly of marine
shales. The physil suite is composed primarily of I/S; lesser amounts of illite,
kaolinite and chlorite are present. Bentonite beds contain kaolinite and beidellite
661

with A1 hydroxide in the interlayer position. The bentonite beds were presumably
subjected to acid conditions at some stage of their development (Gill et al., 1972).
On the basis of a regional analysis of the non-physil minerals, Jones and Blatt
(1984) concluded that " two-thirds of the Pierre Shale samples are heavily influenced
by atmospheric contributions of volcanic ash." The conclusion is based on the clay
size of much of the quartz and an increase in the quartz/feldspar ratio with distance
from the volcanic source. This would imply that much of the I/S is formed in the
marine environment. The similarity in I/S content and composition in the marine
and nonmarine Pierre Shale suggests it is unlikely that a major protion of the I/S in
the marine shales formed from air transported ash. However, Schultz et al. (1980)
noted that the shales overlying bentonite beds were more montmorillonitic than
other shales, indicating the influx of ash did not stop abruptly after the ash beds
were deposited. The ash in the shales may have been water transported.
The physil suite of the marine Upper Cretaceous Wapiabe shales of western
Alberta, Canada, consists of illite and kaolinite. In addition to these two physils, the
overlying marine Belly River shales contain varying amounts of I/S and either
vermiculite or smectite (Campbell and Lerbekmo, 1963). Authigenic kaolinite and
chlorite are present in the continental sandstones. Chlorite is present in the
sandstone containing volcanic detritus. To the east montmorillonite is present in
altered volcanic fragments (Carrigy and Mellon, 1964). Kaolinite and illite, in
varying proportions, are the major physils in the fluviatile Athabasca Group
sediments of northeastern Alberta (Holve et al., 1981).
In one well in the Mackenzie River Delta the physil suite consists of approxi-
mately equal parts kaolinite and illite and minor montmorillonite and chlorite
(Bayliss and Levinson, 1970). In other wells in the area I/S is the predominant
physil (50 to 70%). With depth (3000 to 4000 m) the proportion of illite layers
increases from about 30% to 50% to 75% (Weaver) Kaolinite comprises approxi-
mately 70% of the physil suite of the Upper Cretaceous rocks on Ellesmere Islands
(Arctic). Minor amounts of illite, I/S and smectite are present (MacBustin and
Bayliss, 1979). Kaolinite is abundant throughout the mid-latitudes and as far north
as 80°N latitude. This is in keeping with the idea that the Cretaceous was a time of
great warmth over the globe. Tropical to subtropical conditions extended to at least
45"N; temperate climates extended almost to the poles (Frakes, 1979). Kaolinite is
generally more abundant in the more northern latitudes, suggesting the climate was
more humid than in the lower latitudes. This is confirmed by the relative abundance
of coal beds in the northern latitudes. Bertherine is believed to be the major physil
in the 3,000 to 4,000 m thick shallow marine to brackish-deltaic sediments of the
Canadian Archipelago (Masood et al., 1986).

Coastal Plains
The southern and eastern coastal areas of North America contain of a wedge of
Mesozoic and Cenozoic sedimentary rocks that dip gently towards the Gulf of
Mexico or the Atlantic Ocean. In Mexico, the western Gulf Coast, and Florida
carbonates are predominant. Sandstones and shales predominate on the Atlantic
662

Coastal Plain. A major reef band extends along the east coast of Mexico, along the
northern coast of the Gulf of Mexico and along the outer edge of the Atlantic coast
continental shelf. The climate in the Gulf Coastal Plain was tropical to subtropical
and relatively and.
Data on the physils of the Lower Cretaceous of the Gulf Coast are sparce and
there is only a minor amount of data on the Upper Cretaceous. Near Dallas, Texas,
the physil suite of the lowermost Upper Cretaceous Woodbine Formation deltaic
complex contains an average of approximately 28% illite, 36% kaolinite and 36%
I/S (40 to 50% smectite). Kaolinite is relatively more abundant in the fluviatile
facies of the deltaic complex (Hawkins et al., 1974). Montmorillonite, probably I/S,
is relatively abundant in the Upper Cretaceous of central Texas. The major physil in
the Taylor Formation is montmorillonite; kaolinite is relatively abundant in near-
shore facies. Montmorillonite is the dominant physil in the Austin Chalk. In east
Texas the Eagle Ford Formation I/S (40 to 100% smectite) is the predominant
physil (70 to 80%); varying amounts of illite, kaolinite and chlorite are present
(Weaver).
Lower Cretaceous shale samples from various wells in Louisiana and Mississippi
have highly variable physil suites. Illite is generally the most abundant physil but
I/S and kaolinite are predominant in some samples. The other suites consist
primarily of well-crystallized illite-muscovite with nearly as much kaolinite and
chlorite. The mica in the latter suite suggests metamorphic rocks in the Appa-
lachians were the probable source (Weaver). The Upper Cretaceous Tuscaloosa
Formation (sandstones) contains authigenic chlorite and kaolinite. Chlorite, as rims,
precipitated shortly after burial from intrastratal leaching of ultramafic and
volcanoclastic detritus. Kaolinite crystallized shortly thereafter. At a burial tempera-
ture of approximately 130°C kaolinite is no longer present and a second generation
of chlorite is developed, possibly from the kaolinite (Dahl, 1984). The Fe/Mg ratio
decreases below a depth of 3300 m (Beskin, 1984).
During the Upper Cretaceous the Mississippi Embayment extended as far north
as southern Illinois. Kaolinite is the dominant physil in the northern deltaic
sediments and I/S ( - 1:l) in the outer neritic environment. Intermediate environ-
ments contain approximately equal amounts of kaolinite, illite and I/S (Pryor and
Glass, 1961).
Upper Cretaceous volcanoes have been identified along the western flank of the
Mississippi Embayment, and bentonite beds of Late Cretaceous age are present in
Arkansas, Oklahoma, Louisiana, Texas and Alabama. Kaolinite beds in southwest-
ern Arkansas are believed to have formed from volcanic ash (Ross et al., 1928).
To the east, clastic sediments were deposited in an arc extending from Alabama
through middle Georgia, South Carolina and eastern North Carolina. To the south
in Florida a thick sequence of carbonate sediments was deposited (Fig. 9-30).
Illite-mica is the predominant physil in the Lower Cretaceous carbonate residues;
minor amounts of I/S and chlorite are present. The suite is similar to that of the
Lower Cretaceous of Mississippi and probably all the eastern Gulf Coast. Sharp-
peaked 2M mica is present in the lower part of the section; peak width increases
upward in the section. Kaolinite is abundant in the lower portion, 170 to 270 m
663
HUMBLE COASTAL P . HUMBLE GULF GULF
CARLTON N 0 . 1 T L S N O . I C O L L I E R NO.1 8 2 6 - G NO.l 3 7 3 NO.1

~SMECTITE BRUCITE 15455

Fig. 9-30. North-south cross section showing the distribution of the clay mineral suites in the Cretaceous
carbonate section of southern Florida. Clay mineral analyses were made using insoluble residue fraction.
From Weaver and Stevenson, 1971, Geol. SOC.Amer. Bull., 82,3457-7460.

thick, of the Upper Cretaceous. This interval is presumably equivalent to the


kaolinite-rich Tuscaloosa Formation to the north.
The increase in the width of the l O A illite peak upward in the section,
culminating in the deposition of kaolinite, suggests a metamorphic source, southern
Appalachians, with a systematic increase in the intensity of weathering with time.
Montmorillonite and I/S (4:l to 3:2) are the predominant physils in the Upper
Cretaceous rocks overlying the kaolinite-rich zone. This indicates a decrease in
weathering intensity and the presence of more volcanic material in the source area.
The increase in I/S and illite from north to south suggests there may have been a
southern source area contributing detritus to the carbonate shelf (Weaver and
Stevenson, 1971).
The Tuscaloosa Formation (lower Upper Cretaceous) forms an arc of kaolinite-
rich sediments flanking the Appalachians and extending from Alabama, through
Georgia and South Carolina, into North Carolina. The sediments primarily uncon-
formably overlie metamorphosed crystalline rocks. The sediments consist of sand
and gravels with lenses and beds of clay. Updip the sediments are fluvial and
deltaic, grading south and east into marine deposits. The source was primarily
weathered crystalline rocks in the Appalachian belt.
In northwestern Alabama kaolinite is the predominant physil in the Tuscaloosa.
To the south and seaward in central Alabama montmorillonite is predominant.
Varying amounts of illite are present in both areas and illite and kaolinite are the
dominant physils in the overlying sediments (Clarke, 1965). In Georgia and western
South Carolina, lenses, a few centimeters to 20 m thick, of high purity, commercial
664

x
100

BLACK CREEK
80

60

40

20

0 10 20 30 40 50 60 70. 80
MILES DOWNDIP
Fig. 9-31. Change in the percentages of kaolinite, montmorillonite, and illite in the Middendorf, Black
Creek, and Peedee Formations of North Carolina coastal plains. Sediments grade from non-marine to the
left, updip, to marine on the right. From Heron and Wheeler, 1964.

kaolinite are abundant. Some lenses contain minor amounts of montmorillonite


and/or illite. Montmorillonite is the dominant physil in the marine sediments. The
kaolinite-rich interval at the base of the Lower Cretaceous carbonate rocks of
Florida is presumably the time equivalent of the Tuscaloosa. Much, probably most,
of the Cretaceous kaolinite is detrital, derived from weathered crystalline rocks to
the north and northwest; however, the presence of vermicular kaolinite books
suggests that some of the kaolinite may have formed from feldspar after deposition.
Volcanic material from volcanoes in the Caribbean area may also have been a
source.
The abundance of detrital kaolinite in the middle Cretaceous rock of the
southeastern United States indicates the climate was probably warm and humid and
relief was moderate. In general this was a time of near-maximum temperature
during the Cretaceous (Frakes, 1979). The westward flow of the warm Tethys
currents and its continuation as the North Equatorial Current (Berggren and
Hollister, 1974) presumably provided the conditions necessary to produce relatively
high rainfall in the area.
In northern South Carolina and North Carolina kaolinite is the dominant physil
in the nonmarine sediments and montmorillonite and I/S in the marine sediments
(Fig. 9-31). Minor illite is present in all environments and glauconite in the marine
sediments (Heron and Wheeler, 1964; Heron, 1960). The same physil-environmental
relations exist to the north, in Delaware, New Jersey and New York. Minor chlorite
is present in the marine sediments (Groot and Glass, 1960). In most of Virginia and
parts of southern Maryland, montmorillonite is the predominant physil in the
nonmarine Potomac Group (Moncure and Force, 1976).
665

While it is evident that the physils in the continental and deltaic sediments in the
Atlantic Coastal Plain were derived from the Piedmont and Appalachian Moun-
tains, the origin of the montmorillonite in the marine rocks is less clear. During the
opening of the Atlantic Ocean extensive volcanism occurred during the late Lower
Cretaceous extending into the Upper Cretaceous (Jansa and Pe-Piper, 1985). In
some way these volcanic rocks were presumably the source of the montmorillonite.
Some volcanic material could have altered to montmorillonite directly in the marine
environments or it could have altered on the land and been transported to the
marine environment. Volcanic ash should erode more easily than normal soils and
could have been transported before it altered to kaolinite. Another possibility is that
the montmorillonite was from the lower portion of the soil profile, where weathering
was less intense, and that size segregation occurred during transport. The coarse
kaolinite would have been deposited in the high-energy shoreline environments and
the fine-grained montmorillonite transported to the marine environments.

Europe

In Europe, as in North America, there was an Early Cretaceous regression during


which much of western and central Europe emerged and marine sedimentation was
restricted to several rifts acd basins. This was followed by a gradual rise in sea level,
and by Late Albian time shallow seas again covered much of Europe. The Arctic
and Tethys Seas were again connected. As the rate of seafloor spreading accelerated
in the Late Cretaceous, the Tethys engulfed nearly all of Europe except the Baltic
Shield. The Chalk series was deposited at this time. Near the close of the Creta-
ceous, continued convergence of Eurasia and Africa led to the development of the
Alpine orogeny and a marine regression.
In western Europe the lower portion of the Lower Cretaceous is called the
Wealdian. It is characterized primarily by continental fluviatile and lacustrine
sediments. The facies extends from the London Basin to Belgium, the Paris Basin,
Spain, Switzerland, Germany and north Africa. In these areas the physil suite is
composed primarily of varying amounts of kaolinite (and bauxite) and illite (Millot,
1964; Perrin, 1971; Lopez-Aquayo and Martin-Vivaldi, 1972; Persoz, 1982; Storr,
1975; Mendez and Galan, 1976). In France and Spain kaolinite is mined from these
deposits. The presence of coals and bauxites in western Europe indicates the climate
was warm and humid. Extensive kaolinite and bauxite developed on the emergent
areas (Ardens, Massif Central, Vosges, Arihge and the PyrCnCes-Orientales)(Millot,
1964) and was transported to the flanking low areas.
However, in the southeast of France, shallow water carbonates were deposited.
These contain a variety of detrital physil suites which contain kaolinite, degraded
illite and chlorite, I/S and a relatively large amount of montmorillonite (Chamley
and Masse, 1975; Porthault, 1979). Weathering conditions were less severe than to
the north and west. A similar physil suite occurs in the Lower Cretaceous marls of
the Transdanubian Central Mountains of Hungary. Montmorillonite is abundant in
666

the Mecsek Mountains, where it is associated with volcanogenic sediments (VicziAn,


1975). Volcanism apparently began earlier here, than in western Europe.
The global sea level began to rise in the late Lower Cretaceous (Aptian-Albian).
The eustatic change is believed to be due to the rapid increase in new crust
generated at the spreading centers on the ocean floors. Coincident with the rise in
sea level and the beginning of the most extensive marine transgression identified in
the post-Precambrian, volcanism increased. Aptian and Albian age bentonites are
present in northwestern Europe. They formed from both primary and reworked
ashfalls (Jeans et al., 1982). This period of volcanic activity coincided with the
extensive volcanic activity that occurred in the Rocky Mountain region (Wyoming
bentonites). This suggests a major global crustal disturbance occurred near the end
of Lower Cretaceous time. The location of volcanoes supplying ash to northwest
Europe is believed to have been either the North Sea area (where rifting was
occurring), the North Atlantic or local sources. Volcanoclastic rocks of this age are
abundant in the Scotian Shelf, offshore eastern Canada (Jansa and Pe-Piper, 1985).
Fuller’s earth beds in the Middle Cretaceous range from pure montmorillonite
derived directly from ash falls (Hallam and Sillwood, 1968) to beds with an
appreciable content of illite. The latter deposits were formed from water-transported
ash and Paleozoic illites (Cowperthwaite and Fitch, 1972).
In England (Fig. 9-32), once smectites became a major component, they con-
tinued as a major physil throughout the Cretaceous. The smectitic material ranges
from a good montmorillonite to an I/S with up to 60% illite layers. It is commonly
accompanied by illite, which is frequently the dominant physil. Minor amounts of
kaolinite are present in some samples. Some of the commercial fuller’s earth beds
are pure montmorillonite, probably of volcanic origin (Hallam and Sillwood, 1968).
Glauconite-rich sediments (greensands) are relatively abundant in the Aptian,
Albian and Cenomanian sediments of western Europe. Jeans et al. (1982) demon-
strated that the glauconite in England formed from volcanic debris of mafic
composition (Fe-rich). He suggested that the montmorillonites formed from volcanic
debris of acid or alkaline composition. Jeans et al. equate the glauconite deposits
with those of equivalent age in New Jersey (U.S.), though evidence for a volcanic
origin of the latter deposits is laclung.
During the Upper Cretaceous, chalk was deposited over much of western Europe.
In England montmorillonite (and I/S) and illite, in varying proportions, are the
dominant physils. The relative amount of smectite apparently increases from the
Lower to the Upper Chalk. In the Lower Chalk there are areas in which the physil
suite is composed of kaolinite, mica, smectite, chlorite and vermiculite (Perrin, 1971;
Hallam and Sillwood, 1968; Jeans, 1968; Weir and Catt, 1965). This physil suite,
plus the glauconite-rich rocks, indicate this was a transitional facies between the
older kaolinite-illite suite and the younger smectite-illite suite. In the North Sea the
chalk has been buried deep enough that the montmorillonite has been diagenetically
altered to I/S.
Montmorillonite is the predominant physil in the chalk deposits of the Paris
Basin (Millot, 1970) and is a major component, along with illite of the chalks of
central Germany (Heim, 1957). In northwestern Germany, aggregate grains of
~

STAGE

TIJRONIAN
90 m.y.

XNOMANIAN

95m.v.

7
ALBIAN

APTIAN

IISm..

IARREYIAN

I21m.v.

IWTERIVIAN

126 m.y.

Fig. 9-32. Summary of lithology, clay mineralogy and distribution of volcanogenic material in the
Cretaceous sediments (Wealden Clay-Middle Chalk) of southern England and their relationship to
Cretaceous igneous rocks in the southern North Sea and to primary bentonites in the Cretaceous of north
Germany. From Jens et al., 1982. Copyright 1982 Miner. Soc. London.

montmorillonite have been shown to have formed from volcanic glass fragments
(Valeton, 1960). Cenomarian sediments along the western flank of the Paris Basin
were derived from the Armorican Massif. The continental nearshore facies contains
detrital kaolinite, smectite and mica. In the marine environment (carbonates), these
physils were believed to have been transformed into ferriferous montmorillonite,
glauconite, zeolites and cristobalites (Louail et al., 1979). In the Upper Cretaceous
carbonates of the Aquitaine Basin (southwestern France) illite and I/S, in ap-
proximately equal amounts, are the dominant physils in most samples. Kaolinite is
relatively abundant in the lower Senonian oil-bearing carbonates. The physil suite
does not vary with the dolomite/calcite ratio but does vary with the composition of
the pore water. The physil suite in rocks containing depositional water is illite and
I/S. Where the salinity is high,due to the solution of underlying Triassic evaporites,
the I/S has been converted to illite. The kaolinite was formed during a period of
emergence by the influx of fresh waters (Rumeau and Kulbicki, 1966). In southeast-
ern France (Provence) the most common physil suite contains illite, smectite and
kaolinite. Smectite is the only physil in some samples (Cornet, 1977).
Sepiolite in red beds of northern Italy are believed to have formed diagenetically
from I/S (Veniale, 1966). Sepiolite, apparently authigenic, is also present in the
668

carbonate rocks of the province of Segovia, Spain. In laterally equivalent dolomites,


kaolinite, illite and montmorillonite all occur as the major physil (Mingarro Martin
and Lbpez De Azcona, 1975). The Cenomanian rocks of the western Pyrenees grade
from a diagenetic facies containing illite and I/S, with minor kaolinite and chlorite,
into an anchizone facies composed primarily of illite, with minor chlorite (Dunoyer
De Segonzac et al., 1968).
During the Lower Cretaceous, the sediments deposited in the Negev, southern
Israel, were largely continental. The physils are primarily detrital illite and kaolinite,
with kaolinite frequently being the only physil. Minor smectite is present in the few
marine sediments (Bentor et al., 1963). During the early Lower Cretaceous the area
was transgressed from the north by the Tethys. Illite is the predominant physil,
generally more than 65% of the physil suite. Kaolinite and smectite are equally
abundant; minor palygorskite is present in samples rich in gypsum. Illite increases
to the south, near the crystalline Precambrian source area. In the Middle and Upper
Turonian clays, smectite increases sharply in abundance. This suggests a volcanic
source. In the uppermost Cretaceous smectite is dominant and commonly the only
physil present. Palygorskite is present in brackish bay sediments. The sequence is
similar to that in western Europe, though the first occurrence of abundant smectite,
probably indicative of volcanism, occurs at a later time. The volcanic source was
presumably different from that which supplied the volcanics to western Europe.
In the earliest Cretaceous, large areas of Europe were above sea level, weathering
was relatively intense (warm, humid) and abundant kaolinite was formed. Deposi-
tional environments were largely continental. The physil suite was composed largely
of kaolinite and illite. The presence of the illite indicates weathering was not
extreme, perhaps like the present southeastern United States. Towards the end of
the Lower Cretaceous, volcanism increased (coinciding with the first major phase of
drifting), marine environments became more extensive and the climate became
cooler. All three factors would favor the development of smectite over kaolinite.
Though there has been considerable discussion on the origin of the smectite, the
global increase in volcanism and smectite in the late Lower Cretaceous and the
Upper Cretaceous leaves little doubt that most of it formed from volcanic material,
both in the marine environment and on land, under relatively mild climatic
conditions. The large amount of volcanic activity may have, in part, been a factor in
lowering the temperature.

Africa

Lucas (1962) described two thin sections of Cretaceous red marls from the
Coastal Meseta region of Morocco. The physil suite in one section is composed of
80 to 90% palygorskite and 10 to 20% illite. The suite is similar to that in the
underlying Triassic. In the other section “open” illite (illite-I/S) comprises 95 to
100% of the suite. The presence of red mark and conglomerates in the sections
suggests a continental-lacustrine-likeenvironment. Palygorskite-rich beds are associ-
669

ated with, but not in, the Upper Cretaceous phosphate deposits of Morocco. It is
also present in the coastal region of Gabon (Millot, 1970).
A 4172 m deep well in the Douala Basin, Cameroun (Dunoyer de Segonzac,
1969), penetrated a thick section of Upper Cretaceous sediments which grade from
continental at the base to open marine in the uppermost Cretaceous. The physil
sequence is typically diagenetic. The physil suite in the upper section ( - 12 m)
consists of montmorillonite, I/S (with high smectite), kaolinite and minor illite and
Ch/V. With depth kaolinite disappears. The montmorillonite is systematically
converted to illite (80%) and chlorite (20%). The deepest illite still contains ap-
proximately 5% smectite layers.
The diagenetic conversion is relatively far advanced for the burial depth. The well
is very near the west coast of Africa. As rifting occurred between Africa and South
America the geothermal gradient in the area of the west African coast was
presumably greater than it is at the present time. It is likely that the Cretaceous
sediments were subjected to higher temperatures than occur at present. The base of
the well presently has a temperature of 160°C. Based on the crystallinity index and
the decrease in l O A peak width when heated, it is likely the lower part of the
section has been exposed to temperatures in the range of 200 to 250°C.
The Upper Cretaceous of north-central Tunisia consists of marls and shallow
marine carbonates. The physil suites are complex and show no particular pattern.
Kaolinite and I/S are generally dominant, with illite and chlorite present in minor
amounts. Smectite is dominant in the Lower Cenomanian (Bismuth et al., 1982).
In Nigeria the Cameroun Rift, aligned southwest-northeast, extends from the
present Atlantic coast into the Sahara Desert. It is also referred to as the Benue
Trough and during the Upper Cretaceous extended from the incipient South
Atlantic Ocean to the Tethys Sea. In general, the section grades from deltaic in the
lower part to marine in the upper section as the seas transgressed into the trough.
The physils in the deltaic sections are primarily illite, chlorite and kaolinite. The
suite suggests a source area of active and rapid erosion with high relief and
moderate rainfall. As relief in the source area lowered, the physil suite was
dominated by mixed-layer physils (I/S, I/V, Ch/V and Ch/S) along with kaolinite.
This suite grades upward into one in which smectite is predominant. This suggests a
further decrease in weathering intensity and perhaps some volcanism in the Atlantic.
In the uppermost Cretaceous in coastal coal-bearing deposits, the illite, chlorite,
kaolinite suite reappears, suggesting the source area was rejuvenated (Santonian
Orogeny) (Enu, 1980).

South America

In the Campos Basin, Brazil, the marine Cretaceous shales contain a physil suite
composed of I/S (Fe-rich smectite), kaolinite and biotite. There is no apparent
diagenetic change with depth (3,500 m) (Couto Anjos, 1985). In the Sergipe Basin
and Platform, the physil suite consists of illite, I/S, smectite and kaolinite. Kaolinite
increases eastward, towards the sea (L.E. Neves, personal communication).
670

CENOZOIC

In the Cenozoic era the plates gradually assumed their present form and position
(Fig. 9-33). Both the North and the South Atlantic continued to widen. During the
Paleocene the Greenland part of the American Plate began to separate from
Norway and the Atlantic was connected with the cold waters of the Arctic Ocean.
The African Plate continued to converge on the Eurasian Plate, gradually closing
the Tethys Sea. When the Tethys was consumed during the Oligocene and Miocene,
the colliding plates, formed a huge chain of mountains extending from the Atlas to
the Alps and east to the Himalayas where India collided with the Eurasian Plate.
The closing of the Tethys altered the global circulation patterns and climate. The
closing of the Tethys and the opening of the Arctic to the Atlantic had a major
impact on the distribution of the physil suites during the Cenozoic. Though the
continents were scattered and local factors influenced the make-up of the physil
suites, the global cooling trend is reflected in the general decrease in kaolinite and
palygorskite from the Paleogene to the Neogene. Smectite is relatively abundant
throughout the Cenozoic.

North America

In North America, the oceans and marine sedimentation were restricted to the
fringes of the continent except for a brief epeiric sea invasion, the Tejas Sea, of the
continental interior during the Paleocene.
Extensive changes occurred along the west coast. At the beginning of the
Cenozoic the Pacific Ocean extended inland over the site of the present Cascade
Range and to the foothills of the Sierra Nevada. Continued westward movement of
the continent and subduction led to the development of the Coast Ranges. The
Rocky Mountains continued to grow. The Sevier Orogeny, which lasted from the
Late Cretaceous to the Middle Paleocene, affected a belt from Utah into Canada.
The Laramide Orogeny, whch produced the Rocky Mountains, extended into the
Eocene. Other, smaller areas (i.e., Black Hills, Teton Range) were elevated during
this period. The Great Basin (Basin and Range Province) formed in the Oligocene
when the Pacific Plate began a northwesterly movement.
The basins produced by t h s orogenic activity were filled with detritus from the
adjacent highs. Abundant low-grade coal beds were formed in many of the basins.
The Eocene is distinguished by the development of a system of large lakes that
formed in Colorado, Utah, and Wyoming, of which the Green River is an example.
Many of these lake deposits contain oil shales. The abundance of coal and oil shales
indicates the climate was moderate and humid.
Volcanism in the western United States was extensive throughout the Cenozoic.
In the Eocene, andesitic volcanism occurred in the San Juan Mountains (southwest
Colorado) and the Absaroka Mountains and Yellowstone Plateau (northwestern
Wyoming). The most extensive volcanism occurred in the Miocene when thick
671

basaltic flows covered the Columbia Plateau and the Snake River Plain. In the
Pleistocene and Recent andesitic volcanism occurred in the Cascade Mountains.
During the Miocene the entire Rocky Mountain area was raised over 1000 m and
the climate became drier and cooler. During the Miocene and Pliocene continental
sediments, derived largely from the west, spread as far east as the Mississippi River
and south unto the western Gulf Coastal Plain.
A narrow belt of Cenozoic sediments fringes the Atlantic and Gulf Coast. The
Appalachian Mountains and eastern interior supplied most of the sediments to the
western Atlantic and eastern Gulf of Mexico. Carbonate rocks were deposited in
south Georgia and Florida. Sediments from both the Rocky and Appalachian
Mountains were funnelled into the Gulf Coast rivers, particularly the Mississippi,
and a thick clastic wedge was deposited in the Gulf of Mexico.
The mean global temperature remained relatively warm in the Paleocene and
Eocene and then became progressively cooler. The mean global precipitation
remained low in the Paleocene, was high in the Eocene and Oligocene, and was
moderate for the remainder of Cenozoic time (Frakes, 1979). Though there were
numerous sizable excursions, sea level became progressively lower during Early to
Late Cenozoic time.

Western North America


Deposition in California was primarily in deep fault-bound basins. The rocks are
largely sandstones and shales. The physil suites are extremely variable and it is
difficult to correlate suites between relatively closely-spaced wells. As might be
expected, due to the abundant volcanic activity, montmorillonite and I/S are the
most abundant physils. In the deeper wells the diagenetic conversion of montmoril-
lonite to I/S can be observed. Illite is the next most abundant physil and is
dominant in many samples. Chlorite, Ch/S, and Ch/V are of relatively minor
importance but occur as the major physils in some sandstones. Kaolinite appears to
be relatively scarce. This may be due to the relatively rapid rate of erosion and
deposition (Quaide, 1956; Milne and Earley, 1958; Weaver).
The Sespe Formation, an Upper Eocene to Lower Miocene red bed sequence up
to 4700 m thick, outcrops along the westernmost edge of southern California. The
rocks are primarily coarse-grained and have a quartz/feldspar ratio of one (arkose).
Depositional environments are: alluvial fan, river and floodplain, and
playa-lacustrine. Montmorillonite and I/S are the dominant physils. Montmoril-
lonite is relatively more abundant in the sandstones than in the finer-grained rocks
and is believed to have formed diagenetically from volcanic debris. A slightly
degraded illite (asymmetrical peak) is ubiquitous but in subordinate amounts.
Chlorite and kaolinite occur in trace amounts. Adjacent marine rocks have similar
physil suites. Corrensite occurs locally in evaporite gypsum beds. It apparently
formed from montmorillonite in playa lake environments. Aside from some of the
montmorillonite, the physils are largely detrital. The source rocks were largely
granitic and on the basis of the physils the climate was probably arid to semi-arid
(Flemal, 1967); however, it appears likely that the immature suite of physils was
due, in part, to the high relief and rapid erosion.
673

Extensive Eocene sedimentary kaolinite deposits (Ione Formation) occur along


the western flank of the Sierra Nevada Range. The kaolinite was formed by
weathering of the crystalline rocks in the Sierra Nevada Range. Another sizable
kaolinite deposit of Paleocene age occurs in the Santa Ana Mountain region of
southern California. Both residual and sedimentary deposits, derived from weather-
ing of crystalline rocks, are present in this area (Patterson, 1967).
Montmorillonite and I/S are the predominant physils in Eocene (Tyee Forma-
tion) turbidite beds in the central Oregon Coast Range. Illite is concentrated in the
uppermost portion of each bed (Cummings and Beattie, 1963). The vertical physil
distribution suggests the platy illite settled more slowly from the turbidity flows
than the montmorillonite, which was presumably flocculated.
The Eocene nonmarine coals (Tulameen) of southern British Columbia were
developed in a volcaniclastic sequence. Bentonite partings in the coal originally
consisted of glassy rhyolitic tephra. During initial alteration in the coal swamps the
glass was altered to smectite-cristobalite-clinoptilolite.Where leaching was more
intense, a smectite-kaolinite suite developed. In the southern part of the area, a later
thermal event metamorphosed the smectite to a regularly interstratified I/S with
55% illite layers (Pevear ef al., 1980).
Numerous kaolinite deposits occur in western Oregon and Washington and
northeastern Idaho (Patterson, 1967; Murray and Patterson, 1975). The deposits in
the western area are primarily of Eocene age and of sedimentary origin. Some are
associated with coal. The deposits in western Washington and Idaho (Latah
Formation) are of Miocene age. They occur both as residual clay weathered from
granitic material and as transported lacustrine sediments (Ponder and Keller, 1960).
Note that the kaolin deposits along the west coast, of Paleocene and Eocene age,
were formed during a time of high global rainfall and were formed on the windward
side of the coastal mountain ranges. However, it is not clear why kaolinite is so
scarce in the thick basin deposits. Presumably it is because of rapid rates of erosion
for most of the area. Kaolinite apparently formed locally in areas where the relief,
or at least erosion, was moderate and a deep weathering profile had time to develop.
During the Cenozoic a large number of saline lakes existed in western North
America. Though the Green River Lake system is by far the largest, many existed in
which a wide variety of exotic minerals were precipitated. Quaternary lake deposits
are discussed in Chapter IV.
In the Miocene-Pliocene lacustrine borate beds at Boron, California, montmoril-
lonite is the major physil and is most abundant in borate-bearing rocks. Illite is
present in all rock but is most abundant in the borate-barren siltstones and
claystones. Trace amounts of kaolinite and Ch/V are present. Both the illite and
montmorillonite are believed to be detrital. The supply of detrital illite remained
essentially constant; the montmorillonite influx increased periodically during pulsat-
ing episodes of volcanism. The volcanic waters were the source of the boron (Gates,
1959). Montmorillonite, along with minor illite, is also the major physil in the
borate-bearing strata in the Death Valley area (Droste, 1963).
Papke (1969) described 11 lacustrine Miocene and Pliocene montmorillonite
deposits from southern and western Nevada (Basin and Range Province). It is
614

believed the montmorillonites, up to 7 m thick, formed from volcanic ash shortly


after it fell into alkaline lakes.
In the Ash Meadows and Amargosa Flat areas of the Amargosa Desert, southern
Nevada, large Pliocene-Pleistocene pluvial lake deposits contain sepiolite beds as
much as 1.2 m thick. Large deposits of mixed-layer kerolite/stevensite are associ-
ated with the sepiolite. Isotopic data and the lack of abundant soluble salts suggest
the physils precipitated in low salinity waters. Khoury et al. (1982) suggest the lake
was fed by spring water from Paleozoic carbonate aquifers. Calculations indicate
only minor evaporative concentration is required for these two physils to precipi-
tate.
Because of the wide variety of local source areas, the western interior basins have
a wide variety of physil suites. Aside from the Green River Basin the physils have
not been investigated in detail. Illite, I/S, smectite, and kaolinite all occur as
dominant physils.
Kiersch and Keller (1955) described Pliocene bentonite deposits (Cheto) from
northeastern Arizona. Latitic ash falls filled streams and lakes and under fresh
water conditions altered to Ca, Mg montmorillonites. Authgenic or diagenetic
smectite and illite (broad peak) are present in altered tuff beds in the lacustrine Big
Sandy Formation (Pliocene) of western Arizona; zeolites and K-feldspar are also
abundant (Sheppard and Gude, 1973). Well crystallized kaolinite occurs as tonstein
partings in coal beds throughout the Rocky Mountain region. These form as the
result of volcanic ash falls in the acidic coal-swamp environment (Bohor, 1983;
Pollastro et a[., 1983).
During the Eocene a large system of interior lakes, the Green River System,
covered large areas of western Wyoming, northeastern Utah, and northwestern
Colorado. The Green River Formation consists dominantly of carbonate rocks.
During the initial stage the lake was a large fresh-water lake. It was during this
period that the oil shales (Tipton Shale Member) were deposited. During an
intermediate stage and conditions produced a restricted highly saline environment.
In the final stages climatic changes caused a return to fresh-water conditions and
lake expansion (Bradley, 1964). Over 50 authigenic mineral species have been
described, including talc, sepiolite, stevensite, saponite, chlorite, and a wide variety
of zeolites and carbonate minerals; however, most of the minerals are detrital. Well
over 100 thin, less than a few centimeters thick, tuff beds are present in the Green
River Formation. Most of these beds have altered to analcine; authigenic albite, and
less commonly K-feldspar, are major constituents in some tuffs; a few have altered
to montmorillonite (Iijima and Hay, 1968).
In the Uinta Basin, the Green k v e r Formation is approximately 1000 m thick.
The formation can be subdivided into nine physil zones which can be correlated
over a distance of at least 11 miles (three wells) (Weaver). Illite and I/S, along with
appreciable kaolinite and minor chlorite, comprise the physil suite in the lower part
of the formation. This is a detrital suite deposited during the early fresh water
phase. Higher in the section kaolinite disapppears and chlorite increases in abun-
dance; analcite is present. This interval presumably represents the high salinity stage
of the lake. Near the top of this interval I/S is absent and chlorite is at a maximum.
675

The chlorite could be authigenic; unfortunately, it was not studied in detail. Well
crystallized montmorillonite is the dominant physil in the upper part of the
formation and the overlying Uinta Formation. This is presumably the final fresh
water phase that developed as the lake was filled with sediments. The high feldspar
content suggests the montmorillonite and feldspar formed in the lake from volcanic
detritus.
Tank (1972) analyzed a well from the center of the Green River Basin, Wyoming.
Illite is relatively abundant (generally > 40%) throughout the section and is the only
physil in some of the oil shales. Montmorillonite is relatively abundant throughout
the section but is inversely related to loughlinite ( - Na sepiolite), which is only
present in the middle, high salinity interval. Chlorite is present only in the silty and
sandy beds in the lower part of the section. Tank suggests all of these physils are
au thigenic.
Petrographic evidence of loughlinite replacing a number of other minerals
indicates it formed authigenically or diagenetically (Fahey et al., 1960). The inverse
relation between montmorillonite and analcine in tuff beds suggests the former
crystallized when the lake waters were relatively fresh and the latter when salinity
(Na) was relatively high (Surdam and Parker, 1972). The restriction of the chlorite
to the siltstone-sandstone intervals suggests it could be diagenetic. The fact that
some of the oil shales contain illite as the only physil is interpreted by Tank to
indicate an authigenic origin for illite. However, most marine black shales contain
an illite-rich or illite-only physil suite. It is possible that other physils are prefer-
entially destroyed in the organic-rich environment. K-Ar analyses should solve the
problem.
The Eocene Golden Valley Formation of North Dakota provides an example of
physils deposited in a more aggressive lacustrine environment than the Green River
Formation. The physil suite of the clay beds from the middle part of the formation
-
is composed predominantly of kaolinite ( 70%) with minor amounts of montmoril-
lonite and illite. Montmorillonite, with varying amounts of illite and kaolinite and
minor chlorite, is the major physil in the underlying and overlying sandstones. The
physils are detrital and were transported into a freshwater lake from a western
source (Freas, 1962).
Montmorillonite is the dominant physil in the Middle Cenozoic fluvial deposits
of southwestern North Dakota. Zeolites are present in many samples (Stone, 1973).
The major physils in the continental Fort Union Formation (Paleocene) of south-
western North Dakota are illite and montmorillonite; varying amounts of kaolinite
and I/S are present. A few bentonite beds are present. Sepiolite (?) is locally
abundant in backswamp and floodplain deposits (Maisano, 1975). Pedogenic
montmorillonite is the dominant physil in the Middle and Upper Cenezoic eolian
and fluvial sandstones of Wyoming and Nebraska. The physils formed under
semiarid conditions from unstable grains in the volcaniclastic sandstone (Stanley
and Benson, 1979). Montmorillonite and palygorskite are the dominant physils in
the continental Pliocene limestones and sands of New Mexico. Palygorskite is
apparently a weatheirng product of montmorillonite (Frye et al., 1974).
676

Gulf Coast
During the Cenozoic a thick wedge of clastic sediments was deposited in the Gulf
of Mexico. The depositional pattern was similar to that of the present. Several cycles
of sedimentation occurred as the sea periodically advanced and retreated. During
the transgressive phase, thin sequences of shallow marine sands, clays, and marls
were deposited. These units alternate with thick deltaic sequences deposited during
the regressive phase of sedimentation. The shoreline moved progressively seaward.
The modern Mississippi River, which dominates the Gulf region, developed at least
by Miocene time. Over 5000 m of Miocene to Holocene sediments have been
deposited in the Mississippi delta area. The detritus was derived largely from the
volcanic-rich midcontinent and Rocky Mountain region. Montmorillonite, devel-
oped in the soils and lakes, was the major physil delivered to the Gulf. A lesser
volume of physils, largely illite, was derived from the Appalachian Mountains. The
Appalachians were a more important source of detritus in the early Cenozoic than
in more recent time.
During the Eocene a major depocenter existed in southwest Texas. Over 7000 m
of sediment were deposited, primarily in deltaic and inner neritic environments. The
physils in the Lower Eocene Wilcox Formation consist of 50 to 70% I/S ( -= l p m
fraction) and varying amounts of illite, chlorite, and kaolinite. The deeply buried
sediments show a typical sequence of diagenetic changes. Kaolinite decreases and
chlorite and illite increase with depth. The proportion of illite layers increases from
40% (60°C) to 60 to 80% (210°C) with depth (Boles and Franks, 1979). The physil
suite and changes with depth are typical of the thicker sections of sediments
throughout the Gulf Coast region. The process of burial diagenesis is discussed in
more detail in Chapter VII.
A belt of Eocene age (mostly Early Eocene) rocks containing scattered kaolinite-
bauxite deposits occurs along the updip edge of the Eocene deposits flanking the
Gulf of Mexico and southwestern North Atlantic. The belt starts in east Texas and
follows the edge of the Mississippi embayment as far north as southern Illinois; it
swings south through western Kentucky and Tennessee, then east through Missis-
sippi and Alabama and northeast through central Georgia and into western South
Carolina. The kaolinites are primarily sedimentary (detrital) and were deposited
near the edge of the Eocene shoreline (Patterson, 1967; Murray and Patterson,
1975). These deposits are frequently associated with thin lignite beds. Aside from
the kaolin beds, kaolinite is apparently relatively abundant in the Lower Eocene
rocks in general. It has been reported as the major physil in several formations in
Texas (Simons and Taggart, 1954; Roberson, 1957).
In the Lower Eocene Wilcox Formation of Texas, kaolinite is the dominant
physil in the updip fluviatile sediments. Kaolinite decreases and montmorillonite
increases downdip; the latter is dominant in the shallow marine sediments. The
kaolinite formed by weathering in coastal swamps; the montmorillonite was ap-
parently transported to the offshore area by longshore currents (Griffin, 1962).
In Texas, Upper Eocene and Oligocene bentonite deposits outcrop along the full
length of the Coastal Plain (Roberson, 1964). In the lower part of the section
(Jackson Group) the bentonites are composed of well-crystallized montmorillonite
677

which is believed to have altered from volcanic ash in coastal lacustrine environ-
ments. Clinoptilolite and opal are present in the base of the section (Eargle, 1968).
The upper part of the section (Catahoula Formation) contains an abundance of ash,
tuff, and tuffaceous sandstones. The bentonites contain a poorly crystallized
montmorillonite (apparently an I/S with 50-60% illite layers). The high content of
clastic material in these latter deposits indicates the bentonites are formed from
reworked bentonites or water-transported volcanic material. The source was to the
southwest (Roberson, 1964).
It is of interest that in the Texas deposits, and many of the Cenozoic deposits,
fresh, unaltered volcanic glass beds occur intimately mixed with completely altered
bentonite beds. It is not clear if the difference in the rate of alteration is due to
slight differences in composition or to the availability or composition of water.
Senkayi er al. (1984) described a bentonite in a Late Eocene lignite bed in
east-central Texas that was partially altered to kaolinite (tonstein) as humic solu-
tions moved downward through the bentonite layer. In Arkansas and Mississippi
bentonite beds are present in the Oligocene age Wilcox Formation (Grim and
Giiven, 1978). These are approximately the same age as the younger Texas be-
ntonites.
The physils in the marine Lower Eocene Wilcox sediments of central Louisiana,
700 to 5000 m burial depth, show the typical changes incurred during burial
diagenesis. The shallow physil suite is composed primarily of montmorillonite;
varying amounts of chlorite, kaolinite, and illite are present. Montmorillonite was
systematically converted to I/S and the thermal stability of chlorite increased with
depth (Burst, 1959). However, during the early Cenozoic the drainage basin of the
Mississippi River lay largely in the eastern United States (Paleozoic and metamor-
phic rocks) and appreciably more illite, chlorite, and I/S must have been supplied
to the eastern Gulf than during the Neogene when most of the detritus came from
the montmorillonite-rich western source. Thus, the physil changes in the Eocene
section may not be primarily due to diagenetic changes, as they are in the younger
sediments.
In addition to the two studies of thick Lower Eocene sections that have been
discussed, a number of studies have been made of younger, thick sedimentary
sections in coastal and offshore eastern Texas and Louisiana: (Weaver, 1957:
Miocene-Oligocene, Texas; Perry and Hower, 1970: Oligocene, Texas, Oligocene-
Eocene, Texas, Pleistocene, Louisiana, Miocene, Louisiana; Weaver and Beck, 1971:
Pliocene-Miocene, Louisiana; Hower et al., 1976: Miocene-Oligocene; Morton,
1985: Miocene-Oligocene, Texas; Freed, 1980: Miocene-Oligocene, Texas; Weaver:
Pliocene-Oligocene, Texas; and others). Wells have been studied to a depth of
approximately 6000 m. The sediments were primarily deposited in marine-deltaic
and deep marine environments. In general the physil suites and diagenetic changes
with depth are similar. The fine fraction of the shallow samples contains from 70 to
90% dioctahedral smectite and varying amounts of illite, kaolinite, and chlorite.
With depth the smectite converts to I/S, kaolinite decreases, and chlorite increases.
The depths at which the various reactions occur are controlled largely by the
geothermal gradient. The vast amount of detrital smectite was derived from the
678

Mesozoic and Cenozoic altered volcanic material in the mid-continent and Rocky
Mountain region. However, it seems likely that a fair amount of volcanic debris was
transported by water and air to the Gulf region and altered to montmorillonite at
the site of deposition, probably after burial.
East from the Mississippi delta area Cenozoic sediments thin drastically. Down-
dip from the Paleocene-Middle Eocene continental kaolinite-bauxite deposits of
Mississippi and Alabama, the approximate equivalent age marine rocks contain
montmorillonite, clinoptilolite, opal-CT, authigenic feldspar, and quartz (Reynolds,
1966; Roquemore et al., 1984). The authors could find no evidence of a volcanic ash
or biogenetic precursor and suggested that the clinoptilolite precipitated from pore
and vein fluids. It is not known whether the montmorillonite is detrital. In view of
the widespread volcanic activity in the Lower Cenozoic and the common occurrence
of zeolites and opal-CT with many of the altered volcanics, it seems likely that
appreciable volcanic material was present in the primary Mississippi and Alabama
deposits.

Eastern North America


During the Paleogene, a relatively thin layer of predominantly carbonate rocks
was deposited in the southeastern United States. A northern clastic facies gradually
moved south and by Miocene time extended into central Florida. A few analyses of
the carbonate rocks indicate montmorillonite, along with opal-CT, is the predomi-
nant physil (Weaver, 1968). The updip portion of the shallow marine, upper Eocene
Twiggs clay of Georgia is composed largely of montmorillonite and opal-CT,
derived from diatoms. Seaward the opal decreases and calcite increases (Carver,
1972). Clay deposits in South Carolina consist largely of montmorillonite and
opal-CT, with some zeolite (Heron et al., 1965). Montmorillonite, along with
clinoptilolite and K-feldspar, is also the dominant physil in the carbonate rocks on
the continental shelf and Blake Plateau (Weaver, 1968). In the coastal area of
central Georgia large amounts of commercial grade, detrital kaolinite were de-
posited (Patterson and Buie, 1974).
Paleogene clays composed primarily of montmorillonite and opal-CT, with and
without zeolites, extend from Texas to South Carolina. Though there is little
evidence that the minerals were derived from volcanic material, the abundant
volcanic activity in the Caribbean, Atlantic, and Mississippi embayment areas
suggests volcanic material was supplied to the southern and eastern Coastal Plain.
Bentonite beds are present in the Paleocene in Delaware (Jordan and Adams, 1962)
and offshore Florida.
In the area from south Florida to North Carolina, montmorillonite is the major
physil in the Neogene coastal to shallow marine clastic sediments. The Miocene
sediments contain large commercial deposits of phosphate (Florida, Georgia, North
Carolina) and palygorskite-sepiolite (Florida and Georgia). Most of the palygorskite
formed from montmorillonite in shallow, brackish water lagoons and tidal flats.
Formation of palygorskite was restricted to Lower Miocene time but reworked
palygorskite occurs in younger Miocene sediments (Weaver and Beck, 1977).
679

Clinoptilolite is commonly present in the montmorillonitic marine sediments. Con-


tinental shelf and Blake Plateau sediments have a similar physil suite.
To the north, in New Jersey, montmorillonite, I/S, and glauconite occur as major
physils in Eocene marls. The Miocene physils are largely kaolinite and illite.
As in the central and western Gulf, smectite is the major physil in the eastern
Gulf and in the eastern United States. Most of the smectite in the central and
western Gulf was transported by rivers from the mid-continent and Rocky Moun-
tain region. Some of this material was probably transported by clockwise flowing
Gulf currents into the southeastern United States and north along the Atlantic
Coast. In part, this would account for the preferential concention of smectite in the
marine sediments. However, some volcanic ash was undoubtedly transported to the
area by wind and marine currents. Bentonite beds are present in the carbonate rock
on the continental shelf of the southeastern United States (Hathaway et al., 1970).
Volcanic activity was widespread in the circum-Caribbean area (Donnelly, 1973), in
the central eastern North Atlantic and the Labrador Sea (Jansa and Pe-Piper, 1985),
and in eastern Mexico (Axelrod, 1981).
In contrast to what might be expected, kaolinite is the major physil in some
Cenozoic sediment in the Canadian Arctic. On the Ellesmere Islands, Eastern
Canadian Arctic Archipelago (79'-81 O N latitude), the sediments grade upward from
deltaic to alluvial. The average physil composition of the Paleogene sediments is
75% kaolinite and 20% illite; Neogene sediments contain 50% kaolinite and 35%
illite. Varying amounts of chlorite, vermiculite, I/S, and smectite are present. The
physils are thought to be primarily detrital (MarcBustin and Bayliss, 1979). How-
ever, the distribution of kaolinite correlates well with the estimated paleotempera-
ture (Frakes, 1979), generally decreasing (65% to 15%)upward through the Neogene.
there is an abrupt decrease in kaolinite in the late Pliocene, coinciding with the
beginning of Arctic glaciation. The correlation of kaolinite with temperature sug-
gests, that, at least, in the Neogene some kaolinite formed in the surrounding soils.

Europe

Though the Tethys sea was still in existence at the beginning of the Cenozoic, it
was rapidly displaced or destroyed as numerous mountain ranges began to appear in
Europe and Asia as southern and northern land masses collided. The western Tethys
was closed by a scissor-like motion that began at the west end and moved eastward.
The Pyrenees and Altas Mountains emerged at the close of the Eocene. Movement
in the European Alps and Carpathians began in the middle Eocene but the most
important folding occurred during the Oligocene. Most of the uplift in the Himalayan
area occurred during the Middle Miocene.
In the Early Cenozoic most of Europe between the Tethys (Alpine area) and the
Baltic shield was occupied by a series of shallow seas and lagoons. The eroded
remnants of the Hercynian Mountains formed a discontinuous barrier between the
southern warm Tethys and the cold northern seas which were connected to the
680

North Atlantic Ocean. Following a period of mountain building, a number of


residual seas, during the Miocene and Pliocene, existed adjacent to the mountains
and received detritus from them. The largest volume of Cenozoic sediments was
deposited in the North Sea and Pannonian (Hungary) Basins. It should be noted
that the Mediterranean Sea is not a remnant of the Tethys but a new sea formed in
an opening related to the formation of the Atlantic Ocean (Ager, 1980; Ziegler,
1982).
In Great Britain, Cenozoic sediments are largely restricted to the London and
Hampshire Basins (reviewed by Perrin, 1971; Sellwood and Sladen, 1981; Shaw,
1981). Most of the sediment was derived from the west and the depositional
environments changed from alluvial in the west to brackish and shallow marine
eastward. Kaolinite is abundant in the western alluvial sediments. Illite and I/S
(high smectite) are abundant in the eastern sediments; minor chlorite is present.
Discrete smectite beds are common in the marine sediments in the London Basin.
These beds are frequently associated with zeolites and were presumably derived
from volcanic ash. Gilkes (1978) found that in the Hampshire Basin illite was
abundant, in excess of 80%, in the upper Eocene and Oligocene sediments. As there
was no apparent source for detrital illite, he suggested it was neoformed in
calcareous lakes and lagoons.
The Cenozoic section in the North Sea is considerably thicker than that in Great
Britain. I/S with - 90% smectite layers is the dominant physil in the lower
Cenozoic of the Viking Graben. Diagenetic changes, increases in illite layers,
commonly occur at a depth of 2000 to 2500 m in Paleocene sediments (Pearson et
ul., 1982). In a 3000 m section, through a submarine fan in the southeastern part of
the central Cenozoic basin, montmorillonite is the dominant ( - 70%) physil in the
Paleocene, Eocene, Oligocene, and Lower Miocene (Karlsson et al., 1979). In the
younger sediments smectite systematically decreases (to 20%) and illite and chlorite
(starting in uppermost Miocene) increase. Kaolinite remains relatively constant at
- 10%.
Ash beds and volcanic material are abundant in the Paleocene and Eocene and
are presumably the source material for much of the smectite. The volcanic activity is
believed to be related to the opening of the Atlantic. Volcanic derived smectite has
been reported from the Lower Cenozoic of Germany and Denmark and the Goban
Spur area of the eastern Atlantic (DSDP Sites 549 and 550) (Knox, 1985). As
volcanic activity decreased and the climate became cooler, with a decrease in
weathering intensity, the proportion of detrital illite, and later chlorite, increased.
Farther north, on the Norwegian continental shelf, smectite is the dominant clay
throughout the claystone-rich Cenozoic section (Roaldset, 1985).
Montmorillonite is the major physil in the Paleogene sediments, largely marine
clays, of Denmark (Tank, 1963; Nielsen, 1974). Montmorillonite, with varying
amounts of illite, is predominant in the Paleocene and Lower Eocene. In the
younger sediments montmorillonite decreases and kaolinite (20 to 60%) increases.
Tuff beds occur in the Lower Eocene. Much of the montmorillonite in the lower
part of the section presumably formed from volcanic material. Detrital physils
become increasingly more abundant in the younger sediments.
681

The deep ocean which was centered in the southern part of the North Sea
periodically expanded and contracted, and transgressed southeast into the Paris
Basin. Fig. 9-34 summarizes the distribution of physils in the Paris Basin. Environ-
ments range from marine in the western part of the basin to continental at the
eastern edge of the basin. The physils reflect the evolution of the basin as described
by Ager (1980).
The initial Cenozoic transgression occurred in the early Paleocene (Dano-
Montien); smectite, along with illite, was deposited in the western marine environ-
ment. By YprCsien time a full marine-lagoon-lacustrine-continental sequence of
facies was developed. The source of the sediments was primarily the Massif Central,
to the south. The physils grade from illite and smectite in the marine sediments to
kaolinite and kaolinite/smectite in the more shoreward environments (Fig. 9-34).
Emergence occurred during Upper Lutetian and Bartonian time; a karstic surface
developed and lagoonal deposits were abundant. Attapulgite (palygorskite) and
sepiolite were neoformed in the lagoonal and lacustrine environments. Gypsum with
interbeds of attapulgite are present in the Ludian (Upper Eocene) deposits. During
the Oligocene (Stampien) the deposits were primarily “semi-marine” and freshwater
limestones; smectite, attapulgite, and sepiolite occur in the latter deposits. The
general uplift of northwestern Europe at the end of Oligocene time caused the
complete removal of the sea from the Paris Basin. Freshwater limestones, with
smectite, attapulgite, and sepiolite, were deposited during the Lower Miocene
(Aquitanian). Younger deposits are continental kaolinitic clays and sands contain-
ing varying amounts of illite, smectite, and I/S.
Detailed discussions of the distribution and origin of the physils in the Paleogene
sediments of the Paris Basin have been provided by Trauth et al. (1969) and Trauth
(1977). The former authors divide the sediments, based primarily on the physils, into
four zones. The lower three zones form an evolutionary sequence in which the
physils reflect the processes of alteration, transformation, and neoformation. In the
fourth, upper zone, the physils are detrital.
The lower zone (Upper Paleocene) contains the Al-rich physils (A1 Zone)
kaolinite, kaolinite/smectite, and beidellite that were formed by the weathering of
the physils in the underlying and adjacent chalk deposits. The second zone (Lower
Eocene), or Fe Zone, contains ferrous beidellite and glauconite, occurring in
transgressive marine sandstones. The beidellite was believed to have been derived
from vertisols on the basin slopes or from weathered chalk as the sea transgressed
over it. K was extracted from the sea water and the Fe beidellite converted to
glauconite. The third zone (mostly Upper Eocene), the Mg Zone, contains neoformed
attapulgite and sepiolite, along with some illite and smectite. The sediments are
chalk, dolomite, and gypsum. Varying amounts of attapulgite and sepiolite are
present, along with smectite, in all three rock types. The chain structure physils are
believed to have formed from material supplied by the destruction of smectite and
Mg from the water (Trauth, 1977). As in other areas, smectite is the source of A1 for
the formation of attapulgite. Oxygen isotope studies indicate the calcite and gypsum
formed from continental waters (Fontes et al., 1967). The physils in the overlying
sediments are largely detrital.
682

Palygorskite (or attapulgite) and sepiolite are abundant in the Cenozoic of


France and Spain. Millot (1970) has reported the occurrence of neoformed attapul-
gite and sepiolite, commonly associated with limestones and mark, in lacustrine
deposits in Mormoiron (Eocene), Salinelles and Sommikres (Middle Oligocene)
(these two deposits were more recently described by Trauth (1977)), Cormeilles-en-
Parisis (Eocene), Herbeville (Eocene), Lagny, SCzanne, Aquitaine (Oligocene and
Eocene), and LeLocle (Miocene of Switzerland). Chanley and Colomb (1967) found
neoformed palygorskite with montmorillonite in the Upper Miocene lacustrine beds
of the Cucuron Basin (Vaucluse). Millot suggested deposits in the basins of Lower
Brittany were formed in a marine environment, probably peri-marine.
During much of the Cenozoic the Hercynian massifs were extensively weathered,
producing thick kaolin deposits. Millot believes the material necessary for the
neoformation of chain physils and associated carbonate minerals in the alkaline
lakes was produced by the chemical weathering that occurred on the massifs. The
precipitation of sepiolite is easy to understand; however, the mobilization of enough
A1 to allow precipitation of palygorskite is a problem. Presumably montmorillonite
or other physils were the source of the Al.
EstCoule-Choux (1984) described the occurrence of palygorskite in a shallow
Cenozoic basin or gulf in the west-central portion of the Armorican Massif. The
palygorskite is associated either with smectite or Fe-rich illite (10 to 15% Fe,O,). All
three physils are considered to have been neoformed from ions supplied from the
surrounding lateritic crusts formed in a tropical climate. They are present in both
marine and continental littoral environments and formed when the climate was dry
and evaporation occurred.
Stevensite, along with sepiolite, occurs in the upper Paleocene evaporate sedi-
ments (gypsum and dolomite) of the Mormoiron Basin (Trauth, 1977). Trauth
believed the precursor was a dioctahedral smectite.
The Aquitaine Basin is bounded to the north and east by the massifs of Armorica
and the Massif Central and to the south by the Pyrenees. At the end of the
Cretaceous the sea receded from the northern Aquitaine, and during the Lower
Eocene rivers draining the Massif Central carried kaolinite from the thick lateritic
soils to the basin. Coarse alluvial sands and gravels along with lenses of kaolinite
were deposited. These deposits are overlain by micaceous sands containing illite and
montmorillonite. This suggests the lateritic crust on the massif had been breached.
The alluvial sediments were subjected to lateritic weathering and some of the mica
was altered to kaolinite, and in some instances, gibbsite. Flora and fauna indicate
the climate was tropical (Millot, 1970).
More than 30 kaolinite quarries are present in the Charentes region. The kaolins
are associated with lignite beds but are detrital. The Upper Paleocene deposits
contain well-crystallized kaolinite derived crystalline rocks of the Massif Central
region. Lower Eocene disordered kaolin was derived from the weathering of
sedimentary rocks on the border of the basin. Some recrystallization and gibbsite
formation occurred in the deposits associated with the acidic lignite deposits
(Dubreuilh et al., 1984).
Farther to the southwest, in the marine portion of the basin, illite and montmoril-
F A C I ~MARIN F A C I ~LAGUNAIRE FACIIESCONTINENTAUX D'ALTE~ATIONET,
FACIIS
LACUSTRES FACI~SCONTINENTAUX DETRITIQUES

OUATERNAIRE
I KAOLINITE (SMECTITE)
PLIOC?NE ILLITE
1 INTERGRADES
& INTERSTR.(10-14M)
MIOC~NE 1 I
BURDIGALIEN SMECTITE. KAOLINITE.ILLITE
AQUITANIEN I SMECTITE,ATTAP.,SEPIO.
SMECTITE
ILLITE 1 SMECTITE,ATTAP.,SEPIO. I
KAOLINITE
STAMPIEN
I
ATTAPULGITE, SEPIOLITE, ILLITE
1
LUDIEN
ILLITE,SMECTITE ATTAPULGITE
I I
AUVERSIEN (KAOLINITE) 1 SMECTITE
SUP~RIEUR ATTAP. ,SEPIOLITE ' I
LUTEITIEN
ILLITE,SMECTITE
- I
1
I
INF .+MOYEP IUOLINITE
I
(ATTAPULGITE) I KAOLINITE

I
CUISIEN KAOLINITE,SMECTITE SMECTITE ALUMINEUSE
YPR& IEN INTERSTRATIFIES KAOLINITE-SMECTITE SMECTITE ALUMINO-MAGNESIENNE
SPARNACIE) ILLITE I KAOLINITE I INTERSTR. KAOLINITE-SMECTITE
SMECTITE ( ILLITE) KAOLINITE
THANETIEN GLAUCONITE I
I I
DANO-MONTIEN SMECTITE(ILLITE
I I
Fig. 9-34. Physil distribution in Pans Basin Cenozoic sediments, from west (left) to east (right). From Sittler el a/.. 1978.
m
00
P

F O S S k R H k N A N F O S S t ! R H O O A N I E N

WESSE MAVENCE BRutwSAL


LANMU
ALYCE YdNE SAWIE AIX - MARSEILLE L A MOUEDOC

Fig. 9-35. Distribution of physils in the Rhine and R h h e Grabens. From Sittler er af.,1978. Copyright 1965 Mkm. Sew. Carte Gkol. Alsace Lorraine.
685

lonite comprise approximately 254%of the Lower Eocene physil suite. The amount
of illite and montmorillonite is more abundant in the uppermost Lower Eocene
sediments and becomes increasingly more abundant through the Middle and Upper
Eocene section, primarily carbonates. Chlorite also increases upward in the section
as kaolinite decreases to approximately 20 to 254%(Klingebiel and Latouche, 1964).
The physils in the Oligocene calcareous lacustrine sediments are largely detrital
illite derived from the Pyrenees to the south (Kulbicki, 1953). Many of the beds
contain Fe-rich illite as the only physil, and Millot (1970) suggests that because of
its purity it is probably neoformed.
Kaolinite is the dominant physil in the Lower Eocene lignitic shales and
sandstones of the Bouxwiller syncline (Bas-Rhin). Illite becomes predominant in the
overlying limestones. Both are apparently detrital and reflect a change in climate
(Trauth ef al., 1977).
In the Province area (southeast France) the basins were filled with continental
sediments, largely clays and carbonates, from the Upper Cretaceous to Miocene
time. Illite and montmorillonite are predominant; minor chlorite, kaolinite, and I/S
are present. Neoformed palygorskite is locally abundant in the Eocene carbonate
sediments. Some of the montmorillonite is detrital and some is believed to have
precipitated from solution, particularly that in the carbonate rocks. The lack of
kaolinite in the Eocene is due to continued tectonic uplift of the source area, which
limited the time available for intense weathering (Sittler, 1965; Cornet, 1977).
The Rhine Graben (FossC RhCnan) and RhCne-SaCne Trough (FossC Rhodanien)
extend from central Germany through eastern France to the Mediterranean. They
form a narrow belt bounded by normal faults and filled with Cenozoic sediments.
The depression is filled with a variety of continental, marine, and evaporite
sediments. During the Miocene parts of the Rhine Graben were covered with
basaltic lavas. The Eocene sediments are primarily continental, the lower and
middle Oligocene marine and upper Oligocene and Miocene continental.
Fig. 9-35 shows the distribution of physils in a section extending from the Hesse
Depression, north of the Rhine Graben, to Languedoc near the Mediterranean, a
distance of approximately 1000 km (Sittler, 1965). The regional distribution of the
physil suites is quite variable. This is due to the variety of both source areas and
depositional environments. In general, the Eocene is characterized by an abundance
of detrital kaolinite, as is much of the Eocene of western Europe. Locally, palygors-
kite and sepiolite are present in lacustrine deposits. Illite, montmorillonite, and I/S
are more abundant in the southern portion of the trough, indicating weathering was
less intense in the southern source areas. Some of the I/S formed diagenetically
from smectite. Montmorillonite and illite are predominant in the Oligocene sedi-
ments; varying amounts of kaolinite, chlorite, and I/S are present. These physils are
primarily detrital, though in the salt beds chlorite is believed to have formed from
degraded physils. Much of the illite at the edge of the troughs is degraded and is
partially to completely reconstituted in the center of the basin (potash beds).
The flora indicate the climate during Eocene time was warm and humid, became
cooler and drier during Oligocene time, and again became warm and humid during
Miocene time, but less so than during Eocene time.
686

Farther north, in the northern and central portion of the Rhine Graben, the
Cenozoic (Upper Eocene to Miocene) physil suites are composed primarily of
smectite, I/S, and illite, with varying amounts of chlorite and kaolinite. The
geothermal gradient is relatively high and most of the smectite has been converted
to I/S, with the proportion of illite layers increasing with depth (Heling, 1974,
1978).
A well (Pierrefeu) drilled in the southern portion of the Languedoc Depression at
the southern end of the Rhodanian Trench (near the Mediterranean coast) penetrated
3000 m of Oligocene clays and limestones. The physil suites reflect the effect of
burial diagenesis. Montmorillonite, illite, and chlorite are present in the shallow
sediments. The upper rocks are evaporitic and contain a relatively high content of
chlorite which was probably neoformed. With depth the montmorillonite is con-
verted to I/S and at total depth ( - 180°C) is almost, but not entirely, converted to
illite. Minor kaolinite is present in the sediments below the evaporite zone and
persists to the bottom of the well (Dunoyer de Segonzac, 1969). The sequence is
similar to that in the southernmost well in Fig. 9-35, which is located slightly to the
west of the Pierrefeu well.
As is evident from the preceding discussion, extensive deposits of kaolinite were
developed during Eocene time when a warm humid climate prevailed and lateritic
soils were developed. These soils on weathering crusts were extensively developed,
on a wide variety of rocks, on the Massif Amoricain and Massif Central. These
crusts are best preserved in the Massif Amoricain. Numerous transported or
sedimentary deposits were derived from the Massif Central, wluch was periodically
“rocked” by Alpine tectonic activity. Relatively coarse, kaolinite-rich sediments
were deposited in the fringing piedmont deposits to the west (Charentes and
Dordogne) and north (Grande Brenne). Some of these deposits were further altered
after deposition. Finer grained kaolin was transported farther into and deposited in
lagoons along the southern border of the Paris Basin. Erosion on the tectonically
quiescent Massif Armoricain was less intense, but fine-grained kaolinite was de-
posited in “little inland hollows” (Rennes, Taulven, Central Brittany) (Esteoule and
Esteoule, 1976).
The Cenozoic of the Iberian Massif (Spain) consists largely of thin, flat-lying
continental sediments. In places evaporites are present. Palygorskite and sepiolite
are abundant in some of the lacustrine deposits.
The largest palygorskite deposit in Spain is located in the province of Caceres in
the western part of the country. The palygorskite, of Miocene age, occurs as lenses
(3-8 m thick) in lacustrine marly beds. The total thickness of the Cenozoic
sediments is 20 m. Palygorskite is most concentrated in the center of the basin
where it can form 75 to 85% of the beds. Minor chlorite, montmorillonite, dolomite,
and quartz are present in the clay beds. The major physil in the overlying alluvial
fans is montmorillonite. The authors (Galan et a[., 1975) believed the palygorskite
formed by direct precipitation from waters rich in Si, Al, and Mg. They refer to
similar Cenozoic deposits near Cadiz, Guadalquivir, Ebro, Vallecas, and Galacia
which are also believed to have formed by neoformation.
In the Tajo Basin abundant palygorskite and sepiolite occur in gypsum and
687

calcareous gypsum deposits as much as 250 m thick. Illite, kaolinite, Ch/S, and
smectite (both montmorillonite and stevensite) are present in varying amounts,
commonly comprising more than 50% of the physil suite (Garcia Palacios, 1977).
Detrital illite and kaolinite are most abundant in the edges of the basin and sepiolite
in the center. The palygorskite, sepiolite, and stevensite are believed to have formed
in the evaporite environment both by neoformation and aggradation of degraded
physils from the weathered zone.
Palygorskite and sepiolite also occur associated with continental Pliocene marly-
calcareous beds near Lebrija in southern Spain. Sepiolite is abundant in the lower
part of the section and gradually decreases upward as palygorskite and illite
increase. In the early stages the climate was relatively dry and conditions favored
the precipitation of sepiolite and calcite. Si was derived from diatoms, as suggested
by the presence of diatomite beds. As the climate became wetter, the amount of
detrital degraded illite and ions in solution supplied to the basin increased. Under
these conditions some of the illite, supplying the needed Al, was transformed to
palygorskite. At the same time some palygorskite grew from solution. This cycle was
repeated eight times (Galan and Ferrero, 1982).
In a summary of the Spanish Cenozoic deposits of sepiolite and palygorskite,
Galan and Castillo (1984) concluded they were all formed in lacustrine or peri-
marine environments. They formed by neoformation from solution and by transfor-
mation of chlorite and illite under semi-arid climatic conditions during periods of
tectonic calm.
There is a considerable difference of opinion as to whether palygorskite in any
given deposit or in most deposits is neoformed or transformed from smectite or
illite. There is little doubt that it forms both ways; however, my impression is that
major deposits form by transformation. Sepiolite forms by direct precipitation.
Other Spanish basins contain more typical detrital physil suites. In the Ebro
Basin, in the Navarra Province (south of the Pyrenees), the Eocene blue marls
contain a physil suite containing kaolinite, illite, montmorillonite, and sepiolite
(Arrese er al., 1966). The Upper Miocene continental deposits of the Granada Basin
contain muscovite and montmorillonite, along with minor amounts of paragonite,
chlorite, and kaolinite (Ortega-Huertas et al., 1979). Along the northeast border of
the Tajo depression, the Miocene sediments grade from sandstones near the edge to
shales and gypsum near the center of the basin. Montmorillonite, possibly formed
by the weathering of illite in the source area, is predominant; illite (I/S) and
kaolinite are also present. Towards the basin center illite becomes predominant as
montmorillonite decreases. The illite formed by aggradation of the degraded illite,
I/S, and montmorillonite (Martin et al., 1976). In the upper central part of the
Madrid Basin a deep-water lacustrine evaporite sequence, containing sodic salts,
contains layers of mudstone which contain 80 to 90% illite and minor amounts of
muscovite, kaolinite, chlorite, and I/Ch. Density currents are believed to be the
transport mechanism for the mud (Ordonez and Aquayo, 1982).
The Guadalquivir Basin in southwestern Spain opened into the Atlantic Ocean.
Montmorillonite is the predominant physil at the base of the Upper Miocene and
the top of the Lower Pliocene. Illite and kaolinite plus chlorite (mostly kaolinite)
688

increase towards the Miocene-Pliocene boundary. The pattern is believed to be the


result of the physils being derived from two sources and a regional transgression
climaxing near the end of the Late Miocene and beginning of the Early Pliocene.
The montmorillonite was derived from a continental source on the southeastern
flank of the basin where the climate was semi-arid; the illite, kaolinite, and
chlorite(?) were transported by marine currents from a humid tropical source. As
the transgression progressed, the physils from the distant source increased and then
decreased as the marine influence decreased. Locally, palygorskite and sepiolite
occur in lacustrine sediments near the top of the Pliocene (Latouche and Viguier,
1976).
The Lower Miocene marine marls and calcareous oozes of Sicily contain
montmorillonite, zeolite (heulandite-clinoptilolite), opal-A, and opal-CT. Radio-
laria, diatoms, and sponge spicules are abundant. There is no evidence that the
minerals formed from volcanic material (in place). The zeolites are believed to have
formed because of the high Si environment (Barbieri et al., 1981). The Pliocene
marine sediments of the Caltanissetta Basin contain carbonate rocks in the Lower
Pliocene and clays in the Upper Pliocene. The clay suite consists largely of illite,
kaolinite, and smectite with minor palygorskite, chlorite, and I/S. Palygorskite and
illite decrease upward in the section as smectite and kaolinite increase. The lower
physil suite is believed to have been derived from both Africa (Tunisia) and Sicily.
In the Upper Pliocene, tectonic movements increased the importance of the Sicilian
source (smectite and kaolinite) (Chamley, 1976).
Analyses of Upper Miocene Messinian and Pliocene samples from 1 2 wells
scattered throughout the Mediterranean and along the edges showed the contrast
between the Pliocene and Messinian physil suites:

smectite chlorite illite Ch/S kaolinite palygorskite


Pliocene 30 7 33 2 15 13
Messinian 54 10 21 3 8 4

The Messinian, which is composed largely of evaporite minerals, has a relatively


high content of smectite. The combination of minerals indicates a semi-arid environ-
ment with strongly contrasted seasons and suggests the smectite formed in the soils
developed on the low-relief, poorly drained coastal plain. Up to 25% chlorite, along
with minor palygorskite, is present in the dolomitic mark in the eastern Mediter-
ranean (Chamley et al., 1978). The authors suggest they could have grown in the
evaporitic sequence. The change in the physil suite from the Messinian to the
Pliocene may indicate a change in climate that influences weathering but probably
also is due to source changes influenced by the invasion of the Atlantic into the
Mediterranean and the rising water level.
The Cenozoic history of the Bohemian Massif (western Czechoslovakia and parts
of Austria, West and East Germany, and Poland) has a complex history of kaolinite
formation extending from the Ordovician to the Pliocene. An extensive kaolinitic
weathering crust, over the whole massif, developed from Late Cretaceous through
689

Early Eocene time. During the Middle and Late Eocene the area was transgressed
but kaolinization continued in the higher areas. The whole area was uplifted during
the latest Eocene and Early Oligocene, and kaolinization was more extensive. The
sea transgressed from the north during the Middle Oligocene and much of the
weathering crust was eroded. During the Late Oligocene and Miocene the sea
withdrew. Erosion occurred in the uplifted area and fringing deltaic deposits rich in
kaolinite were deposited. Volcanic rocks were deposited and altered to kaolinite and
montmorillonite. During the Late Miocene and Pliocene the whole area was again
uplifted and weathered fluviatile sediments were deposited. Laterites developed on
the volcanic rocks of the Vogelsberg area. Though the climate cooled following the
subtropical conditions of the Early Cenozoic, it remained warm and humid enough
up to the Pliocene for kaolin to be the major weathering product (StGrr, 1975).
Commercial Czechoslovak ball clays occur in western Bohemia and southern
Slovakia. These were deposited in Pliocene to Quaternary shallow lakes. The parent
rocks were slates and granites. The deposits of western Bohemia contain poorly
crystallized kaolinite and appreciable mica and illite. The southern Slovakia kaolins
contain a minor amount of smectite (Sindelar and Kraus, 1976). In central Slovakia,
Upper Miocene-Lower Pliocene rhyolite tuffs altered to montmorillonite in fresh-
water lakes. Deposits average 25 m thick. In some areas the rhyolite altered to
kaolinite (Kraus and Zuberec, 1976).
The upper Neogene sediments of the Polish Lowland contain an abundance of
beidellite. I/S and illite increase downward in the section and are believed to have
formed from beidellite in areas where the K content af pore water was relatively
high (near salt domes). Kaolinite is predominant in young Pliocene beds (Wyrwicki
and Wiewiora, 1981). The Upper Miocene sediments of the Carpathian Foredeep
(southern Poland) has a physil suite composed of varying amounts of I/S, illite, and
kaolinite. The amount of smectite in the I/S ranges from 60% to nearly 100% but
does not vary with the depositional environments, which range from hypersaline to
brackish. The I/S is believed to have formed during the continental weathering of
pyroclastic material (Srodo6, 1984).
The Upper Miocene and Pliocene sediments in the southern part of the Great
Hungarian Plain contain a physil suite composed of montmorillonite, kaolinite,
mica, chlorite, and amorphous material. Chlorite values are reported to be as high as
50%in the deeper part of the section (4500 m) (Varshnyi, I., 1975).
In the Tokaj Mountains (northeastern Hungary) montmorillonite, derived from
the weathering of Tertiary volcanics, is present in sediments ranging in age from
Eocene to Miocene. It becomes mofe abundant in the younger rocks. In the
Transdanubian Central Mountains of Hungary detrital poorly crystallized kaolinite
is abundant in the Eocene coal-bearing beds; varying amounts of illite and
montmorillonite are present. Illite and chlorite become increasingly more abundant
in the younger sediments, up through the Pliocene. Illite, I/S, and chlorite, with
lesser amounts of kaolinite and montmorillonite, are the major physils in the
Neogene of the southern Great Plain, in northeastern Hungary, in the Zala Basin
(southwestern Hungaq), and in the foreland of the Transdanubian Central Moun-
tains (Viczihn, 1975). The Neogene sediments of the Vienna Basin contain an I/S,
690

illite, kaolinite suite. The 1/S reflects a typical diagenetic sequence with illite layers
increasing from 25% to 80%with depth (2803 m) (Johns and Kurzweil, 1979).
In the Negev (southern Israel), marine limestones, chalk, and clay were deposited
from the Middle Cretaceous through the Early Eocene. Smectite comprises at least
80% of the physil suite. Palygorskite is locally present in brackish bay sediments.
From the Late Eocene to the end of Pliocene time sediments were deposited in a
variety of environments ranging from marine to eolian. The marine calcareous shale
section (Upper Eocene to Pliocene) contains physil suites composed of smectite,
illite, and kaolinite with minor palygorskite. In the lower part of the section smectite
is predominant (54%). Illite generally increases upward, to 48%, with kaolinite
varying between 21 and 44%. The physil suite reflects the interplay of two source
areas: smectite and palygorskite derived from local Cretaceous deposits and illite
and kaolinite supplied from the Nile River (derived from the Nubian sandstones).
In the Miocene-Pliocene freshwater lake deposits, the physil suite in the lower
part of the section is dominated by smectite with some kaolinite. In the upper part
of the section illite and kaolinite are predominant; only minor smectite is present.
The lower physil suite was apparently derived from Cretaceous-Eocene sediments
on the western shore of the lake and the suites in the upper part of the section were
derived from Nubian sandstones on the eastern shore. The shift in source is due to
tectonic tilting (Bentor et al., 1963).
Palygorskite is a common physil in the Cretaceous and Cenozoic limestones and
marls of the Middle East. It is also a common constituent of the soils. For
references see Wiersma (1970), Callen (1984), and Shadfan and Dixon (1984). Some
of the palygorskite in the soils is inherited from the underlying rock, some
transported by wind, and some neoformed. Most of the Cenozoic palygorskite
apparently formed in continental lacustrine-like environments under semi-arid
conditions. Present-day conditions favor the preservation of the fibrous physils in
the soil profile.

Africa

Millot (1970) has reviewed the literature on the origin of physils in the Cenozoic
basins of west Africa. During the Lower and Middle Eocene humid tropical
conditions prevailed and, as in much of Europe, lateritic weathering crusts devel-
oped in the interior of Africa. The western edge of the continent and more interior
depressed areas were invaded by the transgressing Atlantic. Kaolinite derived from
the weathering crusts was deposited in the coastal areas. Leached material was
carried to the lacustrine and shallow marine environments where it precipitated as
smectite, palygorskite, and sepiolite. Some of the smectite was probably detrital,
derived from the less intensely weathered, low relief, inland area.
Relatively pure palygorskite, with varying amounts of sepiolite, deposits are
present in sediments ranging in age from the Paleocene to Middle Eocene in
Morocco, Senegal, Ivory Coast, Dahomey, eastern Sudan, western Niger, and
Gabon. In Senegal-Mauritania the only physils in a bed 475 m thck are palygors-
691

kite and sepiolite. The chain physils are commonly associated with carbonate rocks,
which usually also contain montmorillonite and kaolinite. In general, kaolinite is the
dominant physil in the nearshore areas. Seaward, westward, montmorillonite and
palygorskite increase, with palygorskite increasing relative to montmorillonite. In
some areas sepiolite is present in the westernmost deposits. The general distribution
suggests palygorskite increases in a seaward direction and is neoformed in a marine
environment. The palygorskite beds are interbedded with phosphate and glauconite
beds which were deposited in shallow marine environments. Montmorillonite is the
major physil in the phosphate and glauconite beds. Phosphate and glauconite
apparently do not occur in the palygorskite beds, but detrital palygorskite can occur
in phosphate beds. The general rule for the phosphate series is that montmorillonite
is predominant in the lower part of the section consisting largely of open marine
carbonates. Upward, palygorskite increases in abundance, commonly being the only
physil in the youngest beds, dolomitic clays and phosphate beds, whch were
deposited in restricted environments (Boujo et al., 1980).
The situation is similar to that in the Miocene of the southeastern United States
where, in a peri-marine environment, marine phosphate beds occur interbedded
with brackish water-lagoonal palygorskite beds (Weaver and Beck, 1977). The
African coastal deposits, probably had a similar origin. Minor tectonic activity or
sea level changes can rapidly change coastal basins from marine to brackish.
Because one bed is marine does not mean the adjacent beds are marine. Also, some
of the African palygorskite beds contain appreciable kaolinite, whch suggests both
are probably detrital.
Following the extensive development of palygorskite, which was related to the
relatively high temperature conditions and tectonic quiescence that existed during
the lower Paleogene, the various western African basins were inundated with
kaolinitic detritus (Terminal Continental Series). As a result of renewed tectonic
activity in the lateritic highland areas, the initial influx of kaolinite ranged from the
Upper Eocene to the Miocene according to the tectonic evolution of the different
regions.
In both western Africa and Europe, much of the Paleocene and Lower and
Middle Eocene were times of tectonic quiescence. The climate was warm and
humid. Under this environmental situation lateritic weathering occurred in the high
areas. Residual kaolinitic crusts developed and relatively large volumes of Si, Mg,
Ca, K, and other soluble ions were transported to the low-lying fringing areas where
palygorskite and sepiolite formed. Rainfall must have been seasonal, as mild
evaporitic conditions prevailed in the low-lying basins. In Europe the chain clays
formed largely in alkaline lakes, whereas in Africa most of the depressions occurred
in coastal areas and waters apparently fluctuating from normal marine to brackish.
It is likely the palygorskite formed when the waters were brackish. Though the
sepiolite and some of the palygorskite was neoformed, it is likely much of the
palygorskite formed from the transformation of montmorillonite.
Under conditions where kaolinite develops in the high elevation areas,
montmorillonite commonly develops in the soils in the more flat-lying fringing areas
where leaching is more restricted and ions accumulate (Fig. 3-15). In turn, some of
692

Table 9-5
Physils in the i2 pm Fraction of Cenozoic and Recent Sediments in the Niger Delta. From
Lambert-Aikhionbare and Shaw (1982).
Formation 4 kaolinite 4 smectite 4 illite
Recent 35-60 30-50 10-15
Agbada Fm. 40-75 10-35 15-25
Akata Fm. 30-60 20-50 10-30

this montmorillonite is normally transported into the shallow lakes or peri-marine


lagoon-like depressions. During dry periods evaporation increases the Si and Mg
concentration to the level where palygorskite is the stable phase. Montmorillonite
transforms to palygorskite. Montmorillonite is the source of the A1 necessary for the
formation of palygorskite.
Montmorillonite is the predominant physil (80 to 100%) in the phosphate-
evaporite series of Tunisia. Varying, but minor, amounts of kaolinite, illite, paly-
gorskite, and sepiolite are present throughout the section (Lucas et al., 1979). In the
Upper Cenozoic gypsum-dolomitic marl sequence of Morocco (Le Jbel Ghassoul),
illite, chlorite, montmorillonite, and palygorskite are present in variable combina-
tions and amounts. Some of the fine-grained evaporitic dolomite beds contain
saponite and stevensite, which is believed to have formed from montmorillonite in
the high Mg waters (Trauth, 1977).
The Niger Delta contains a thick sequence of Cenozoic sediments similar to the
sequence in the Mississippi Delta. The physil suite of the shales and muds is
relatively uniform from the Eocene to the Recent (Table 9 - 9 , indicating weathering
conditions in the drainage basin have not changed significantly.In some wells
smectite persists to a burial temperature of 120°C. Other wells show a relatively
systematic conversion of smectite to I/S, with illite layers increasing with depth. At
least one well shows a relatively abrupt increase in the proportion of illite layers
over a short interval. The factors controlling the rate of burial metamorphism in the
Niger Delta are not yet evident (Lambert-Aikhionbare and Shaw, 1982; Braide,
1982). The relatively low content of smectite in the physil suite may account for
some of its erratic behavior.

South America

The Cenozoic sediments of South America are primarily continental, derived


from the Andes. The Andes and associated ranges contain much volcanic material.
During the Miocene some marine sediments were deposited near the mouth of the
Amazon and east-central Argentina.
In the Resende Basin, Rio de Janeiro State, Brazil, the fluvial Miocene to
Pleistocene sediments contain primarily smectite in the lower part and kaolinite in
the upper part. The change in physils presumably reflects an increase in rainfall
during more recent times, probably from semi-arid to humid (Amador and Zalan,
693

1980). A large commercial Pliocene kaolinite deposit, up to 85 m thick, occurs along


the Rio Jari, near the mouth of the Amazon River. The deposit is believed to have
formed by the alteration of arkosic sands deposited in either a deltaic or lacustrine
environment (Murray and Partridge, 1981).
The Eocene and post-Eocene sediments of northwestern Venezuela (Lake Mara-
caibo region) grade from fluviatile in the northwest to bathyal in the southeast. The
fluviatile facies expands in the younger sediments. Analyses of samples from 14
wells indicate kaolinite is the predominant ( > 40%) physil in the fluviatile sedi-
ments; varying amounts of illite and smectite are present. Chlorite and I/S increase
seaward as kaolinite decreases; illite remains relatively constant. The bathyal physil
suite averages 60% I/S, 25% illite, 10%chlorite, and 5% kaolinite. The proportion of
illite layers in the I/S does not increase regularly with depth, indicating some of it is
detrital; however, in the deeper sections there is a progressive increase in illite layers
(to 7:3 ratio) with depth, indicating burial metamorphism has been effective
(Weaver).
It is of interest to note that in northwestern Venezuela detrital pyrophyllite is a
major component, 20 to 70%, of clay beds ranging in age from Cretaceous to
post-Eocene. Illite and minor kaolinite are commonly associated with the pyrophyl-
lite (Palacio, 1967). The pyrophyllite was apparently derived from schists in the
mountains to the south.

PLEISTOCENE

During the Pleistocene, glacial till and glacio-lacustrine sediments were deposited
over much of northern North America, Europe and Asia. These deposits are largely
fine-grained scoured material that reflect the composition of the source rock. Post
depositional weathering has modified the deposits to some extent, but in general,
the physils can be related to the source rocks.
Extensive eolian loess deposits occur south of the glaciated regions of North
America and northern Europe. There are extensive deposits in northern China,
flanking the Gobi Desert, and in the steppe region of Siberia. Smaller deposits occur
in a number of areas. Loess is a relatively homogeneous, nonstratified, unconsoli-
dated deposit consisting largely of silt-sized grains but with a fair amount of
clay-sized material. Either illite or smectite is generally the dominant physils. The
wind-blown material is believed to have been derived from glacial outwash and
deserts (China and Siberia).

PALEOATLANTIC PHY SILS

History

Briefly, the plate motion that led to the formation of the Atlantic Ocean started
by the opening of rifts and outpouring of basalt during the Triassic, about 200 m.y.
694

ago. As drifting continued, Laurasia and Gondwana separated and the first marine
sediments were deposited in the North Atlantic during the Middle Jurassic (Cal-
lovian, 155 m.y. B.P.). By the end of the Jurassic period, South America and Africa
began to separate to form the South Atlantic. Africa gradually rotated counter-
clockwise, narrowing and finally closing the Tethys.
During the early stages of ocean development evaporite deposits were relatively
widespread and continued to be deposited until Middle Cretaceous time. As the
Atlantic continued to widen, normal marine conditions became more extensive.
However, during Middle Cretaceous time several major anoxic events occurred: late
Barremian-Ablian, late Cenomanian-early Turonian, and Coniacian-Santonian
(minor). Large amounts of organic carbon were deposited and preserved. Much of
the world’s oil has been generated in these black shales. The condition that allowed
the development of these organic-rich muds is believed to be a widespread, warm
equable climate, which allowed a high production of organic material and ocean
basins that were periodically salinity stratified (evaporation). Limited circulation
and increasing salinity with depth produced the anaerobic conditions that allowed
the preservation of much of the organic water (Arthur and Natland, 1979).
By the latest Cretaceous, circulation between the North and South Atlantic was
well established and oxic conditions generally prevailed. In the North Atlantic,
active spreading of the Mid-Atlantic Ridge separated Greenland and Scandinavia
about 60 m.y. B.P. (Paleocene) and by the Lower Eocene (50 m.y. B.P.) Greenland
separated from Spitsbergen, allowing cold polar waters to enter the North Atlantic
and initiate the present deep and bottom water circulation pattern. In Miocene time
(18 m.y. B.P.) the Tethys was severed by the junction of Eurasia and Africa, forming
the Mediterranean Sea which was partially isolated from the Atlantic. This further
modified the North Atlantic circulation pattern. The initiation of glaciation about 3
m.y. ago was responsible for the formation of the cold Labrador Current which
displaced the warm, northward-flowing branch of the Gulf Stream (Berggren and
Hollister, 1974). The shifting currents throughout the Mesozoic-Cenozoic influenced
the distribution of the physils in the Atlantic.
The Mesozoic-Cenozoic section of the western North Atlantic has been divided
into six formations (Jansa et al., 1979) which can be correlated with equivalent units
in the eastern North Atlantic (Jansa et al., 1977). The Upper Jurassic and most of
the Lower Cretaceous rocks are primarily pelagic limestones, argillaceous in the
lower part. These are overlain by organic black shales of Aptian to Cenomanian age
deposited under anoxic conditions. The younger sediments are primarily clays, silty
clays and siliceous clays deposited under oxic conditions. During Late Paleocene
through Middle Eocene time, biogenic silica was deposited in the deeper North
Atlantic; calcareous-siliceous sediments were deposited at shallower depths. The
siliceous material is primarily opal-cristobalite. The high production of siliceous
organisms is believed to be due to an increase in volcanic activity (source of Si) and
the introduction of cold Arctic bottom waters which caused the upwelling of
nutrients.
695

Physils

On the basis of studies of Holocene physils in the oceans, we know that some are
autochthonous, having been formed from volcanics, and some are detrital. Both
types of physils are present in the Mesozoic-Cenozoic sediments in the ocean basins.
The problem is to determine the relative abundance of the two types. There is a
difference of opinion.
Kossovskaya et al. (1975) concluded from a study of three DSDP cores from the
western and eastern North Atlantic (Sites 8, 9 and 12, Leg 2) that most of the
Cretaceous and Cenozoic physils in the North Atlantic formed by epigenetic
alteration of basaltic material. In sediments ranging in age from Miocene to
Cretaceous, ferrimontmorillonite (with 20-258 illite and 10%kaolinite) is by far the
predominant physil. The iron content of the ferrimontmorillonites is typically in the
range of 6 to 13%,which is appreciably more than for most other montmorillonites.
The K,O content is high, ranging from 2.5 to 3.58, suggesting the physil is an I/S.
The high Fe and K values are characteristic of smectites-I/S forming from deep
ocean basalts at slightly elevated temperatures (see Chapter VI). The clays com-
monly have a breccia-like fragmentary structure which the authors believe indicates
they formed from hyaloclastics. They cite the presence of abundant ferric oxide and
MnO as further evidence that the smectite formed epigenically. Further, they note
that much of the ferrimontmorillonite has an acicular form and that the proportion
of acicular material increases with depth.
Clauer et al. (1984) examined Cretaceous and Paleogene samples from the same
general area and concluded that the smectite, identified as beidellite, was detrital
and of pedologic origin. They found authigenic laths growing on the “fleecy”
detrital smectite particles. Chemically the two types of smectites are very similar.
Isotopic studies indicated the laths crystallized in the pores shortly after deposition
but after contact with sea water was lost. Thus, it appears the detrital smectite
adjusts to its new environment by recrystallizing, producing a major morphological
change but little chemical change. The laths are presumably the acicular material
observed by Kossovskaya et al. (1975).
The interpretations conflict. Some background information may help resolve the
conflict. First, studies of recent deep sea physils indicate that most of the smectite
forming from basaltic material has a hydrothermal origin and only a relatively small
percentage of the basalt alters to smectite. Second, during the Mesozoic and Early
Cenozoic the Atlantic Ocean was smaller than at present and almost completely
surrounded by large land areas. Volcanism was probably more extensive on the
continents than in the deep sea. An appreciable amount of detritus must have been
transported from the continents to the sea, and much of this material was likely
smectite formed by the continental weathering of volcanic material.
Kastner (1981) plotted the percentage of lithologic units, in DSDP cores from all
+ +
oceans, containing clinoptilolite palygorskite smectite versus age (Fig. 6-17).
She considers this to be an authigenic mineral assemblage. The relative abundance
of this assemblage is at a maximum, 40%,in the Late Cretaceous. It is less than 10%
for most other periods. As much of the palygorskite is probably detrital, it is
696

unlikely that the 40% value represents the volume of authigenic silicates in the
Upper Cretaceous deep sea sediments.
In contrast to the conclusions of Kossovskaya et al. (1975), most scientists who
have studied the Mesozoic-Cenozoic physils of the Atlantic concluded the physils
are largely detrital and reflect climatic conditions, tectonic activity and current
patterns (e.g., Chamley, 1979; Chennaux et al., 1985).
Based on a study of the grain-size distribution of quartz silt (2-6 pm, moderately
well sorted and negatively skewed) in North Atlantic pelagic samples, ranging in age
from Early Cretaceous to Late Miocene, Lever and McCave (1983) concluded that
the majority of the physils and quartz in the < 2 p m fraction was atmospheric dust
derived from both Africa and North America. Maximum input occurred between
paleolatitudes 20-30"N.

Mesozoic
The oldest sediments along the western edge of the Atlantic are Middle to Late
Jurassic age (Hollister et al., 1972). Site 100 was drilled at the northern edge of the
Bahama Platform. The hole bottomed in basalt. The basalt is overlain by 43 m of
greenish-gray limestone containing plant fragments. This grades upward into a red
clayey limestone approximately 32 m thick. The depositional environment is con-
sidered to be upper bathyal.
The physils (Weaver) in the section immediately overlying the basalt consist of
A
I/S with a ratio of 1:l.The phase produces a broad 10 peak and no 5 A peak and
is presumably Fe-rich. It may be a form of glauconite or celadonite. The upper
portion of the greenish-gray limestone contains I/S (2:3), relatively well-crystallized
illite, kaolinite and palygorskite (near the top). The red limestone contains I/S, with
ratios ranging from 1:4 to 3:2, illite, palygorskite, and traces of kaolinite. Palygors-
kite is the dominant physil in some samples. The physil suite contains no
montmorillonite and appears to be primarily detrital.
The Jurassic was also cored in Hole 105, due east of Cape Hatteras on the
continental rise. The physil suite consists largely of illite and montmorillonite. The
former is predominant in the lower part of the section (Oxfordian) and the latter in
the upper part of the section. There is no information on the nature of the
montmorillonite (Zemmels et al., 1972). The physils would appear to be detrital and
reflect a source area in a more temperate environment than the source area for the
southern well (Hole 100). Thus, though basalt formed the floor of the narrow
Atlantic Rift it does not appear that an appreciable portion of the overlying physils
were formed by the submarine alteration of the basalt to smectite.
DSDP Site 330, located on the eastern end of the Falkland Plateau, penetrated a
Middle to Upper Jurassic sequence that in Jurassic time was adjacent to the coast of
South Africa. The basement is continental gneissic and granitic rock on which a
kaolinite-rich soil was developed. The overlying Jurassic section was deposited
during a shallow water marine transgression of the basement. The environmental
sequence ranges from coastal swamp, to beach, to shelf (Barker and Dalziel et d..
1976). Kaolinite and illite are the predominant physils in the lower part of the
section. Montmorillonite is present in moderate amounts. I t increases in abundance,
697

as kaolinite decreases, in the shelf sediments (silty clay); illite remains relatively
abundant (40 to 70% of the physil suite) (Zemmels et al., 1976). The illite and
kaolinite were presumably derived from the adjacent African continent. The upward
increase in the montmorillonite/kaolinite ratio may indicate a moderation in
weathering intensity or a differential settling of physils-kaolinite deposited near
shore and montmorillonite on the open shelf.
Smectite is the most abundant physil in the deep sea Cretaceous sediments as it is
in the Cretaceous epicontinental deposits, where much of it formed by subaerial
weathering. Chamley (1979) reported that most of the Atlantic smectites are the
A1-Fe type with associated V and Li contents not representative of volcanogenic
origin. These data, along with the rare earth content, led him to conclude that most
of the smectite in the Atlantic was terrigenous.
Illite is commonly the dominant physil in both the southern Atlantic and parts of
the Pacific Oceans. Kaolinite is relatively abundant off the southwest coast of
Africa. Palygorskite is common in sediments of Aptian age and younger. In most
instances it occurs offshore of continental deposits and much, if not all, is ap-
parently detrital. An organic-rich black shale facies is widely distributed in the
Atlantic. The physil suite is similar to that in the adjacent oxidized sediments
(Chamley, 1979).
At least in the Atlantic, the Cretaceous physils have a latitudinal zonation,
presumably related to climatic conditions. This is best illustrated in a series of wells
along the western coast of Africa (DSDP Legs 40 and 41). Off the tip of South
Africa the physil suite is composed of illite (40 to 95%) and kaolinite (up to 50%).
To the north illite remains predominant, commonly 80 to 90%, kaolinite decreases
and montmorillonite increases but is < 20%. The sediments are largely turbidites
and organic clays. The physils suggest a relatively high relief with moderate rainfall
in the source area.
Farther north, off the northwest African bulge, montmorillonite becomes the
predominant physil. Varying amounts of illite, I/S and kaolinite are present.
Northward palygorskite and chlorite (minor) increase and kaolinite decreases. Near
the northern portion of the bulge palygorskite is similar in abundance to montmoril-
lonite, though illite, I/S and kaolinite comprise the bulk of the physils in the Lower
Cretaceous. The sediments are primarily turbidites interbedded with pelagic clays;
the physil suites (smectite and palygorskite) suggest a relatively low relief source
with warm, arid to semiarid conditions and variations in seasonal humidity.
Weathering was moderate and palygorskite apparently formed in brackish coastal
waters and/or alkaline lakes (Chamley, 1979).
Off the coast of central Portugal (Site 398) and in the Bay of Biscay (Site 400)
montmorillonite becomes predominant, comprising 60 to 90% of the physil suite.
Illite becomes relatively abundant in the uppermost Cretaceous (Chamley, 1979).
The climate was apparently less arid and cooler than in North Africa where
palygorskite formed. Montmorillonite appears to be relatively more abundant than
on the European continent, suggesting that some of it formed from volcanics in the
marine environment or after shallow burial. Fig. 9-36 shows the distribution of
physils in a typical North Atlantic well.
698
SITE 400 I
LEG 48 DSDP CLAY MINERALOGY

SAMPLES COLOR . CLAY MINERALS ASSOCIATE MINERALS

1-1-85l1021
2-2-105
3-3-13
4-4-88
5-1-80
6-2-78
7-2-45

9-3-20
9-1-11
10-2-86
11-2-5

13-3-53
13-6-63145.60)
14-2-38
15-4-75

17-1-43

20-5-132136)
I I I I I
23-4-50
24-6 -12711071

26-3-88

3l-1-124 161,1001

IIIII

51-5-671771
sz.?.i36u~,n9)
53-1-Ul5.121
59-3-951741
55-2.78177. E7J

57-2-321301

58-2-81(49.881
69.1-106
-
Fig. 9-36. Physil distribution in DSDP Site 400, Bay of Biscaye. From Chamely, 1979, Deep Drilling
Results in the Atlantic Ocean: Continental Margins and Paleoenvironment. Copyright 1979 Amer.
Geoph. Union.
699

The physil zonation along the western flank of the Atlantic is generally similar to
that along the east flank, though in the southern latitudes kaolinite is less abundant
and montmorillonite is more abundant than along the east flank; illite is present in
moderate amounts. The high montmorillonite content may be related to volcanism
in the southern Andes (Baker, Dalziel et al., 1976). Palygorskite is apparently
present, in abundance, farther south than along the eastern South Atlantic. Roberts
(1981) reports 40 to 50% (of the physil suite) palygorskite in the Upper Cretaceous
from near the Rio Grande Rise ( - 28"s latitude). Roberts suggests that during the
Upper Cretaceous the eastern portion of the RIO Grande Rise subsided slightly,
forming a semi-enclosed basin in which the palygorskite formed. With continued
tilting and subsidence, the palygorskite was reworked and transported seaward as
the environment favorable for the formation of palygorskite migrated westward.
Chamley (1979) notes that there is a short-term increase in the amount of illite
and chlorite in the Late Cretaceous and attributes this to the separation of Canada
and Greenland which created strong north-south marine currents. These currents
transported physils from the cold region of minimal weathering, south into the
central Atlantic, where montmorillonite was forming in a moderate weathering
environment.
Though the Cretaceous global temperatures are believed to have been relatively
uniform, the latitudinal zonation of the deep sea physils indicates there are signifi-
cant latitudinal variations in climate. The zonation also demonstrates that a major
portion of the marine physils is detrital.
Well-crystallized smectite is by far the most abundant physil in sediments of Late
Jurassic to Late Paleogene age. Chamley (1979) believes the abundance of detrital
smectite is due to the wide extent of low relief land areas, a hot climate and seasonal
changes in humidity. It is also likely that the relative abundance of volcanic material
on the continents was a contributing factor.

Cenozoic
Smectite is the major physil in the Paleogene sediments, but palygorskite-sepiolite
are locally abundant along the east coast of South America and the west coast of
Africa and southern Europe. Much of this material is detrital and was apparently
formed in semi-closed and low-lying basins around the margins of the Atlantic.
Much of the palygorskite-sepiolite was reworked and transported seaward; in some
instances the marginal basins may have subsided and become part of the open
ocean. (In the Blake Plateau Region Lower Cretaceous tidal-flat deposits are now
beneath 2607 m of sea water and 99 m of younger sediments (Enos and Freeman,
1979).) The distribution of the chain physils indicates they were formed where the
climate was warm to hot and alternately wet and dry (Weaver and Beck, 1977;
Chamley, 1979).
From the Upper Eocene to the Late Pleistocene there is a general increase in the
illite and chlorite (primary minerals), associated mixed-layer physils, quartz and
feldspar; smectite decreases (Fig. 9-36, 9-37) (Chamley, 1979; Latouche, 1978). In
the Pleistocene sediments montmorillonite comprise c 20% of the physil suite in the
northern part of the North Atlantic (DSDP Leg 80). This trend is related to
700

CLAY
S I T E 550 S I T E 549 S I T E 548
MINERAL
+ ' +

YIDDLE 40

,OWER 40 40
'LIOCENE
0

JPPER 40 40 40
{IOCENE
0 0 0

MIDDLE & LOWER


MIOCENE
_-___---
OLIGOCENE

---------
JPPER
5OCENE

----
80 80
PALEO- 4o
40
CENE
0 0

RQ 0 rn m
SMECTITE ILLITE KAO. CHLORITE KAO.+
CHLORITE

Fig. 9-37. Physil distribution in Cenozoic sediments in three wells in the northern North Atlantic. From
Chennaux ei nl., 1985.

world-wide cooling and an increased deep ocean circulation caused by the move-
ment of cold polar waters into the Atlantic (opening of the Norwegian Sea).
Evidence for these two processes was obtained from oxygen isotope studies of deep
sea fauna and the presence of hiatuses in the deep sea Cenozoic record (Roberts and
Montadert, 1979).
There was a further drop in bottom water temperature in the North Atlantic
approximately in the Middle Miocene. This coincided with a subsidence of the
Iceland-Faeroes Ridge and an increased southern flow of cold water (Roberts and
Montadert, 1979). In samples from Leg 80, slightly southwest of Great Britain,
701

Chennaux et al. (1985) detected a relatively abrupt increase in the illite and chlorite
at the approximate boundary between the Middle and Upper Miocene (Fig. 9-37).
This presumably reflected an increase in the influx of cold water physils. Weaver
and Beck (1977) noted that in the southeastern United States the crystallization of
palygorskite stopped abruptly in the Middle Miocene. Faunal data indicates this
coincided with a decrease in water temperature.
The Cenozoic physils in the South Atlantic show a trend similar to those in the
North Atlantic. As can be seen in Fig. 9-36, the primary minerals began to increase
in abundance in the late Eocene to Oligocene time interval and continued to
increase in younger sediments. The change in the physil suite is apparently related
to climatic cooling and an increase in marine current activity (Robert, 1980). The
simultaneity of the change in the composition of the physil suites in the South and
North Atlantic suggest the decrease in temperature was global in scope. The
increase in kaolinite in the younger sediments near the equator demonstrates the
overriding effect climate can have on the distribution of physils.
Jacobs (1974) found that in the Antarctic deep sea sediments an increase in illite,
and to a lesser extent chlorite and kaolinite, started in the early Miocene, reached a
maximum (40-45% illite, 35-50% montmorillonite, 5-158 chlorite and kaolinite

CRETACEOUS AND CENOZOIC AVERAGE VARIATIONS


1 NE ATLANTIC

lite (I ) Sepiolite (Se)


hlorite(C) Mixed- Attapulglte (At) Cristobalite Smectite
tldspors layers Kaolinlte (Palygorskite) Zeolites (sm)
mphiboles (M-L) (K) Quartz

Temperate Warm o Cold Volcanic El Volcanic


Warm-humid El Desert 0 Temperate,
wor
4 Y / 0

Increase in areas

Fig. 9-38. Geographical changes in average physil mineralogy in northeast Atlantic Ocean; climatic
influence is indicated. From Chamley, 1979, Deep Drilling Results in the Atlantic Ocean: Continental
Margins and Paleoenvironment. Copyright 1979 Amer. Geoph. Union.
4
8
Table 9-6. Chains of causes responsible for some of mineralogical events recognized in North Atlantic sediments. From Chamley (1979).
Period Region Mineralogical Main cause Secondary possible causes
event
Middle-earl y Everywhere Irregular increase of CLIMATE CURRENTS TECTONICS PHYSIOGRAPHY
Paleogene to Illite, Chlorite. Quartz. (World cooling; less (increase of (relief rejuvenation) (vegetation partial
Pleistocene Feldspars. mixed-layers seasonal contrasts in deep-sea denudation)
humidity: sea-level meridian
lowering) circulation)
Early Eocene N E Atlantic Appearance and increase PHYSIOGRAPHY CLIMATE TECTONICS CURRENTS
(at least) of Sepiolite and (closed or semi-closed (hot; strong (marginal instability) (local: turbidites)
Attapulgite (Palygorskite) marginal basins) contrasts in
humidity)
Santonian to N E Atlantic Strong increase of TECTONICS CURRENTS CLIMATE PHYSIOGRAPHY
Maastrichtian (at least) Primary Minerals: lllite (global event: (general: meridian (cooler under northern (continental
and associate Minerals separation of Canada circulation of latitudes) geological
and Greenland?) sea-water masses) environment)

Albian N E Atlantic Sudden appearance of PHYSIOGRAPHY CLIMATE TECTONICS CURRENTS


(at least) Attapulgite (Palygorskite) (closed or (hot; contrasts (marginal instability) (local:
semi-closed marginal in continental differential settling
basins) humidity) or turbidites)

Most of Everywhere Large abundance of CLIMATE PHYSIOGRAPHY TECTONICS CURRENTS


Cretaceous well-crystallized (warm; - extension of
(large (marginal instability) (local:
and Smectite rather arid with short flattened continental differential settling)
Paleogene humid seasons) areas; hydromorphic
soils)
Middle upper NW Atlantic Strong and fast supply TECTONICS PHYSIOGRAPHY CURRENTS (CLIMATE?)
Jurassic of Minerals inherited (oceanic initiation) (continental (local) (temperate and
from rocks and soils geological humid?)
(mineralogical mixture) environment)
703

each) in the Upper Miocene and then remained constantly high. The illite in the
latter samples has a much higher degree of crystallinity (narrower peak width) than
that in the montmorillonite-rich (72-86%) older sediments. The change in the physil
suite is believed to reflect a decrease in temperature and the development of the
glacial ice sheet on Antarctica. Chemical weathering was curtailed and the ice
transported unweathered detritus to the surrounding ocean.
In both the Pacific and Atlantic Oceans there is a general increase in illite and
chlorite starting at times ranging from Late Miocene to Pleistocene. This increase is
related to global cooling and some of its by-products, e.g., glacial scour, loess,
increased intensity of cold currents flowing toward the equator.
It is apparent, that aside from some (minor?) smectite, the physils filling the
Atlantic since the Jurassic were largely derived from the continents. The lateral and
temporal distributions of the various physils are closely related to both latitudinal
and global climates. The physil distribution has been modified by deep ocean
currents operative since the Late Eocene. The current pattern has been largely
determined by movement of the plates, primarily closing of the Tethys Ocean and
opening of the Arctic Ocean. In Fig. 9-38 Chamley (1979) has summarized the
average distribution of Cretaceous and Cenozoic physils in the northeastern Atlantic
Ocean and in Table 9-6 indicated some of the factors which have affected temporal
changes in the physil suites.
705

Chapter X

LITHIFICATION AND PETROLOGY

Mostly what I want to talk about in this chapter are the physical changes that
occur as a mud is converted to a physilite, primarily shale. The chemical changes
that occur during lithification are discussed in other chapters. First, the bulk
changes are discussed and then, the petrography.

Compaction

Compaction in argillaceous sediments has been thoroughly reviewed in a book by


Rieke and Chilingarian (1974). For other reviews, see Meade (1966). In addition to
the geologic literature, the soil science literature contains a large amount of
information on compaction, particularly at shallow depths.
Once physil plates, flocs, and aggregates of various types settle to the bottom of a
body of water, the physil concentration increases and the charged physils with
various coatings of absorbed organic material, oxides, and hydroxides, coalesce to
form a continuous network-mud. These muds have a porosity between 70 and 90%
(50-80% water by weight) and interstitial water and ions readily exchange with the
overlying water and ions. As physils accumulate, the overburden pressure causes
compaction by particle rearrangement and deformation.
Porosity decreases rapidly for the first 500 m and then slows. The rate of porosity
decrease, compaction and consolidation is influenced by a number of factors,
mainly type of physil, absorbed cations, composition and pH of interstitial fluids,
decomposition of organic material, chemical diagenetic processes, rate of deposition
and availability and distribution of permeable zones (Chilingarian, 1983).
As a high electrolyte concentration decreases the size of the electric double layer
and promotes coagulation, marine slurry-muds commonly have a lower porosity and
water content than fresh water sherry muds.
Fig. 10-1 contains curves showing changes in mud-shale porosity for a number of
stratigraphic sections and experimental values for kaolinite, illite and montmoril-
lonite. The experimental studies with the pure physils suggest that porosity and
porosity reductions are a function of particle size: kaolinite > illite > montmorillonite
(Chilingar and Knight, 1960). The curves also indicate that porosities for labora-
tory-compacted clays are considerably higher than naturally compacted shales and
that porosities for Paleozoic shales are less than those for Cenozoic shales. In part,
this is a function of mineral composition and grain size. The natural shales have
706

- Dickinson (1951)
Gulf Coast Tertiary
-- Hedberg (1936)
- Venezuela Tertiary
Meade ( 1 9 6 6 )
Comcosite- Recent M i o
- Athy (1930)
Oklahoma -Paleozoic
----- Classen (Unpub.)
Gulf Coast -Tertiary
....... Magara (1968)
Japan - Tertiary
-Chillingar and Knight
(1960)experimental

Porosity (*/.I
Fig. 10-1. Selected curves showing changes in “shale” and clay porosity with depth. From Hinch, 1978,
reprinted by permission of American Association of Petroleum Geologists.

more quartz than the laboratory clays and the Paleozoic shales have more illite and
less montmorillonite than the Cenozoic shales.
In addition to mineral composition and particle size, time and temperature
appear to be factors in porosity reduction in physilites. To a large extent, time and
temperature tend to increase particle size and decrease surface area. Well data for
montmorillonitic shales of nearly identical composition in sections ranging in age
from Pliocene-Miocene to Mississippian indicate shale density increases with age
(Weaver, 1968). When porosity-temperature trends for wells from various areas are
compared to porosity-depth trends, there is closer grouping of trends for the former
than the latter. This suggests temperature plays a significant role in the process of
pore reduction. The influence is most pronounced at temperatures higher than
12OoC, suggesting chemical diagenesis is a factor (McCulloh et a/.,1978).
Hinch (1978) summarized the results of Amoco Production Company’s studies of
porosity reduction in Gulf Coast mud-shales. Hinch concluded that while gravita-
tional compaction is probably the dominant cause of the rapid loss of porosity at
depths shallower than 1000 m, some other mechanism is operative below 1000 m.
He believes that cementation, pressure solution, and recrystallization do not play a
significant role in the reduction of shale porosity and fluid expulsion in the Gulf
Coast, but that a thermophysical process is a major mechanism.
Gulf Coast Cenozoic shales have approximately 80,000 times the surface area of
an equal volume of sand. Calculated pore size decreases from 100 A at 1000 m to
about 25 A at about 4000 m. As a consequence, shale pore water is either absorbed
or close enough to grain surfaces to be structured to some degree. Structured and
absorbed water is bonded to the physil surface and is more difficult to remove than
the “free” water encountered in larger pores. Thus, in the deeper parts of shale
sections, where pores are small, an increase in temperature, thermal energy, is
required to mobilize the structural water so that it can be squeezed by overburden
pressure into more porous beds or fractures. This is a reasonable explanation for the
Gulf Coast area where chemical diagenesis consists largely of the conversion of
smectite to I/S. However, at higher temperatures ( > 200°C) pressure solution,
recrystallization and crystal growth are the major factors that reduce porosity to less
than 5%.
The Gulf Coast mud-shales and a large number of mud-shale sections from other
areas are overpressured and undercompacted. In most sedimentary sections, poros-
ity and water content of physilites decreases systematically with depth; however, in
many sections, particularly where sedimentation was rapid and the physilite content
high, the trend is interrupted by a zone containing abnormally high porosities and
pressures. Overpressuring may be due to the development of permeability seals,
release of smectite interlayer water, and or organic maturation, amont other things
(Rieke and Chilingarian, 1974; Weaver and Beck, 1971).
One of the main consequences of compaction that is of interest to the geologist is
flake orientation. Orientation is something that can be measured and provide some
insight to the depositional and tectonic history of a physilite. It has been generally
accepted that under fresh water conditions physils are not flocculated and settle
face-to-face with fairly good orientation. Under marine conditions, the physils are
presumed to flocculate and form a more open cardhouse or honeycomb structure.
However, the association is not that simple and the factors that control the
depositional fabric of physils are not well understood.
O’Brien (1971) observed with the electron microscope, that there was little
difference in the fabric of kaolinite or illite floccules formed in saline or salt free
water. Fig. 10-2 shows the similarity of floccules of illite formed in salt water and
distilled water. The floccules are composed of units of face-to-face bonded flakes
that are organized in an edge-to-face arrangement. There is some difference in detail
but the interpretation is somewhat subjective. SEM pictures of Pliocene and
Holocene marine sediments (OBrien et al., 1980) indicate the turbiditic clays have
an edge-to-face flocculated fabric, whereas hemipelagic clays have a fairly well
developed parallel orientation (Fig. 10-3). The turbiditic flocculated material pre-
sumably formed in a relatively concentrated turbidite “cloud” and the hemipelagic
clays formed from the settling of scattered, isolated physil flakes. This illustrates
that clay fabric is not necessarily related to salinity. The importance of organic
material and pelletizing organisms in clay sedimentation (p. 295) further indicate
that the fabric of many muds may not be closely related to water chemistry or
currents. Bioturbation further complicates the situation.
One would expect that overburden pressure would cause preferred orientation in
708

Fig. 10-2. Illite floccules formed in distilled water (top) and salt water (1 gm/l NaCI) (bottom), freeze
dried. The floccules are basically similar. The sketches illustrate the fabric of illite (top) and kaolinite
(bottom) floccules. Both show a combination of E-E and E-F bonding. Bar = 2 pm. From OBrien, 1971.
Copyright 1971 The Clay Miner. Soc.Courtesy N.R. OBrien.

muds but the early experimental data are far from conclusive (Rieke and
Chilingarian, 1974). In general, the experimental data suggests applied pressure does
produce an increase in particle orientation (measured by x-ray peak intensity)
(Martin, 1962; Quigley and Thompson, 1966). More recent TEM and SEM studies
709

Fig. 10-3.SEM pictures showing face-to-face orientation of hemipelagic clays (c) and open edge-to-face
arrangement of turbiditic clays. (d) Bar = 1 pm. From OBrien et al., 1980. Copyright 1980 Elsevier Pub.
Co. Courtesy N.R. OBrien.

have confirmed this conclusion. TEM studies of samples with a high void ratio
(3-4) from the Gulf of Mexico and compacted in the laboratory showed that at the
relatively low overburden pressure of 392.3 kPa (4.0 kg/cm2) and void ratio of 1.3,
the particles were forced into clumps and packets which are randomly oriented. At
710

load pressure of 32 kg/cm’ and higher, the clumps and packets (domains) devel-
oped a high degree of parallelism (Bowles et ul., 1969).
The various studies indicate that under relatively low pressures, the physil flake
with EE and EF association are reorganized into F F units which have been referred
to as stacks, tactoids, packets, clusters, books, aggregates, domains (see Bennett et
al., 1977).
SEM studies of microfabric features in sediments from marine, brackish, and
fresh water, flood plain, lacustrine, aeolian, and glacial deposits (Collins and
McGown, 1974) showed that there was no unique relationships between microfabric
and depositional environments; however, there may be a dominant fabric feature or
set of features characteristic of any one sediment. Commonly, different types of
microfabric features were observed in any one type of sediment.
TEM and SEM studies of core samples of Mississippi prodelta sediments (Faas
and Crockett, 1983) confirm that orientation occurs under relatively low overburden

‘Single plate like’


particles and chains

Very high void ratio


> 3.0

Domain particles
and chains

High void ratio


> 2.5

Medium to high
void ratio
-1 5 - 2 5

Low void ratio


<I 5

=- Very low void ratio


<I 2

Fig. 10-4. Proposed clay fabric model for smectite-and-illite rich submarine sediments. From Bennet et
al.. 1981. Copyright 1981 Soc. E o n . Paleo. Miner.
71 1

--
Fig. 10-5. SEM showing changes in deep sea clay fabric (south Atlantic) with depth. Upper sample at a
depth of 10.8 m contains 54% water and has a void ratio of 2.8, both EE and EF, as well as FF bonding
is evident. Middle sample, 73.5 rn, contains 47% H,O and V.R. of 1.56. F F orientation is predominant in
deeper two samples where domains are developed. Bar = 1 pm. From Faas and Crocket, 1983.
712

pressures. Fig. 10-4 illustrates the sequence of clay fabric development with increas-
ing depth. Individual flakes rotate as water is expressed to form F-to-F oriented
particles, small domains; these coalesce to form larger composite domains. Random
orientation changes to preferred orientation at 250 kPa overburden pressure ( - 50 m
depth) as a result of linkage destruction and incorporation of the linking material
into the larger domains, surrounded by a smaller-particle homogeneous matrix
(Faas and Crockett, 1983). Fig. 10-5 contains SEM pictures showing the open EE
and EF structure at shallow depths and the FF orientation of small domains at a
depth of 90 m.
Fissility (p. 6, Fig. 7-25) is usually the product of preferred orientation.
It is not entirely clear why all physilites do not contain parallel oriented physil
flakes. Bioturbation and early cementation are two factors that inhibit orientation.
The presence of silt-and sand-sized grains also limits the amount of orientation that
can be achieved during compaction. Orientation tends to increase with increasing
organic content and decrease with content of carbonate minerals (Odom, 1967).
Organic matter is believed by some to act as a dispersing agent, allowing physil
plates to settle individually F to F (Moon and Hurst, 1984). In his study of
Pennsylvanian sediments, Odom (1967) found that orientation of physil particles
ranged from parallel to nearly random within a few vertical inches.

Sedimentary structures

Compared to sandstones and limestones, there has been relatively little study of
the sedimentary structures in physilites. The literature up to 1980 has been reviewed
by Potter el al. (1980). Table 10-1 from Potter et al. (1980), contains a list of
sedimentary structures in physilites and a brief statement as to their origin. The
largest variety of structures are primarily formed during the deposition of the
physilites and related rocks. Compaction and deformation “soft-sediment’’ struc-
tures are produced during or immediately after deposition. Diagenetic structures are
formed during burial, usually while the sediment has appreciable permeability.
In addition to structures produced by currents, desiccation, compaction and ion
and fluid migration, a variety of structures are produced by the many organisms
that live in and on muds (for a review, see Potter et al., 1980). In addition to fecal
pellets, biogenic structures have been grouped into three categories. Distinct mor-
phological structures (trails and burrows) are known as trace fossils. Many of these
patterns or structures have been given generic and specific names, even though the
organisms that made them may be unknown. When the trace fossils are con-
centrated and individual traces indistinct, they are referred to as burrow-mottled
textures. Nonspecific mixing of the sediment is referred to as bioturbation. The
paleontologists have developed a complex terminology for trace fossils. Classifica-
tions are based on such things as whether the animal was crawling, grazing, feeding,
dwelling, resting or escaping, To some extent, trace fossils can provide information
on environments and the clay petrologist should give them more consideration.
Many sedimentary structures can be observed directly but more detail can
713

usually be seen in radiographs (technique of x-radiography). Excellent radiographs


of shales have been published (Nuhfer et al., 1979; Cluff, 1980; Cluff et a/., 1981).
The structure or texture observed in radiographs is due to differences in density.
Thus, some of the better pictures are of black shales where pyrite, which is x-ray
dense, contrasts with physils and quartz (Fig. 10-6). Pyrite is commonly present in
burrows and associated with organic matter in organic-rich laminae.
Based on studies of Devonian dark mudrocks from the Appalachians, Nuhfer et
al. (1979) classified mudrocks into four basic types: thinly laminated ( < 2 mm),
lenticular, laminated, homogeneous (includes bioturbated) and sharply bonded
(> 1 cm). To the four basic types (Nuhfer et al., 1979) should be added, at least,
several bioturbated types ranging from isolated burrows to a high reworked mottled
structure, slump structures, and types in which shale or clay clasts, grains, or pellets
are present.

Table 10-1
Sedimentary Structures in Shale and Their Origin (After Potter ei al., 1980)

Part A. Primary Structures

Stratification
Parallel horizontal
Episodic suspension in still water
Massive
Continuous, rapid sedimentation from suspension or bioturbation
Parallel discontinuous
Episodic suspension with some bottom currents
Len ticular-wavey
Episodic traction transport with possibly some deposition from suspension
Varves
Suspension grading with rapid sedimentation in spring slow sedimentation in winter
Ripple marks and flaser bedding
Traction transport of silt, sand, and mud aggregates as ripples with some deposition from suspension
Cross bedding
Traction transport of silt and sand in the “flat-bed mode
Sole marks
Bottom scour followed by deposition
Graded beds and Bouma cycles
Deposition by turbidity currents
Massive sand beds
Deposition by grain flow
Convolute lamination, dish structures, and fluid escape pipes
Formed by fluidized sediment flow
Pebbly mudstone and conglomerate beds
Deposition by debris flow
Clay clasts
Local erosion and deposition of cohesive clay layers
Raindrop imprints
Subbaerial impact by rain drops
714

Table 10-1 (continued)

Part B. Compactional and Deformational Structures

Mud Cracks and slicensides


Desiccation and shrinkage either by subaqueous syneresis or subaerial drying
Load casts
Ball and pillows
Soft sediment displacement of sands and silts into underlying mud
Flame structure
Soft sediment displacement of sands and silts into underlying ds and sliding downslope
Mud lumps and diapirs
Large-scale upward displacement of plastic mud and shale

Part C. Diagenetic Structures

Concretions

i
Nodules
Sept aria local cementation, commonly early, without major displacement of
Geodes mud matrix; commonly form around organic nucleus
Spherulites
Cone-in-cone
Crystal growth
Crystal casts
Crystal growth, commonly salt
Color banding
Probably diffusion and generally obscure, but may be related to weathering

Cole and Picard (1975) describe eleven classes of primary structures and six
classes of secondary sedimentary structures in the Eocene fine-grained Green River
Formation (lacustrine) of Utah and Colorado (Fig. 10-7). Though many of these
rocks are referred to as oil shales, they are in fact argillaceous carbonate rocks rich
in organic matter and authigenic silicates. Nevertheless, they are illustrative of the
range of structures that could possibly be encountered in physilites and calcareous
physili tes.
Thin section and SEM studies of physilites can provide information on the origin
of the physils and the environment in which they were deposited. Post-depositional
distortions and crystal growth can best be observed by combining these two
techniques. Many examples of the use of SEM studies have been presented elsewhere
in this book, so I will not dwell on that technique.

Composition

Quartz, calcite, dolomite, feldspar, pyrite, and organic material are usually the
most abundant non-physils in muds and physilites. Quartz is the only mineral that
has been studied in any detail and its abundance related to depositional environ-
715

Fig. 10-6. X-ray radiographs of brownish-black New Albany shales (Devonian) (A) characterized by
medium to coarse laminations. Few synaeresis joints, sporadic burrowing, and few primary sedimentary
structures such as ripples and cross-bedding are present. (B) also characterized by medium to coarse
laminations and weak bioturbation, but with much more extensive snyaeresis development. (C) char-
acterized by very fine and closely spaced pyritic laminae and an occasional large pyrite module (as at
lower left). (D) characterized by slightly thicker and more irregular laminae and by numerous small
pyrite modules. Short vertical lineation at the center of the radiograph is a late stage (post-compaction)
synaeresis joint filled with calcite. Bar = 1 cm. From Harvey et al., 1977.Courtesy R.D. Harvey.

ments. Fig. 10-8 shows how the bulk composition (x-ray) of Holocene muds change
along a traverse from the mouth of the Mississippi River to the deep Gulf of Mexico
(Shaw and Weaver, 1965). The quartz content decreases from 52% to 17%.
Blatt and Totten (1981) used the sodium bisulfate fusion technique to destroy all
minerals except quartz and feldspar in mudock samples from the Permian Blaine
Formation of western Oklahoma. They found a systematic decrease in the amount
of quartz from 57% at the sand-mud line to 11%270 km from shore (Fig. 10-9). The
716

STRATIFICATION T Y P E S SECONDARY STRUCTURES

LI
PARALLEL

DISCONTINUOUS
PARALLEL ,
CURVED NONPARALLEL

STRUCTURELESS]
LOOP

BRECCIATED

WAVY PARALLEL ALGAL

n
GRADED DISRUPTION

Fig. 10-7. Diagrammatic illustration of major primary stratification types and secondary structures in the
Eocene Green River shales of Wyoming (Cole and Picard, 1975). From Potter et al., 11980. Copyright
1980 Springer-Verlag.

&bb,bl
FELDSPAR

CARBONATES

RIVER
I O X L
OUARTZ

LAGOON (BAY)
I-

DELTA (SHELF) NEAR SHORE


(SLOPE 1
-

DEEP

IOU NUMBER INDICATES SAMPLES ANALYZED

Fig. 10-8. Plots showing the average mineral composition of a series of recent mud samples from the
vicinity of the Mississippi River delta. Numbers indicate the number of samples analyzed. From Shaw
and Weaver, 1965. Copyright 1965 Soc. Econ. Paleo. Miner.
717

Fig. 10-9. Aerial variation in the percentage of quartz in Blaine shales in relation to distance from a
shoreline established using palynologic and lithofacies data. A single sample point on the map may
represent several samples located too close to each other to be separable at the scale of the map. From
Blatt and Totten, 1981. Copyright 1981 SOC.Econ. Paleo. Miner.

mean grain size decreased seaward from approximately 37 pm to 7.5 pm. In a


similar study of the Pierre Shale in midcontinental North America, Jones and Blatt
(1984) found a similar decrease in the percent and grain size of quartz from the
shoreline (west) seaward (east) (Fig. 10-10). The quartz in the Pierre is extremely
fine with the primary mode in the range of 1.4 to 2.7 pm. The fine size suggests
much of the quartz, along with volcanic ash, was wind transported.
Fig. 10-11 contains SEM pictures of quartz concentrates from the Cambrian
Conasauga shale and slate. The shale quartz grains tend to be subequant and are
rounded as is typical for detrital quartz. The slate quartz is similar in size but is
more platy and angular; the surfaces much smoother than -the shale quartz. The
slate quartz has recrystallized during low grade metamorphism.
I was able to establish (1962) in the Cretaceous of the Washakie Basin, Wyoming,
a close correlation between electric log resistivity values of shales and their quartz
content. Quartz and resistivity values were highest in flood-plain deposits (10 ohms)
and decreased progressively from prodeltaic shales (8 ohms), through shallow marine
(6 ohms) to deep marine (4 ohms). Of all the factors studied, the quartz content and
resistivity values of shales were the best environmental indicators. The relation
between resistivity and quartz content works only if the carbonate content is low
and there is no drastic change in the physil suite.
c
P
U

L
r
50h
V
+

. * .

'.
:
a'
10

I
0
I
100
I
200 300
6
400 500 600 700 800
Approximate distance to the nearest shoreline (km)

?-.

r
9.0

fe
LL
+
8.0

0 7.01
0
1. .*-
a
a

5.0I
loo 200 300 400 500 600 700 aoc
Approximate distance to the nearest shoreline (krn)

Fig. 10-10. Relation between quartz concentration in the quartz+ feldspar + clay fraction (upper), mean
grain size of quartz+feldspar (lower) and distance to nearest shoreline. Blatt and Totten (1981) trends
refer to data from Blaine Shale (Fig. 9-9). From Jones and Blatt, 1984. Copyri3ht 1984 SOC.Econ. Paleo.
Miner.

In carbonate sections, shale resistivity closely reflects the carbonate content of


the shale.
Shaw and Weaver (1965) observed a relatively good relation between the
quartz/feldspar ratio in shales and that in associated sandstones. Shales associated
with mature quartzite had a Q/F ratio around 30, those associated with arkoses had
a ratio of 1.2, and shales associated with various types of graywackes have average
ratios ranging from 7 to 12.
719

Fig. 10-11. Rounded quartz from Conasauga shale (A) and angular, recrystallized quartz from Conasauga
slate (b). White bar = 2 pm.
720

Color

The color of shales has been reviewed by Pettijohn (1975) and Potter et a/. (1980).
though shales occur in a wide variety of colors, the significance of shale color has
not received the attention that the color of sandstones has. To a large extent, the
color of shales is controlled by organic mater and Fe. The inherent color of the
various physils is a factor only when the amount of organic matter and Fe oxides is
small.
Kaolinite and montmorillonite have a white to light neutral color. Illite, chlorite
and glauconite commonly have a green color. The green color is apparently due to
the Fe2+ in the octahedral sheet, even though Fe3+is usually more abundant. The
greenish color of many Paleozoic marine shales probably reflects the color of the
physils (illite and chlorite).
Fe oxides in physilites are normally fine-grained and form a thin film on the
minerals, only a few percent can produce a relatively deep color. In oxidizing
environments where the Fe3+/Fe2+ratio is high and the organic content low, a red
color results; usually the pigment is hematite. Thus, continental shales and mud-
stones are commonly red, as are some deep sea muds. There is a large amount of
literature on the origin of red beds (Van Houten, 1973; Turner, 1980). The present
consensus is that the oxidation state of Fe is so easily changed that the color of red
beds is acquired at the most recent site where oxygenated fluids were present, at the

5 0

30

I 5
z
P 10
K
a
V

0
$ 0.5
K
0
03
w
V
Q
WO2
n OLIVE GRAY (5Y 4/11

01
GREENISH GRAY (5GY 611)

0 02 04
MOLE FRACTION Fo" - 06
mFc'*
mFa,, mFo,,,
+
08 10

Fig. 10-12. Shale colors as related to content of organic carbon and oxidation state of iron. m represents
the number of moles of iron per gram of rock. From Potter et al.. 1980. Copyright 1980 Springer-Verlag.
721

side of deposition, during burial, or by weathering of the outcrop. The Fe is


normally obtained by leaching of various Fe silicates and ilmenite. The Fe content
of red shales is not particularly higher than that of gray and black shales but the
Fe,O,/FeO ratio is higher ( - 3-5)(Tomlinson, 1916).
The black and gray colors are normally due to the presence of organic matter
(measured as organic carbon). Unless the production or organic matter is exces-
sively high or the rate of deposition rapid, the preservation of appreciable organic
matter (3-10%) is restricted to anoxic environments such as restricted marine
basins, either deep or shallow, marshes, swamps and some lagoonal and estuarine
deposits. These reducing environments also decrease the Fe3+/Fe2+ ratio causing a
progressive shift in “Fe color”: red --+ yellow + green + . Commonly, much of the
Fe is mobilized, concentrated and precipitated as metastable Fe monosulfides
(black) which during early diagenesis is readily transformed to pyrite (Berner, 1984).
To a large extent, the formation of pyrite eliminates the Fe as a color source and the
degree of blackness is largely related to the amount and size of the organic matter.
Fig. 10-12 is a graph devised by Potter et al. (1980) to relate shale color to the
percent organic carbon and the Fe2+/Fe2++ Fe3+ ratio. The relation shown in the
graph accounts for the color of many physilites. Some exceptions are green colors
due to the color of the physilites and purple and black colors due to Mn oxides.
X-ray slides are useful for detecting subtle color differences in clays and shales.
The color changes can commonly be correlated with changes in the physil suite
and/or depositional environments.

Thin sections

In recent years, the microscopic study of clays and shales has not received the
attention it deserves. Unfortunately, I too will not be able to give the subject the
attention it deserves. Though SEM investigations seem to have surplanted optical
microscopic work, it is best that both techniques be used. In many situations, it is
possible to examine the same thin section with the optical microscope and the SEM.
Carozzi (1960) has provided some good microscopic descriptions of residual and
reworked clays such as kaolinites and bentonites and of a variety of shales, e.g.,
silicious shales, calcareous shales and carboniferous shales. Potter et al. (1980)
discussed the features to look for when examining shales and claystones, e.g.,
detrital and authigenic minerals, cements, pellets, organic matter, grain size, bur-
rows. Unfortunately, little detail can be seen in black and white photomicrographs
of clays and shales and they generally are not worth publishing. Potter et al. (1980)
published 22 colored photographs of shales which show some of the basic sedinien-
tary structures. One has to get the “feel” of a thin section by studying it for many
hours at different magnifications. The impressions are difficult to summarize in a
photomicrograph. Because of the small size of physils, it is usually best to cut thin
sections of shales and slates thinner (10-20 pm) than normal (30 pm).
This Page Intentionally Left Blank
723

REFERENCES

Aagaard, P. and Helgeson, H.C., 1983. Activity/composition relations among silicates and aqueous
solutions: 11. Chemical and thermodynamic consequences of ideal mixing of atoms of homological
sites in montmorillonites, illites. and mixed-layer clays. Clays Clay Miner., 31 : 207-217.
Abbot, P.L., Minch, J.A. and Peterson, G.L., 1976. Pre-Eocene palesol south of Tijuana, Baja California,
Mexico. Jour. Sed. Petrol, 46: 355-361.
Ager. D.V., 1980. The Geology of Europe. John Wiley and Sons, N.Y., 535p.
Ahn, J.H. and Peacor, D.R., 1985a. Transmission electron microscopic study of diagenetic chlorite in
Gulf Coast argillaceous sediments. Clays Clay Miner., 33: 228-236.
Ahn. J.H. and Peacor, D.R., 1985b. TEM study of mineralogy and diagenesis of phyllosilicates in
volcanogenic sediments from the Southland Syncline. New Zealand. Geol. SOC. Amer. Prog. with
Abs., Orlando, 510.
Ahn, J.H. and Peacor., R., 1986a. Transmission and analytical electron microscopy of the smectite-to-il-
lite transition. Clays Clay Miner., 34: 165-179.
Ahn, J.H. and Peacor, D.R., 1986b. Transmission electron microscope data for the origin and structure of
“fundamental particles.” Clays Clay Miner., 34: 180-186.
Al-Bakir, Khalaf, F., and Al-Ghadban, A., 1984. Mineralogy, genesis and source of surficial sediments in
the Kuwait marine environments, northern Arabian Gulf. Jour. Sed. Petrol., 54: 1266-1279.
Alberti, A. and Brigatti, M.F., 1985. Crystal chemical differences in Al-rich smectites as shown by
multivariate analysis of variance and discriminant analysis. Clays Clay Miner., 33: 546-558.
Albrecht, P. and Ourisson, G., 1969. Diagtntse des hydrocarbures saturts dans une sCrie sedimentaire
tpaisse (Douala. Cameroun). Geochim. et Cosmochim Acta., 33: 138-142.
Alexander, L.T., Faust, G.T., Hendricks, S.B., Insley. H. and McMurdie, H.F., 1943. Relationship of the
clay minerals halloysite and endellite, Amer. Miner. 28: 1-18.
Alexiades, C.A. and Jackson, M.L., 1965. Quantitative determination of vermiculite in soils. Soil Sci. SOC.
Amer. Proc., 29: 522-527.
Alietti. A,, 1958. Some interstratified clay minerals of the Taro Valley. Clay Miner. Bull., 3: 207-21 1.
Alietti, A., 1960. Sudi una montronite poco ferrifera di Chiampo (Lessini). Soc. Tipografica Editrice
Modenese, 33-44.
Alietti, A. and Meisner. J.. 1980. Structure of a talc/saponite mixed-layer mineral. Clays Clay Miner., 28:
388-390.
Allen, B.L. and Pashai. A,, 1981. Selected mineralogical characteristics of calcretes in West Texas and
eastern New Mexico. Inter. Conf. Aridic Soils, Jerusalem, Abs., 5.
Alling, H.L., 1945. Use of microlithologies as illustrated by some New York sedimentary rocks. Geol.
SOC.Amer. Bull., 56: 737-755.
Almon, W.R., 1981. Depositional environments and diagenesis of Permian Rotliegendes sandstones in
the Dutch section of the southern Northern Sea. In: F.J. Longstaffe (Editor), Clays and The Resource
Geologist. Min. Assoc. Canada. Calgary, 119-147.
Almon. W.R.. and Davies, D.K.. 1979. Regional diagenetic trends in the Lower Cretaceous Muddy
Sandstone, Powder River Basin. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of Diagenesis.
SEPM Sp. Pub. No. 26, Tulsa, 379-340.
Almon, W.R.. Fullerton, L.B. and Davies, D.K., 1976. Pore space reduction in Cretaceous sandstones
through chemical precipitation of clay minerals. Jour. Sed. Petrol., 46: 89-96.
724

Alt, J.C., Laverne, C. and Muehlenbachs. K.. 1983. Alteration of the upper oceanic crust: mineralogy and
processes in Deep Sea Drilling Project Hole 504B. Leg 83. In: R.N. Anderson, J. Honnorez, K.
Becker et al. (Editors), Init. Repts. DSOP, 83: Wash. ( U S . Govt. Printing Office), 217-247.
Altaner. S.P.. Hower, J., Whitney, G. and Aronson, J.L., 1984. Model for K-bentonite formation:
evidence from zoned K-bentonites in the disturbed belt, Montana. Geology, 12: 412-415.
Altschuler, Z.S.. Dwornik, E.J., and Kramer, H., 1963. Transformation of montmorillonite to kaolinite
during weathering. Science, 141: 148-152.
Amador, E.S.. and Zalan, P.V., 1980. Paleoclimatic interpretation of Cenozoic continental sediments
from Resende Basin, Brazil. An. Acad. Brazil Citnc., 52: 591-602.
Anand, R.R., Gilkes, R.J., Armitage, T.M., and Hillyer, J.W., 1985. Feldspar weathering in lateritic
sepiolite. Clays Clay Miner., 33: 31-43.
Ames. L.L. and Goldich, S.S., 1958. A contribution on the Hector, California, bentonite deposit. Econ.
Geol., 53: 22.
Amiri-Garroussi, K., 1977. Origin of montmorillonite in the early Jurassic shales of N W Scotland. Geol.
Mag.. 114: 281-290.
Amouric, M. and Parron, C., 1985. Structure and growth mechanism of glauconite as seen by high
resolution transmission electron microscopy. Clays Clay Miner., 33: 473-482.
Andrews, J.E., Packman. G. et a/., 1975. Site Report 285, 286. Initial Reports of the Deep Sea Drilling
Project, 30, h a s h . (U.S. Gov. Printing Office), 27-131.
Andrews, P.B. and Ovenshine, A.T., 1974. Terrigenous silt and clay facies: deposits of the early phase of
ocean basin evolution, DSDP Leg 29. In: J.P. Kenneth, R.E. Houtz. ei a/., (Editors) Initial Reports of
the Deep Sea Drilling Project. Vol. 29, Wash. (U.S. Gov. Printing Office), pp. 1049-1063.
Angel, B.R., and Hall, P.L., 1972. Electron spin resonance studies of kaolins. Proc. Inter. Clay Cnnf.,
Madrid. pp. 47-60.
Angel. B.R. and Hall, P.L., 1973. Electron spin resonance studies of kaolin. Proc. Int. Clay Conf. Madrid,
47-60.
Antia, D.D.J.. 1986. Kinetic method for modeling vitrinite reflectance. Geology, 14: 606-608.
Antiweiler, R.C., and Drever, J.I., 1983. The weathering of a late Tertiary volcanic ash: importance of
organic solutes. Geochim. Cosmochim Acta, 47: 623-629.
Aoki. S.. Kohyanna, N., and Sudo, T., 1974. An iron-rich montmorillonite in a sediment cnre from the
northeastern Pacific. Deep-sea. Res., 21 : 865-875.
Aoki, S. and Oinuma, K., 1974. Clay mineral compositions in recent marine sediments around
Nansei-Syoto Islands, south of Kyusyu, Japan. Jour. Geol. Soc. Japan, 80: 57-63.
Aomine. S. and Jackson, M.L., 1959. Allophane determination in Ando-Soils by cation-exchange capacity
delta values. Proc. Soil Sci. Soc. Amer., 23: 210-214.
Aomine, S. and Miyauchi, N., 1963. Age of the youngest hydrated halloysite in Kyushu. Nature, 199:
1311-1312.
April, R.H., 1980. Regularly interstratified chlorite/vermiculite in contact metamorphosed red beds. in
Newark Group, Connecticut Valley. Clays Clay Miner., 28: 1-11.
April. R.H., 1981. Trioctahedral smectite and interstratified chlorite/smectite in Jurassic strata of the
Connecticut Valley. Clays Clay Miner., 29: 31-39.
April, R.H., Hluchy, M.M., and Newton, R.M., 1986. The nature of vermiculite in Adirondack soils and
till. Clays Clay Miner., 34: 549-557.
Arnaud, R.J., Sr., and Nortland, M.M., 1963. Characteristics of the clay function in a chernmemic
topodzolic sequence of soil profiles in Saskatchewan. Can. Jour. Soil. Sci., 43: 336-349.
Arnone. R., 1972. Floccule characterization of the Satilla Estuary. M.S. Thesis Ga. Inst. Tech., Atlanta,
GA, 56 p.
Aronson, J.L. and Hower, J., 1976. Mechanism of burial metamorphism of argillaceous sediment: 2.
radiogenic argon evidence. Geol. Soc. Amer. Bull., 87: 738-744.
Arrese. F.. Neira, E.. and Rodriguez, J., 1966. Clay mineral composition of Navarra mark (Spain). Proc.
Internat. Clay conf., Jerusalem, 1966, 101-102.
Arrhenius, G., 1963. Pelagic Sediments. In: M.N. Hull (Editor) The Sea, 3. lnterscience Pub., 655-727.
Arthur, M.A. and Natland. J.H.. 1979. Carboniferous sediments in the North and South Atlantic: The
role of salinity in stable stratification of Early Cretaceous Basins. In: M. Talwani. W.Hay and W.B.F.
725

Ryan (Editors), Deep Drilling Results in the Atlantic Ocean: Continental Margins and Paleoenviron-
ment. Amer. Geoph. Union, Wash. D.C., 375-401.
Artru, Ph. and Gauthier, J., 1968. Evolution gtometrique et diagknttique d’un skrie miogeosynclinale
(Lias inftrieur A Berriasien) d’aprks du sondage de Valvigntres (France sud-est). Bull. Centre Rech.
Pau-SNPA, 2: 101-116.
Ataman, G. and Baysal. O., 1978. Clay mineralogy of Turkish borate deposits. Chem. Geol. 22: 233-247.
Atterberg, A, 1905. Die rationelle klassifikation der sande und kiese. Chem. Zeitschr., 29: 195-198.
Axelrod, D.I., 1981. Role of volcanism in climate and evolution. Geol. SOC.Amer., Sp. Paper 185, 59 pp.
Baadsgaard, H.S.S. and Dodson, M.H., 1964. Potassium-argon ages of sedimentary and pyroclastic rocks.
Quart. Jour. Geol. SOC.Lond., 1205: 119-127.
Badaut, D. and Risacher, F., 1983. Authigenic smectite on diatom frustules in Bolivian saline lakes.
Geochim. Cosmochim. Acta 47: 363-375.
Bailey, S.W., 1963. Polymorphism of the kaolin minerals. Am. Min., 1196-1206.
Bailey, S.W.. 1975. Chlorites. In: J. E. Gieseking (Editor), Soil Components, V.2, Inorganic Components.
Springer-Verlag, New York, 191-264.
Bailey, S.W., 1980a. Summary of recommendations of AIPEA nomenclature committee on clay minerals.
Amer. Miner., 65: 1-7.
Bailey, S.W., 1980b. Structures of layer silicates. In: G.W. Brindley and G. Brown (Editors), Crystal
Structure of Clay Minerals and Their X-ray Identification. Miner. SOC.London, Mono. 5: 1-124.
Bailey, S.W., 1986. Report of AIPEA Nomenclature Committee. Supplement to AIPEA Newsletter,
February, 1986.
Bailey, S.W., Brindley, G.W., Kodama, H., and Martin, R.T., 1979. Report of the Clay Minerals Society
Nomenclature Committee for 1977 and 1978. Clays Clay Miner., 27: 238-239.
Bailey, S.W. and Brown, B.E., 1962. Chlorite polytypism: I Regular and semi-random one-layer
structures. Amer. Miner., 47: 819-850.
Bailey, S.W. and Brown, B.E., 1963. Chlorite polytypism. 11. Crystal structure of a one-layer Cr-chlorite.
Amer. Miner., 48: 42-61.
Bailey, S.W. and Tyler, S.A., 1960. Clay minerals associated with the Lake Superior iron ores. Econ.
Geol., 55: 150-175.
Bain. D.C., 1977. The weathering of ferruginous chlorite in podzol from Argyllshire, Scotland. Geo-
derma, 17: 193-208.
Bain. D.C. and Russell, J.D., 1980. Swelling minerals in a basalt and its weathering products from
Morvern, Scotland: 1. Interstratified montmorillonite-vermiculite-illite.Clay Miner., 15: 445-451.
Bain, D.C. and Tait, J.M., 1977. Mineralogy and origin of dust fall on Skye. Clay MIner., 12: 353-355.
Baker, C.E., 1983. Influence of time on metamorphism of sedimentary organic matter in liquid-dominated
geothermal systems, western North America. Geology, 11: 384-387.
Baker, E.T., 1973. Distribution and composition of suspended sediment in the bottom waters of the
Washington continental shelf and slope. Jour. Sed. Petrol., 43: 812-821.
Ball, D.F., 1966. Chlorite clay minerals in Ordovician pumice-tuff and derived soils in Snowdonia, North
Wales. Clay Minerals, 6: 195-208.
Ball, D.F., 1968. Interstratified illitic clay in Ordovician ash from Conway, North Wales. Clay Minerals,
7: 363-366.
Baltrusaitis, E.J., 1974. Middle Devonian bentonite in Michigan Basin. Amer. Assoc. Petrol. Geol. Bull.,
58: 1323-330.
Bambach, R.K., Scotese, C.R., and Ziegler, A.M., 1980. Before Pangea: The geographies of the Paleozoic
world. Amer. Sci., 26-38.
Bannister, F.A., 1943. Brammalite (sodium-illite), a new mineral from Llandebie, south Wales. Miner.
Mag., 26: 304-307.
Bannister, F.A. and Whittards, W.F., 1945. A magnesium chamosite from the Wenlock limestone of
Wickwar, Gloucestershire. Miner. Mag., 27: 101-110.
Barbieri, M., Ballanca, A. and Neri, R., 1981. Origin of zeolites associated with montmorillonite and
silica phases in Miocene deposits of Sicily. Miner. Petrogr. Acta, 25: 41-55.
Barker, P.F., Dalziel, I.W.D. el a/., 1976. Initial Reports of the DSDP, 36, Wash. (U.S. Gov. Printing
Office), 1080 pp.
726

Barnhisel, R.I.. 1977. Chlorites and hydroxy interlayered vermiculite and smectite. In: J.B. Dixon and
S.B. Weed (Editors), Minerals in Soil Environments. Soil Sci. SOC. Amer.. Madison. Wisc.. pp.
331-356.
Barnhisel, R.I. and Rich, C.E., 1963. Gibbsite formation from aluminum-interlayers in montmorillonite.
Soil Sci. SOC.Amer. Proc., 27: 632-635.
Barrett, P.J., 1975. Textural characteristics of Cenozoic preglacial and glacial sediments at site 270, Ross
Sea, Antarctica. In: D.E. Hayes, A. Frakes el a / . . Initial Reports of the Deep Sea Drilling Project.
Vol. 28, Wash. (US.Govt. Printing Office), pp. 757-73.
Barshad, I., 1948. Vermiculite and its relation to biotite as revealed by base exchange reactions, X-ray
analysis, differential thermal curves and water content. Amer. Miner., 33: 655-678.
Barshad, I., 1957. Factors affecting clay formation. Proc. 6th Natl. Conf. Clays Clay Miner., pp. 110-143.
Barshad, I., 1966. The effect of a variation in precipitation o n the nature of clay mineral formation in
soils from acid and basic igneous rocks. Proc. Int. Clay Conf., Jerusalem, I : 167-173.
Barshad, 1. and Kishk, F.M., 1969. Chemical composition of soil vermiculite clays as related to their
genesis. Contr. Miner. Petrol., 24: 136-55.
Bartholomk. P., Lombard, A.L. and Moulin, C., 1963. Contribution a I’ktude skdimentologique des
argilites Mksozoiques d e I’ouest du Congo. Rkpublique du Congo. Bull. N o 10. 31 pp.
Barzanji, A.F.. Sys, C. and EswBran, H.. 1975. The clay mineralogy of gypsiferous soils of Iraq. Proc.
Inter. Clay Cmf., Mexico City, 657.
Bass, M.N.. 1976. Secondary minerals in oceanic basalt, with specially reference to Leg 34. Deep Sea
Drilling Project. In: R.S. Yeats, S.R. Hart et al., 1976. Initial Report of the Deep Sea Drilling Project.
34, Wash. ( U S . Govt. Printing Office), pp. 393-432.
Bassett. W.A., 1960. Role of hydroxyl orientation in mica alteration. Bull. Geol. SOC.Amer, 71: 449-456.
Bassett, W.A.. 1963. The geology of vermiculite occurrences. Clays Clay Miner., 10: 61-69.
Bassett, R.L. and Palmer, D.P., 1981. Clay mineralogy of the Palo Duro Basin evaporite sequences.
Randall County Core. Texas Bur. Econ. Geol.. Geol. Circ. 81-3: 108-118.
Bates, T.F.. 1959. Morphology and crystal chemistry of 1 : 1 layer lattice silicates. Amer. Miner.. 44:
78-114.
Bates, T.F.. 1962. Halloysite and gibbsite formation in Hawaii. Clays Clay Miner. 9: 307-314.
Bausch, W.M., 1977. Clay mineral provinces in the Upper Jurassic of Europe. Proc. Third European Clay
Conf., Oslo. 8-9.
Bayliss. P. and Levinson, A.A., 1970. Clay mineralogy and boron determinations of the shales from the
Reindeer well, Mackenzie River Delta, N.W.T., Canada. Bull. Canadian Petrol. Geol.. 18: 80-83.
Bayliss P., Loughnan, F.C. and Standard, J.C., 1965. Dickite in the Hawkesbury Sandstone of the Sydney
Basin, Australia. Amer. Miner., 50: 418-426.
Beall, A.O., Jr., 1964. Stratigraphy of the Taylor formation (Upper Cretaceous). East-Central Texas.
Baylor Geol. Studies, Bull. No. 6, 35 p.
Beaver, P.J. and Dumbleton, M.J., 1966. Clay minerals and geomorphology in four Caribbean Islands.
Clay Minerals, 6: 371-382.
Bell, T.E.. 1986. Microstructure in mixed-layer illite/smectite and its relationship to the reaction of
smectite to illite. Clays Clay Miner., 34: 146-154.
Belt Symposium, 1973. Vol. 1, Dept. Geol.. U. Idaho and Idaho Bureau of Mines and Geol., Moscow.
Idaho, 322 p.
Bennett, R.H., Bryant, W.R. and Keller, G.H.. 1977. Clay fabric and geotechnical properties of selected
submarine sediment cores from the Mississippi Delta, U.S. Dept. Commerce, NOAA Prof. Paper 9,
86 p.
Bennett. R.H., Bryant, W.R. and Keller. G.H.. 1981. Clay fabric of selected submarine sediments:
fundamental properties and models. Jour. Sed. Petrol., 51 : 217-232.
Benson, L.V. and Teague. L.S., 1982. Diagenesis of basalts from the Pasco-Basin. Washington - I.
distribution and composition of secondary mineral phases. Jour. Sed. Petrol.. 52: 595-613.
Bentor. Y.K., Bodenheimer, W. and Heller, L., 1963. A reconnaissance survey of the relationship between
clay mineralogy and geological environment in the Negev (Southern Israel). Jour. Sed. Petrol.. 33:
874-903.
Bentor, Y.K., and Kastner, M.. 1965. Notes on the mineralogy and origin of glauconite. Jour. Sed. Petro..
35: 155-166.
727

Bkrczi, I., 1972. Sedimentological investigation of Pre-Pannonian sedimentary formations in the Szeged
Basin, SE-Hungary. Acta Geol., Acad. Scieniarium Hungaricae, 16: 229-250.
Berger, W.H. and von Rad, U., 1972. Cretaceous and Cenozoic sediments from the Atlantic Ocean. In:
Hayes, D.E., Primm, A.C., et al, 1972, Initial Reports of the DSDP, 14, Wash. ( U S . Govt. Printing
Office), 787-954.
Berggren, W.A. and Aubry, M.P., 1984. Rb-Sr glauconite isochron of the Eocene Castle Hayne
Limestone, North Carolina: further discussion. Geol. SOC.Amer. Bull., 95: 364-370.
Berggren, W.A. and Hollister, C.D., 1974. Paleogeography, Paleobiogeography and the history of
circulation in the Atlantic Ocean. In: W.W. Hay (Editor), Studies in Paleo-Oceanography. SOC.Econ.
Paleo. Miner., Spec. Pub. 20, 126-186.
Berg-Madsen, V., 1983. High-alumina glaucony from the Middle Ordovician of Oland and Bornholm,
southern Baltoscandia. Jour. Sed. Petrol, 53: 875-893.
Berner, R.A., 1971. Principles of chemical sedimentology. McGraw-Hill Book Co., New York, 240 p.
Berner, R.A., 1984. Sedimentary pyrite formation: an update. Geochim. Cosmochim. Acta, 48: 605-615.
Berry, R.W. and Johns, W.D., 1966. Mineralogy of the clay-sized fractions of some North Atlantic-Arctic
Ocean bottom sediments. Geol. Soc. Amer. Bull., 77: 183-196.
Beskin, E.A., 1984. Compositional variations of authigenic chlorites in the Tuscaloosa Formation, Upper
Cretaceous of the Gulf Coast Basin. Clay Miner. Soc. Prog. with Abs., Baton Rouge, 25.
Bethke, C.M. and Altaner, S.P., 1986. Layer-by-layer mechanism of smectite illitization and application
to a new rate law. Clays Clay Miner., 34: 136-145.
Bhattacharyya, D.P., 1983. Origin of berthierine in ironstone. Clays Clay Miner., 31: 173-182.
Biggs, R.B., 1967. The sediments of Chesapeake Bay. In: G.H. Lauff (Editor),s Estuaries. Amer. Assoc.
Adv. Sci., Pub. No. 83: 239-260.
Biggs, R.B. and Howell, B.A., 1984. The estuary as a sediment trap: alternate approaches to estimating its
filtering efficiency. In: V.S. Kennedy (Editor), The Estuary as a Filter, Academic Press, pp. 107-129.
Bigham, G.N., 1972. Clay mineral transport on the inner continental shelf of Georgia. Ga. Inst. Tech.,
Atlanta, M.S. Thesis, 49 p.
Bigham, G.N., 1973. Zone of influence - inner continental shelf of Georgia. Jour. Sed. Petrol., 43:
207-214.
Bigham, J.M., Jaynes, W.F. and Allen, B.L, 1980. Pedogenic degradation of sepiolite and palygorskite on
the Texas High Plains. Soil. Sci. Soc. Amer. Jour., 44: 159-167.
Birch, G.F., 1973. Mineralogy and geochemistry of the fine sediment ( 4 63 p ) on the shelf off the
southwestern coast and in the adjacent source area. Dept. Geol., U. Cape Town, Tech. Rpt. 6.47-51.
Birch, G.F., Willis, J.P. and Rickard, R.S. 1976. An electron microprobe study of glauconite from the
continental margin off the west coast of South Africa. Marine Geol., 22: 271-283.
Birkeland, P.W., 1974. Pedology, Weathering, and Geomorphological Research. Oxford University Press,
London, 285 p.
Biscaye, P.E., 1965. Mineralogy and sedimentation of recent deep-sea clay in the Atlantic Ocean and
adjacent seas and oceans. Geol. SOC.Amer. Bull., 76: 803-831.
Biscaye, P.E. and Eittreim, S.L., 1977. Suspended particulate loads and transports in the nepheloid layer
of the abyssal Atlantic Ocean. Marine Geol., 23: 155-172.
Bischoff, J.L., 1972. A ferroan nontronite from the Red Sea geothermal system. Clays Clay Miner., 20:
217-223.
Bish, D.L. and Giese, R.F., Jr., 1981. Interlayer bonding in IIb chlorite. Amer. Miner., 66: 1216-1220.
Bismuth, H., Boltenhagen, C., Donze, P., LeFtvre, J. and Saint-Marc, P., 1982. Etude skdimentologique
et biostratigraphique du Crktack Moyen et Supkrierur du Djebel Semmama (Tunisie du centre nord).
Cretaceous Res., 3: 171-185.
Bjorlykke, K., 1971. Petrology of Ordovician sediments from Wales. Norsk geol. Tidsskr. 51: 123-139.
Bjorlykke, K., 1974. Geochemical and mineralogical influence of Ordovician island areas on epicontinen-
tal clastic sedimentation, a study of lower Paleozoic sedimentation in the Oslo Region, Norway.
Sedimentology, 21: 251-272.
Bjbrlykke, K., 1984. Formation of secondary porosity: how important is it? In: D.A. McDonald and R.C.
Surdam (Editors), Clastic Diagenesis, Amer. Assoc. Pet. Geol. Mem. 37, Tulsa, pp. 277-286.
Black, P.M., 1974. Oxygen isotope study of metamorphic rocks from the Oukgoa District, New
Caledonia. Contrib. Miner. Petrol., 47: 197-206.
728

Black, P.M., 1975. Minerology of New Caledonian metamorphic rocks IV sheet silicates from Outgoa
district. Contrib. Miner. Petrol., 49: 269-284.
Blake, D.F., and Peacor, D.R.. 1981. Biomineralization in crinoid echinoderms: characterization of
crinoid skeletal elements using TEM and STEM microanalysis. SEM/1981, 111, 321-328.
Blanche, J.B. and Whitaker, J.H.MeD., 1978. Diagenesis of part of the Brent Sand Formation (Middle
Jurassic) of the northern North Sea Basin. Jour. Geol. SOC.Lond., 135: 73-82.
Blancher, D.W., Jr., 1973. The petrology of the Maccrady Formation and the Greenbrier Group.
Mississippian, of the Hurricane Ride snycline of southwestern Virginia and West Virginia. Geol. SOC.
Am. Abs. with Program, 5 : 378-379.
Blatt, H., 1979. Diagenetic processes in sandstones. In: P.A. Scholle (Editor), Aspects of Diagenesis, Soc.
Econ. Paleo. Miner. Special Pub. No. 26, Tulsa, pp. 141-157.
Blatt, H. and Schultz, D.J., 1976. Size distribution of quartz in mudrock. Sedimentology, 23: 857-866.
Blatt, H. and Totten, M.W.. 1981. Detrital quartz as an indicator of distance from shore in marine
mudrocks. Jour. Sed. Petrol, 51: 1259-1266.
Blatter, C.L., Roberson, H.E., and Thompson, G.R., 1973. Regularly interstratified chlorite - dioc-
tahedral smectite in dike-intruded shales, Montana. Clays Clay Miner., 21 : 207-212.
Blokh, A.M. and Siderenko, G.A., 1960. Nefedyevite from the Transbaikal. Doklady Akad. Nauk SSSR,
135, No. 3, 701-704.
Blount, A.M., Threadgold, I.M. and Bailey, S.W., 1969. Refinement of the crystal structure of nacrite.
Clays and Clay Min., 17: 185-194.
Bode, G.W., 1973. Grain size analyses, Leg 18, In: Kulm, L.O., von Huene, R. et al.. 1973, Initial Reports
of the Deep Sea Drilling Project, Vol. 18, Wash. (US. Gov. Printing Office), 1061-1067.
Bodine, M.W., Jr., 1978. Clay-mineral assemblages from drill core of Ochoan evaporites, Eddy County.
New Mexico. New Mexico Bureau Mines Miner. Res., Circular 159. 21-31.
Bodine, M.W., Jr. and Rueger, B.F., 1983. Clay-mineral assemblages from the Gibson Dome No. 1 drill
core, Paradox Basin, Utah. Clays Clay Miner., Prog. with Abs., 49.
Bodine, M.W., Jr., and Standaert, R.R., 1977. Chlorite and illite compositions from Upper Silurian rock
salts, Retsof, New York. Clays clay Miner., 25: 57-71.
Bohor, B.F., 1983. Clay mineralogy of a Cretaceous - Tertiary boundary claystone from Montana. Clay
Miner. SOC.Prog. with Abs., 48.
Boles, J.R. and Franks, S.G., 1979. Clay diagenesis in Wilcox sandstones of southwest Texas: Implica-
tions of smectite diageneses on sandstone cementation. Jour. Sed. Petrol, 49: 55-70.
Bonatti, E., 1963. Zeolites in Pacific pelagic sediments. Trans. New York Acad. Sci., 25: 938-948.
Bonatti, E., 1965. Paragonite hyaloclastites and alteration of volcanic glass in the ocean. Bull. Volcano.,
28: 257-269.
Bonatti, E., 1967. Mechanisms of deep-sea volcanism in the South Pacific. In: P.H. Abelson (Editor).
Researches in Geochemistry, 4: 453-491, Wiley, N.Y.
Bonatti. E. and Arrhenius, G., 1965. Eolian sedimentation in the Pacific off northern Mexico. Marine
Geol., 3: 337-348.
Bonattti, E. and Joenson, O., 1968. Palygorskite from Atlantic deep sea sediments. Amer. Miner., 53:
975-983.
Bonorino, F.G., 1966. Soil clay mineralogy of the Pampa Plains, Argentina. Jour. SEd. Petrol, 36:
1026-1035.
Booth, J.S. and Osborne, R.H., 1971. The American Upper Ordovician standard. XV. Clay mineralogy of
insoluble residues from Cincinnatian limestones, Hamilton County. Ohio. Jour. Sed. Petrol., 41 :
840-843.
Borchardt, G.A., 1977. Montmorillonite and other smectite minerals. In: J.B. Dixon, and S.B. Weed
(Editors), Minerals in Soil Environments. Amer., Madison, Wis., 293-330.
Borchardt, G.A., Jackson, M.L. and Hole, F.D., 1966. Expansible layer silicate genesis in soils depicted in
mica pseudomorphs. Proc. Inter. Clay Conf., Jerusalem, 1: 175-185.
Borchert, H., and Braun, H., 1963. Zum chemismus von drei glauconittypedn. Chem. Erde, I , 23: 82-90.
Bornhold. B.D. and Giresse, P., 1985. Glauconitic sediments on the continental shelf off Vancouver
Island, British Columbia, Canada. Jour. Sed. Petrol., 55: 653-664.
Borst, R.L., 1966. A mineralogical study of some Lower Devonian (Helderberg) rocks of the central
Hudson Valley, New York. Jour. Sed. Petrol. 36: 775-793.
729

Bossi, G.E., 1972. A chloritized montmorillonite from the Rio Chiflon Formation (TR) of Argentina.
Clays Clay Miner., 20: 251-258.
Bossi, G.E. and Lenzi, S.R., 1981. Mineralogia das argilas nos grupos Guati e Itarar6 (Permiano) no
subsolo de Gravatai, Rolante e Morungava, no Rio Grande Do Sul, Brasil. Pesquisas, Porto Alegre,
14: 7-21.
Bostick, N.H., 1974. Phytoclasts as indicators of thermal metamorphism, Franciscan Assemblage and
Great Valley Sequence (Upper Mesozoic), California. Geol. SOC.America Spec. Paper 153: 1-17.
Bostick, N.H. and Danberger, H.H., 1971. The carbon ratio rule and petroleum potential in NPC Region
9: Illinois State Geol. Survey Illinois, Petroleum, 95: 142-151.
Boujo, A., Faye, B., Giot, D., Lucas, J., Manivit, H., Monciardini, C. and PrkvBt, L., 1980. The Early
Eocene of the Lake of Guiers (Western Senegal) - reflections of some characteristics of phosphate
sedimentation in Senegal. Soc. Econ. Paleo. Mineral. Sp. Pub. 29: 207-213.
Bouroz, A., 1966. Frtquence des manifestations volcaniques in Carboniftre supkrieur en France. Compt.
Rend., 236: 1025-1028.
Bowles, F.A., Angino, E.A., Hostermann, J.W., and Galle, O.K., 1971. Precipitation of deep-sea
palygorskite and sepiolite. Earth Planet. Sci. Lett., 11: 324-332.
Bowles, F.A., Bryant, W.R. and Wallin, C., 1969. Microstructure of unconsolidated marine sediments.
Jour. Sed. Petrol., 39: 1546-1551.
Bowles, F.A., Jack, R.N. and Carmichael, I.S.E., 1973. Investigation of deep-sea volcanic ash layers from
equatorial Pacific cores. Geol. SOC.Amer. Bull., 84: 2371-2388.
Boyce, R.E., 1972. Leg 11 grain size analyses; In: C.D. Hollister, J.I. Ewing et al., 1972, Initial Reports of
the Deep Sea Drilling Project, Vol. XI, Wash. (US.Gov. Printing Office), 1047-1057.
Boyer, P.S., Guinness, E.A., Lynch-Blosse, M.A. and Stolzman, R.A., 1977. Greensand fecal pellets from
New Jersey. Jour. Sed. Petrol., 47: 267-280.
Boyle, J.R., Voigt, G.K. and Sawhney, B.L., 1967. Biotite flakes alteration by chemical and biological
treatment. Science, 155: 193-195.
Bradley, W.F., 1940. Structure of attapulgite. Amer. Miner., 25: 405-410.
Bradley, W.F. and Weaver, C.E., 1956. A regularly interstratified chlorite-vermiculite clay mineral. Amer.
Min., 41: 497-504.
Bradley, W.H., 1964. Geology of Green River Formation and associated Eocene rocks in southwestern
Wyoming and adjacent parts of Colorado and Utah. U.S. Geol. Survey Prof. Paper 496A. 83 p.
Bradshaw, M.J., 1975. Origin of montmorillonite bands in the Middle Jurassic of eastern England. Earth
Planet. Sci. Lett., 26: 245-252.
Braid, S.P., 1982. Clay mineral burial diagenesis in Tertiary sediments from the eastern flank of the Niger
Delta. Ph.D. Thesis, U. Cincinnati, Ohio, 180 p.
Braitsch, 0..1960. Mineral paragenesis und petrologie der Strassfurtsalz in Reyerhaursen. Kali U.
Steinsalz, I: 1-14.
Braitsch, 0.. 1971. Salt Deposits, Their Origin and Composition. Springer-Verlag Co., Berlin, 297 p.
Brauner, K. and Preisinger, A., 1956. Structure of sepiolite. Petrog. Mitt., 6: 120-140.
Bricker, O.P. and Garrels, R.M., 1965. Mineralogic factors in natural water equilibria. In: S. Faust and
J.V. Hunter (Editors), Principles and Applications of Water Chemistry. John Wiley and Sons. Inc.,
New York, 449-469.
Brigatti, M.F. and Poppi, L., 1980. Vermiculite and its relation to parent materials as revealed by
chemical features. Miner. Petrog. Acta, 24: 123-134.
Brigatti, M.F., 1981. Hisingerite: A review of its crystal chemistry. Proc. Inter. Clay Conf., Bologna,
97-109.
Brigatti, M.F. and Poppi, L., 1984. Crystal chemistry of corrensite: a review. Clays Clay Miner., 32:
391-399.
Brindley, G.W., 1951. The kaolin minerals. In: G.W. Brindley (Editor), X-ray Identification and
Structure of Clay Minerals . Miner. SOC.London, 32-75.
Brindley, G.W., 1959. X-ray and electron diffraction data for sepiolite. Amer. Miner, 44: 495-500.
Brindley, G.W., 1961. Kaolin, Serpentine and Kindred Minerals. In: G. Brown (Editor), “The X-ray
Identification and Crystal Structure of Clay Minerals.” Min. Soc. London, Chap. 11, p. 51-131.
Brindley, G.W., 1980. Order-disorder in clay mineral structures. In: G.W. Brindley and G. Brown
730

(Editors). Crystal Structures of Clay Minerals and Their X-ray Identification. Mineralogical Soc..
London, 125-196.
Brindley. G.W.. 1982. Chemical composition of berthierines - a review. Clays Clay Miner.. 30: 153-155.
Brindley. G.W.. Bish, D.L. and Wan, H.-M.. 1977. The nature of kerolite. its realtion to talc and
stevensite. Miner. Mag., 41: 443-452.
Brindley, G.W. and Brown, G., 1980. Crystal Structures of Clay Minerals and Their X-ray Identification.
Mineralogical Society, London. 495 p.
Brindley, G.W.. and Gillery. F.H.. 1956. X-ray identification of chlorite species. Amer. Miner., 41:
169-1 86.
Brindley. G.W. and Kao, C-C, 1980. Formation, compositions and properties of hyrdroxy-Al-and
hydroxy-Mg-montmorillonite. Clays Clay Miner., 28: 435-443.
Brindley. G.W., and MacEwan, D.M.C., 1953. Structural mineralogy of clays. Ceramics - A symposium.
British Ceramic Soc.,Stoke-on-Trent. 15-59.
Brindley, G.W. and Robinson. K.. 1948. X-ray studies of halloysite and metahalloysite. 1. The structure
of metahalloysite, an example of a random layer lattice. Miner. Mag., 28: 383-406.
Brindley, G.W. and Youell, R:F., 1953. Ferrous chamosite and ferric chamosite. Min. Mag.. 30: 57-70.
Brookins, D.G., 1976. Radiometric age determination of Silurian glauconite: discussion. Amer. Assoc.
Pet. Geol. Bull., 60: 883-887.
Brooks, R.A. and Ferrell, R.E., Jr.. 1970. The lateral distribution of clay minerals in Lake Pontchartrain
and Maurepas, Louisiana. Jour. Sed. Petrol., 40: 855-863.
Brown. G., 1953. The dioctahedral analogue of vermiculite. Clay Miner., 2: 64-70.
Brown, E.H., 1967. The greenschist facies in part of Eastern Otago. New Zealand. Contrib. Miner.
Petrol., 14: 259-292.
Brown, B.E. and Bailey, S.W., 1962. Chlorite polytypism, 1. Regular and semi-random one-layer
structures. Amer. Miner., 47: 819-850.
Brown, B.E. and Bailey, S.W., 1963. Chlorite polytypism. 11. Crystal structure of a one-layer cr-chlorite.
Amer. Miner., 48: 42-61.
Brown. B.E. and Jackson, M.L., 1958. Clay mineral distribution in the Hiawatha sand soils of northern
Wisconsin. Clay Clay Miner., 5 : 213-226.
Brown. L.F.. Jr.. Bailey, S.W., Clive. L.M. and Lister, J.S., 1977. Clay mineralogy in relation to deltaic
sedimentation patterns of Des Moinesian cyclotherms in Iowa-Missouri. Clays Clay Miner.. 25:
171- 186.
Brown, G. and Brindley, G.W.. 1980. X-ray diffraction procedures for clay mineral identification. In:
G.W. Brindley and G . Brown (Editors), Crystal Structures of Clay Minerals and Their X-ray
Identification. Mineralogical Soc., London, 305-360.
Bruce, C.H.. 1984. Smectite dehydration - its relation to structural development and hydrocarbon
accumulation in Northern Gulf of Mexico Basin. Amer. Assoc. Pet. Geol. Bull.. 68: 673-683.
Brun, J. and Chagnon, A., 1979. Rock stratigraphy and clay mineralogy of volcanic ash beds from the
Black River and Trenton Groups (Middle Ordovician) of southern Quebec. Can. Jour. Earth Sic.. 16:
1499- 1507.
Brydon. J.E. and Turner, R.C.. 1972. The nature of Kenya vermiculite and its aluminum hydroxide
complexes. Clays Clay Miner., 20: 1 - 11.
Buckley. H.A., Bevan, J.C., Brown. K.M.. Johnson. L.R. and Farmer. V.C.. 1978. Glauconite and
celadonite: two separate mineral species. Miner. Mag., 42: 373-382.
Bundy, W.B. and Murray, H.H.. 1959. Argillization in the Cochiti mining district. New Mexico. Clays
Clay mlner., 6: 342-368.
Burger. K.. 1964. Stratigraphie und petrographie des neuen kaolin-kohlentonsteins des Flozes H. (EB) in
den Oberen Essener Schichten (Westfal. B) des Rahrkarbons. Fortschir. Geol. Rheinland Westfalen.
12: 451-472.
Burley, S.D., 1984. Patterns of diagenesis in the Shenvood Sandstone Group (Triassic) United Kingdom.
Clay Miner. 19: 403-440.
Burst, J.F., 1958. Mineral heterogeneity in "glauconite" pellets. Amer. Miner., 43: 481-497.
Burst, J.F., 1959. Postdiagenetic clay mineral environmental relationships in the Gulf Coast Eocene.
Clays Clay mlner.. 6: 327-341.
731

Burst, J.F., 1969. diagenesis of Gulf Coast clayey sediments and its possible relation to petroleum
migration. Bull. Am. Assoc. Pet., 53: 73-93.
Burtner, R.L. and Warner, M.A., 1984. Illite/smectite diagenesis and maturation of organic matter in
Cretaceous Mowry and Skull Creek shales of the northern Rocky Mountain region. Clay Miner. SOC.
Prog. with Abs.. Baton Rouge, 35.
Bustillo, A.M. and Iglesia, A.La., 1978. Procesos de formaci6n y diagtnesis en las rocas siliceas d e
Sagunto. Estudios Geol., 34: 167-174.
Butuzova, G. Yu., Drits, V.A., Lisitsyna N.A., Tsipurskii, S.I., and Dmitrik, A.L., 1979. Formation
dynamics of clay minerals in ore-bearing sediments in the Atlantis-2, Red Sea. Lithol. Miner.
Resources, 4: 23-32.
Butze, H., 1955. Lavatstrome and Aschenregen; die vulkanischen Ergcheinungen und Krafte, mit
Berichten von den grossten Vulkankatastrophen der Welt. Leipzig, F.A. Brocklaus, 244 p.
Buxton, B.P., 1968. Measure of the degree of chemical weathering of rocks. Jour. Geol., 76, 518-527.
Bystrom, A.M., 1956. Mineralogy of the Ordovician bentonite beds at Kinnekulle, Sweden. Sver. Geol.
Undersokn. Arsbok., 48: 1-62.
Bystrom, A., 1957. The clay minerals in the Ordovician bentonite beds in Billingen southwest Sweden.
Geol. Forn. 1 Stockholm, Forh, 79: 52-56.
Bystrom-Asklund, A.M., Baadsgaard. H.B. and Folinsbee, R.E., 1961. K/Ar age of biotite, sanidine, illite
from Millde Ordovician bentonites at Kinnekulle, Sweden. Geol. Forn. Stockholm Forh., 83: 92-96.
Caballero, M.A. and Martin Vivaldi, J.L., 1973. Distribution of clay minerals in the Spanish Triassic
sedimentary basins. Internatl. Clay Conf.. Proc., Madrid, 1972, 259-268.
Cadigan, R.A., 1967. Petrology of the Morrison Formation in the Colorado Plateau Region. U.S. Geol.
Sur. Prof. Paper 556, 113 p.
Cahet, G . and Giresse, P., 1983. Le r61e des supports organiques marins et en particulier des pelote
ftcales dans la mintralisation. Bull. Soc. Gtol. France, 25: 523-531.
Cahoon, H.P., 1954. Saponite near Milford, Utah. Amer. Miner, 39: 222-230.
Cailltre, S., 1936. Thermal studies. Bull. SOC.France. Minir., 59: 163-326.
Cailltre, S., 1962. A new type of chlorite. Bull. Groupe France. Argils., 13: 35-37.
Cailltre, S., and Htnin, S., 1949. Experimental formation of chlorites from montmorillonites. Miner.
Mag., 28: 612-620.
Callen, R.A., 1977. Late Cainozoic environments of northeastern South Australia. Jour. Geol. Soc. Aust..
24: 151-169.
Callen, R.A., 1984. Clays of the palygorskite-sepiolite group: depositional environment, age and
distribution. In: A. Singer and E. Galan (Editors), Palygorskite-Sepiolite Occurrences. Elsevier.
Amsterdam, 1-38.
Calvert, C. and Pevear, D., 1983. Paleopedogenic (?) mixed-layer kaolinite/smectite from 9000 ft. in a
Gulf Coast well. Clay Min. SOC.Ann. Meet., Buffalo, Prog. with Abs., 50.
Cameron, E.M. and Carrels, R.M., 1980. Geochemical composition of some Precambrian shales from the
Canadian Shield. Chem. Geol., 28: 181-197.
Campbell, F.A. and Lerbekmo, J.F.. 1963. Mineralogic and chemical variations between Upper Creta-
ceous continental Belly River shales and marine Wapiabi shales in western Alberta, Canada.
Sedimentology, 2: 215-226.
Campbell, F.A. and Williams, G.D., 1965. Chemical composition of shales of Mannville Group (Lower
Cretaceous) of central Alberta, Canada. Amer. Assoc. Pet. Geol. Bull.. 49: 81-87.
Carden, K.L., Beardsley, G.F. and Pak, H., 1971. Particle size distribution in the Eastern Equatorial
Pacific. Jour. Geoph. Res., 76: 5070-5077.
Carmouse. J-P., Pegro, G. and Berrier, J., 1977. Sur la nature des smectites de ndoformation du lac Tchad
et leur distribution spatiale en fonction hydrogiochimiques. C.R. Acad. Sc. Paris, 284, Sirve D.
615-619.
Carozzi. A.V., 1960. Microscopic Sedimentar Petrography. John Wiley and Sons, New York. 485 p.
Carrigy, M.A. and Kramers, J.W., 1973. Guide to the Athabasca oil sands area. Alberta Res. Council.
Edmonton Alta., Info. Series 65.
Carrigy, M.A. and Mellon, G.B., 1964. Authigenic clay mineral cements in Cretaceous and Tertiary
sandstones of Alberta. Jour. Sed. Petrol., 34: 461-472.
732

Carroll, D., 1969. Chlorite in central North Pacific Ocean sediments. Proc. Internatl. Clay Conf., Tokyo,
1: 335-338.
Carroll, D. and Starkey, H., 1960. Effect of sea water on clay minerals. Clays Clay Miner., Proc. 7th Natl.
Conf, 80-101.
Carstea, D.D., 1968. Formation of hydroxy-Al and-Fe interlayers in montmorillonite and vermiculite:
influences of particle size and temperature. Clays Clay Miner., 16: 231-238.
Carstea, D.D., Hanvood, M.E. and Knox, E.G., 1970. Comparison of iron and aluminum hydroxy
interlayers in montmorilonite and vermiculite: 1. Formation. Soil. Sci. Soc. Amer. Proc., 34: 517-521.
Carver, R.E., 1972. Stratigraphy of the Jackson Group in eastern Georgia. Southeastern Geol., 14:
153-181.
Case, E.C., 1926. The environment of vertebrate life in the late Paleozoic of regions other than North
America. Carnegie Inst. Wash. Pub. 375, 211 p.
Castaiio, J.R. and Sparks, D.M., 1974. Interpretation of vitrinite reflectance measurements in sedimen-
tary rocks and determination of burial history using vitrinite reflectance and authigenic minerals.
Geol. SOC.Amer. Special Paper 153, 31-52.
Cauffman, L.B. Jr. and Bayliss. P., 1973. Clay mineralogy and depositional environment of Melville
Island. Clays Clay Mineral., 22nd Ann. Conf., Prog. with Abs., 51.
Cavaroc, V.V., Padgett, G., Stephens, D.G., Kanes, W.H., Boudda, A. and Woollen, I.D., 1976. Late
Paleozoic of the Tindouf Basin - North Africa. Jour. Sed. Petrol. 46: 77-88,
Cecconi, S., Ristori, G.G. and Vidrich, V., 1975. The role of octahedral Fe in biological weathering of
biotite. Proc. Inter. Clay Conf., Mexico City, 623-628.
Cecil, C.B., Stanton, R.W., Neuzil, S.G., Dulong, F.T., and Ruppert, L.F., 1983. Paleoclimatic controls
on coal beds and associated rocks in the central Appalachian Basin. Geol. Soc. Amer. Bull., Prog.
with Abs., 554.
Chamley, H., 1976. MinCralogie des argiles, lithologie et tectonique, dans le Pliockne du Capo Rossello
(Sicile). C.R. Somm. SOC.GCol. France, 18: 39-41.
Chamley, H., 1979a. MinCralogie des argiles, lithologie et tectonique, dans le Plioctne du Capo Rossello
(Sicile). C.R. Somm. SOC.Geol. Fr., 2: 39-41.
Chamley, H., 1979b. North Atlantic clay sedimentation and paleoenvironment since the Late Jurassic.
In: M. Talwani, W. Hay and W.B.F. Ryan (Editors), Deep Drilling results in the Atlantic Ocean:
Continental Margins and Paleoenvironment. Amer. Geophys. Union., 342-361.
Chamley, H. and d’Argoud, G.G., 1978. Clay mineralogy in volcanogenic sediments. In: K.J. Hsu. L.
Montadert et al., (Editors), Initial Reports of Deep Sea Drilling Project, 42. Part 1. Wash. (U.S. Gov.
Printing Office), 395-397.
Chamley, H. and Colomb, E., 1967. Premieres donnCes sur le ddimentation, argileuse du Miocene
superieur dans le bassin de Cucurron (Vaucluse). Prtsence de niveaux lacustres i attapulgite. Des
Stances SOC.Gtol. France, 6: 230.
Chamley, H., Dunoyer de Segonzac, G . and Mtlikres, F.. 1978. Clay minerals in Messinian sediments of
the Mediterranean area. In: K.J. Hsu, L. Montadert et al. (Editors), Initial report DSDP, 42, Part 1.
Wash. ( U S Gov. Printing Office), 389-394.
Chamley, H., Deynoux, M., DArgoud, G., and Trompette, R., 1977. MinCraux argileux des formations
glaciaires el du leur substratum dans le PrCcambrian. Afrique Occidentale. Sci. Geol. Bull. 30:
207-227.
Chamley, H. and Masse, J-P., 1975. Sure la signification des mintraux argileux dans les sediments
Barremiens et Bedouliens de Provence (SE de la France). Internatl. Sedimentological Congress, 9:
25-29.
Chang, H.K., Mackenzie, F.T. and Schoonmaker, J., 1986. Comparison between the diagenesis of
dioctahedral and trioctahedral smectite, Brazilian offshore basins. Clays Clay Miner., 34: 407-423.
Chaudhuri. S. and Brookins, D.G., 1969. The isotopic age of the Flathead Sandstone (Middle Cambrian),
Montana. Jour. Sed. Petrol., 39: 364-368.
Chen, B.Y., 1972. Clay minerals from the alteration of mafic and intermediate igneous rocks in Taiwan
and neighboring islands. Proc. Geol. Soc. China, 15: 45-63.
Chen, P.Y., 1978. Minerals in bottom sediments of the South China Sea. Geol. SOC.Arner. Bull.. 89:
21 1-222.
733

Chennaux, G. and Dunoyer de Sgonzac, 1968. Etude pktrographique de la pyrophyllite du Silurian et du


Devonian au Sahara. Repartition et origene. Bull. Serv. Carte Gkol. Als. Lorr., 20: 195-210.
Chennaux, G., Esquevin, J. and Jourdan, A., 1985. X-ray mineralogy and mineral geochemistry of
Cenozoic strat (leg 80) and petrographic study of associated pebbles. In: P.C. de Graciansky, C. W.
Poag et al., Init. Repts. DSDP, 8, Part 2, Wash. ( U S . Govt. Printing Office), 1019-1043.
Chesnut, D.R., 1984. Source of the volcanic ash deposit (flint clay) in the fire clay of the Appalachian
Basin. Geol. SOC.Amer. Bull., Prog. with Abs., 16: 128.
Chester, R.H., Eldenfield, H., Griffin, J.J., Johnson, L.R., and Padgham, R.C., 1972. Eolian dust along
the eastern margin of the Atlantic Ocean. Marine Geol., 13: 91-105.
Chester, R., Sharples, E.J. and Sanders, G.S., 1985. The concentration of particulate aluminum and clay
minerals in aerosols from the northern Arabian Sea. Jour. Sed. Petrol., 55: 37-41.
Chester, R., Stoner, J.H. and Johnson, L.R., 1974. Montmorillonite in surface detritus. Nature, 249:
335-336.
Chichester, F.W., Youngberg, C.T., and Harward, M.E., 1969. Clay mineralogy of soils formed on
Mazama Pumice. Soil Sci. SOC.Amer. Proc., 33: 115-120.
Chilingarian, G.V., 1983. Compactional diagenesis. In: A. Parker and B.W. Sellwood (Editors), Sediment
Diagenesis, Reidel Pub. Co., Dordrecht, 57-168.
Chilingar, G.V. and Knight, L., 1960. Relationship between pressure and moisture content of kaolinite,
illite, and montmorillonite clays. Bull. Amer. Assoc. Pet. Geol., 44: 101-106.
Chowdhury, A.N., 1982. Smectite, zeolite, biotite and apatite in the Corallian (Oxfordian) sediments of
the Baulking area of Berkshire, England. Geol. Mag., 119: 487-496.
Christ, C.L., Hathaway, J.C., Hostetler, P.B., and Sheppard, A.O., 1969. Palygorskite: new X-ray data.
Amer. Miner. 54: 198-205.
Churchman, G.J., 1980. Clay minerals formed from micas and chlorites in some New Zealand soils. Clay
Miner., 15: 59-76.
(%el, B. and Machajdik, D., 1981. Potassium and ammonium-treated montmorillonite. 1. lnterstratified
structures with ethylene glycol and water. Clays Clay Miner., 29: 4 - 4 6 ,
Cimbilnik6vB. A,, 1971. Chemical variability and structural heterogeneity of glauconites. Amer. Miner.,
56: 1385-1392.
Claridge, G.G.C., 1965. The clay mineralogy and chemistry of some soils from the Ross Dependency,
Antarctica. N.Z. Jour. Geol. Geophys., 18: 186-197.
Clark, F.W., 1924. Data of Geochemistry. U.S. Geol. Survey Bull. 770, 140 p.
Clarke, O.M. Jr., 1965. Clay deposits of the Tuscaloosa Group in Alabama. Geol. Surv. Ala., Reprint
Series 9: 495-507.
Clarke, O.M. Jr., 1973. Gibbsite in weathered granitic rocks of Alabma. Southeast. Geol., 14: 203-212.
Clauer, N., Holtzapffel, T., Bonnot-Courtois, C. and Steinberg, M., 1984. Sr isotopes as indicators for the
origin and evolution of detrital smectites from northern Atlantic Ocean. Clay Miner. SOC.Meet., Prog.
with Abs., 37.
Clauer, N. and Kroner, A,, 1978. Strontium and argon isotopic homogenization of pelitic sediments
during low-grade regional metamorphism: the Pan-African Damara sequence of Northern Namibia
(South West Africe). Earth Planet. Sci. Let., 43: 117-131.
Clayton, R.N., Rex, R.W., Syers, J.K. and Jackson, M.L., 1972. Oxygen isotope abundance in quartz
from Pacific pelagic sediments. Jour. Geophys. Res., 177: 3907-3910.
Clemency, C.V., 1975. Simultaneous weathering of a granitic gneiss and an intrusive amphibolite dike
near SBo Paulo, Brazil, and the origin of clay minerals. Proc. Int. Clay Conf., Mexico City. 15-25.
Cluff, R.M., 1980. Paleoenvironment of the New Albany Shale Group (Devonian-Mississippian) of
Illinois. Jour. Sed. Petrol., 50: 767-780.
Cluff, R.M., Reinbold, M.L. and Lineback, J.A., 1981. The New Albany Shale Group of Illinois. Ill. State
Geol. Surv. Circ. 518, 83 p.
Coakley, R.D., and Rust, B.R., 1968. Sedimentation in an Arctic lake. Jour. Sed. Petrol.. 38: 1290-1300.
Cocker, J.D., 1985. Critical point drying of illite-bearing sandstones: morphology and permeability
effects. Clay Miner. Soc. Meet., Baton Rouge, Prog. with Abs.. 37.
Cole, M.J., 1980. Mineralogy and petrology of very-low-metamorphic grade Archaean banded iron-for-
mations, Weld Range, Western Australia. Amer. Miner., 65: 8-25.
734

Cole, R.D. and Picard, M.D., 1975. Primary and secondary sedimentary structures in oil shale and other
fine-grained rocks, Green River Formations (Eocene), Utah and Colorado. Utah Geology. 2: 49-67.
Cole, T.G. and Shaw, H.F., 1983. The nature and origin of authigenic smectites in some recent marine
sediments. Clays Clay Miner.. 18: 239-252.
Coleman, J.M., 1966. Ecological changes in a massive fresh water clay sequence. Gulf Coast Assoc. Geol.
SOC.Trans., 16: 159-174.
Coleman, J.M. and Wright, L.D., 1971. Analysis of major river systems and their deltas, procedures and
rationale with two examples. La. State U., Coastal Studies Inst. Tech. Rept. 95, 125 p.
Collins, K. and McGown, A,, 1974. The form and function of microfabric features in a variety of natural
soils. Geotechnique, 24, 223-254.
Condie, K.C., 1982. Plate Tectonics and Crustal Evolution. Pergamon Press, N.Y.,310 p.
Cook, M.G. and h c h , C.I., 1963. Negative change of dioctahedtral micas as related to weathering. Clays
Clay Miner., 11: 47-65.
Cooke, G.E., 1977. Pore water constituents of deeply buried pelagic Gulf Coast muds as indicators o f
diagenetic alteration. M.S. Thesis, Georgia Inst. Tech., Atlanta, 173 p.
Coombs, D.S., 1961. Some recent work on the lower grades of metamorphism. Aust. Jour. Sci, 24:
203-21 5.
Copeland, R.A., Frey, F.A., and Wones, D.R., 1971. Origin of clay minerals in a Mid-Atlantic Ridge
sediment. Earth Planet. Sci. Let., 10: 186-192.
Corliss, J.B., Lyle, M. Drymond, J., and Crane, K., 1978. The chemistry of hydrothermal mounds near
the Gallapagos Rift. Earth Planet Sci. Lett. 40: 12-24.
Cornet, C., 1977. Etude prkliminaire des mintraux argileux des shies continentales du CrttacC Suptrieur
et du Tertiaire des fossks Nord-Varois (Provence). Geol. Mkditerrankenne. 4: 379-382.
Correns, C.W., 1968. Introduction to Mineralogy. Springer-Verlag, N.Y., 484 p.
Court. J.E., Goldman, C.R.. and Hyme, N.J., 1972. Surface sediments in Lake Tahoe. California-Nevada.
Jour. Sed. Petrol., 42: 359-377.
Couto Anjos. S.M., 1984. Absence of diagnesis in Cretaceous marine shales from Campos Basin, Brazil.
Inter. Clay Conf. Prog. with Abs., Denver, 43.
Couture, R.A., 1977. composition and origin of palygorskite-rich and montmorillonite-rich zeolite-con-
taining sediments from the Pacific Ocean. Chem. Geol. 19: 113-130.
Cowperthwaite, LA., and Fitch. F.I., 1972. Sedimentation, petrogenesis and radioisotopic age of the
Cretaceous Fuller's Earth of southern England. Clay Miner., 9: 309-327.
Cubitt, J.M., 1979. The geochemistry, mineralogy and petrology of Upper Paleozoic shales of Kansas.
Kansas Geol. Sur. Bull. 217, 117 p.
Cummings, J.C. and Beattie. J.A.. 1963. Distribution of clay minerals in gradded beds of Tyee Formation
(Eocene), Western Oregon. Geol. Soc. Amer. Sp. Paper 76. Abs.. 197.
Curry, J.R., 1965. Late Quarternary history, continental shelves of the United States. In: H.E. Wright, Jr.,
and D.G. Grey (Editors), The Quarternary of the United States. Princeton Univ. Press. Princeton,
N.J.. 723-736.
Curtin. D. and Smillie, G.W., 1981. Composition and origin of smectite in soils derived from basalt in
northern Ireland. Clays Clay Miner., 29: 277-284.
Curtis, C.D.. 1978. Possible links between sandstone diagenesis and depth-related geochemical reactions
occurring in enclosing sandstones. Jour. Geol. Soc. Lond., 135: 107-117.
Curtis, C.D., Ireland, B.J., Whiteman, J.A., Mulvaney. R. and Whittle. C.K., 1984. Authhigenic chlorites:
problems with chemical analysis and structural formula calculations. Clay Miner., 19: 471 -481.
Curtis, C.D. and Spears, D.A., 1971. Diagenetic development of kaolinite. Clays Clay Miner., 19:
219-227.
Dahl, H.M., 1965. Clay mineralogy of some Permian bentonites form the Delaware Basin area, Texas.
Amer. Min., 50: 1637-1646.
Dahl, W.M., 1984. Progressive burial diagenesis in Lower Tuscalosa sandstones. Louisiana and Mi
sippi. Clay Clay Miner., Prog. with A h . , 42.
Dallmeyer, R.D., 1979. 4"Ar/3yAr dating: principles, techniques, and applications in orogenic terrains.
In: E. Jager and J.C. Hunziker (Editors). Lectures in Isotope Geology, Springer-Verlag, Berlin.
77-104.
735

Dalrymple, G.B. and Lamphere, M.A., 1971. 4"Ar/3qAr techniques of K-Ar dating. A comparison with
the conventional techniques. Earth Planet. Sci. Lett., 12: 300-308.
Dalrymple, G.B. and Lamphere, M.A., 1974. 4°Ar/39Ar age spectra of some undisturbed terrestrial
samples. Geochim. Cosmochim, Acta, 38: 715-738.
Damuth, J.E. and Kumar, N., 1975. Amazon Cone: morphology, sediments, age and growth pattern.
Geol. SOC.Amer. Bull., 86: 863-878.
Dandois, par Ph., 1981. Diagenese et metamorphisme des domaines Caledonien et Hercynien de la Vallee
de la Meuse entre Charleville-Mezieres et Namur (Ardennes Franco-Belges). Bull. Soc. Belge Geol.,
90: 299-316.
Dangit, A., 1985. Kaolinization of bauxite: a study in the Vlasenica bauxite area, Yugoslavia. I.
Alteration of matrix. Clays Clay Miner., 517-524.
DAnglejan, B.F. and Smith, E.C., 1973. Distribution, transport, and composition of suspended matter in
the St. Lawrence Estuary. Can. Jour. Sci., 10: 1380-1396.
Davies, D.K., Almon, W.R., Bonis, S.B. and Hunter, B.E., 1979. Deposition and diagenesis of Tertiary-
Holocene volcaniclastics, Guatemala. SEPM Spec. Pub. No. 26, Tulsa, 281-306.
Davis, A.G., 1967. The mineralogy and phase equilibrium of Keuper Marl. Quart. Jour. Eng. Geol., 1:
25-38.
Davis, J.C., 1970. Petrology of Cretaceous Mowry Shale of Wyoming. Amer. Assoc. Pet. Geol. Bull., 54:
487-502.
Davis, R.A. Jr., 1983. Depositional Systems. Prentice Hall, N.J., 669 p.
Davis, R.A. Jr., 1978. Coastal Sedimentary Environments, Springer-Verlag, 420 p.
Dean, R.S., 1962. A study of St. Lawrence lowland shales. Ph.D. Thesis, McGill U. Montreal, Canada
236 p.
Dean, R.S., 1983. Authigenic trioctahedral clay minerals coating Clearwater Formation sand grains at
Cold Lake, Alberta, Canada. Clay Miner. Soc. meeting, Buffalo, Program with Abs., 79.
Dean, R.S., 1985. Burial diagenesis of weathered illite, Beaufort Basin, Canada. Inter. Clay Conf. Prog.
with Abs., Denver, 51.
Dean, W.E. and Gorham, E., 1976. Major chemical and mineralogical components of profundal surface
sediments in Minnesota Lakes. Limnol. Oceanogr., 21: 259-284.
Decommer, H. and Chamley, H., 1981. Environments mtsozoiques du nord de la France, d'aprts les
donnkes des argiles et du palynoplancton. C.R. Acad. Sc. Paris, 293: 695-698.
Deer, W.A., Howie, R.A. and Zussman, J., 1962. Rock-Forming Minerals, Vol. 3 Sheet Silicates. John
Wiley and Sons, Inc., 270 p.
Deffeyes, K.S., 1970. The axial valley: a steady-state feature of the terrain. In: H. Johnson and B.L.
Smith (Editors), The Megatectonics of Continents and Oceans. Rutgers University Press, 194-222.
Degens, E.T., Williams, E.G. and Keith, M.L., 1957. Environmental studies of Carboniferous sediments,
Part I: Geochemical criteria for differentiating marine from fresh-water shales. Amer. Assoc. Petrol.
Geol. Bull., 41: 2427-2355.
Deike, R.G. and Jones, B.F., 1980. Provenance, distribution and alteration of volcanic sediments in a
saline alkaline lake. In: A. Nissenbaum (Editor), Hypersaline Brines and Evaporitic Environments.
Dev. Sed. 28. Elsevier, Amsterdam, 167-193.
DeKimpe, C. and Tardy, Y., 1967. Infrared-spectroscopic study of the conversion of biotite to kaolinite.
Groupe Fr. Argiles, 19: 81-85.
DeLange, G.J., 1986. Early diagenetic reactions in interbedded pelagic and turbiditic sediments in the
Nares abyssal plain (western North Atlantic): consequences for the composition of sediment and
interstitial water. Geochim. Cosmochim. Acta, 50: 2543-2561.
DeLange, G.J. and Rispens, F.B., 1986. Indication of a diagenetically induced precipitate of an Fe-Si
mineral in sediment from the Nares abyssal plain. Western North Atlantic. Marine Geol., 73: 85-97.
Delany, A.C., Parkins, D.W., Griffin, J.J., Goldberg, E.D. and Reimann, B.E.F., 1967. Airborne dust
collected at Barbados. Geochim. Cosmochim. Acta, 31 : 885-909.
Dell, C.I., 1971. Late Quaternary sedimentation in Lake Superior. Ph.D. Thesis, U. of Michigan.
Demaison, G.J., 1974. Relationship of coal rank to paleotemperatures in sedimentary rocks. Jour. Inter.
Petrog. Mat. Org. Disperste Sed. Comptes Rendus, 217-224.
Dennison, J.W., 1983. Internal stratigraphy of Devonian Tioga ash beds in Appalachian Valley and
Ridge Province. Geol. Soc. Amer. Prog. with Abs., 94: 557.
736

Deuser. W.G.. Ross, E.H. and Anderson, R.F., 1981. Seasonality in the supply of sediments to the deep
Sargasso Sea and implications for the rapid transfer of matter to the deep ocean. Deep Sea Research.
28A: 495-505.
Depetris. P.J. and Griffin. J.J.. 1968. Suspended load in the Rio de la Plata drainage basin. Sedimentol-
ogy. 11: 53-60.
Desprairies. A.. 1981. Authigenic minerals in volcanogenic sediments cored during Deep Sea Drilling
Project Leg 60. In: D.M. Hussong, S. Uyeda et al. (Editors). Initial Reports of the DSDP. 60. Wash.
( U S . Gov. Printing Office), 455-464.
Dewis, F.J., Levinson, A.A. and Bayliss, P.. 1972. Hydrogeochemistry of the surface waters of the
Mackenzie River drainage basin, Canada - IV. Born-salinity-clay mineralogy relationships in modern
deltas. Geochim. Cosmochim. Acta. 36: 1359-1375.
Divis. A.F. and McKenzie. J.. 1975. Experimental authigenesis of phyllosilicates from feldspathic sands.
Sedirnentology. 22: 147-155.
Dixon. J.B. and McKee. T.R., 1974. Spherical halloysite formation in a volcanic soil of Mexico. Int.
Congr.-Soil Sci., Trans. 10th. Moscow, VII. 115-124.
Dobryansky. A.F., 1963. La transformation du petrole brut dans la nature. Inst. Francais Petrole Rev.,
18: 41-49.
Dodd. C.G.. Conley. F.R. and Barnes, P.M., 1954. Clay minerals in petroleum reservoir sands and water
sensitivity effects. Clays Clay Miner.. Natl. Acad. Sci.. Natl. Res. Council Pub. 395: 221 -238.
Doehl, F.. Heling, D., Homann, W., Karweil, J., Teichmiiller, M., and Welte, D., 1974. Diagenesis of
Tertiary clayey sediments and included dispersed organic matter in relationship to geothermics in the
Upper Rhine Graben. In: J.H. I l k s and and K. Fuchs (Editors), Approaches to Taphrogenesis. E.
Schweizerbart'sche Veralagsbuchhandlung, Stuttgart, 192-207.
Dolcater. D.L.. Syers, J.K. and Jackson, M.L., 1970. Titanium as free oxide and substituted forms in
kaolinite and other soil minerals. Clays Clay Miner.. 18: 71-79.
Donaldson, A.C.. Renton. J.J. and Presley. M.W., 1983. Pennsylvanian deposy stems and paleoclimates
of Appalachians. Geol. Soc. Amer. Bull.. Prog. with Abs., 560.
Dodd. C.G., Conley, F.R.. and Barnes, P.M., 1955. Clay minerals in petroleum reservoir sands and water
sensitivity effects. Clays Clay Miner., Natl. Acad. Sci.-Natl. Res. Council Pub. 395: 221 -238.
Donnelly, T.W.. 1973. Circum-Caribbean explosive volcanic activity, evidence from Leg 15 sediments. In:
N.T. Edger. J.B. Saunders et al.. Initial Rpt. DSDP, 15. Wash. (U.S. Gov. Printing Office). 969-975.
Donnelly, T.W.. 1980. Secondarily modified sediments of the eastern Pacific: Major-element chemistry of
sites 420. 424, and 425, Deep Sea Drilling Project Leg 54. In: B.R. Rosendahl, R. Keinian et al..
Initial Reports of the Deep Sea Drilling Project, 54. Wash. ( U S . Gov. Printing Office). 329-338.
Donnelly, T.W. and Melson, W., 1973. Chemical studies of amphiboles, pyroxenes, and other minerals
from Caribbean deep-sea sediments, Leg 15. In: N.T. Edgar, J.B. Saunders et al.. Initial Reports of
the Deep Sea Drilling Project, 15, Wash. (U.S. Gov. Printing Office), 963-987.
Donnelly. T.W.. Thompson. G . and Salisbury. M.H.. 1979. The chemistry of altered basalt at Site 417.
Deep Sea Drilling Project Leg 51. In: T. Donnelly. J. Franchateau, W. Bryan. P. Robinson. M.
Flower, M. Salisbury et al., Initial Reports of the Deep Sea Drilling Project 51. 52. 54 Part 2, Wash.
(U.S. Gov. Printing Office), 1319-1330.
Dolt. R.H. Jr.and Batter. R.L., 1976. Evolution of the Earth. McGraw-Hill, N.Y.. 504 p.
Douglas, L.A., 1977. Vermiculite. In: J.B. Dixon and S.B. Weed (Editors), Minerals in Soil Environ-
ments. Soil Sci. Soc. Amer., Madison, Wisc., 259-292.
Doyle. L.J. and Sparks, T.N., 1980. Sediments of the Mississippi, Alabama. and Florida (MAFLA)
continental shelf. Jour. Sed. Petrol., 50: 905-916.
Drever. J.I.. 1971. Early diagenesis of clay minerals, Rio Ameca Basin, Mexico. Jour. Sed. Petrol.. 41:
982-994.
Drever. J.I.. 1974. The Mg problem. In: E.D. Goldberg (Editor). The Sea, 5: 337-357.
Drever. J.I., 1976. Chemical and mineralogical studies, Site 323. In: C.D. Hollister. C. Craddock et al.,
Initial Reports of the Deep Sea Drilling Project, V. 35, Wash. (U.S. Gov. Printing Office), 471-477.
Drever. J.I., 1982. The Geochemistry of Natural Waters. Prentice-Hall. Englewood Cliffs. N.J.. 388 p.
Droste. J.B., 1956. Alteration of clay minerals by weathering in Wisconsin tills. Geol. Soc. Amer. Bull.
67: 911-918.
737

Droste, J.B., 1961a. Clay mineral composition of sediments in some desert lakes in Nevada, California,
and Oregon. Science, 133: 1928.
Droste, J.B. 1961b. Clay minerals in sediments of Owens, China, Searles, Panamint, Bristol, Cadiz, and
Danby Lake Basins, California. Geol. SOC.Amer. Bull., 72: 1713-1722.
Droste, J.B., 1961c. Clay minerals in the playa sediments of the Mojave Desert, California. Calif. Div.
Mines Spec. Rept. 69, 37 p.
Droste, J.B., 1963. Clay mineral composition of evaporite sequences. Northern Ohio Geol. SOC.Mon. No.
1: 47-54.
Droste, J.B., Bhattacharya, N. and Sunderman, J.A., 1960. Clay mineral alteration in some Indiana soils.
Clays Clay Miner., 9: 329-342.
Droste, J.B. and Harrison, J.L., 1958. Division of Mississippian rocks in Indiana by clay-mineral
variation. Geol. SOC.Amer. Prog. with Abs., 1556.
Droste, J.B. and Vitaliano, C.J., 1973. Tioga bentonite (Middle Devonian) of Indiana. Clays Clay Min.,
21: 9-13.
Dubreuilh, J., Marchadour, P., and Thiry, M., 1984. Cadre gtologique et mintralogie des argiles des
Charentes, France. Clay Min., 19: 29-41.
Dumbleton, M.J. and West, G., 1966. Studies of the Keuper Marl. Ministry of Transport R.R.L., Report
40.
Duncan, J.R., Kulm, L.D. and Griggs, G.B., 1970. Clay mineral composition of Late Pleistocene and
Holocene sediments of Cascadia Basin, northeastern Pacific Ocean. Jour. Geol., 78: 213-228.
Dunoyer de Segonzac, G., 1964. Les argiles du Crttack suptrieur dans le Bassin de Doula (Cameroun).
Probltmes de diagtnkse. Alsace-Lorraine Service Carte Gtol. Bull., 17: 237-310.
Dunoyer de Segonzac, G., 1969. Les Mintraux argileux dans la diagtntse. Passage au Mttamorphisme
(Thesis Univ. Strasbourg) Mtm. Serv. Carte Gtol. Alsace-Lorraine. 29, 320 p.
Dunoyer de Segonzac, G., 1970. The transformation of clay minerals during diagenesis and low-grade
metamorphism. Sedimentology, 15: 281-346.
Dunoyer de Segonzac, G. and Chamley, H., 1968. Sur le rble jout la pyrophyllite comme margueur dans
les cycles sedimtntaires. C.R. Acad. Sc. Paris, 274-277.
Dunoyer de Segonzac, G., Ferrero, J. and Kubler, B., 1968. Sur la cristallinitk de l'illite dans la diagentse
et I'anchimttamorphisme. Sedimentology, 10: 137-143.
Dunoyer de Segonzac, G. and Hickel, D., 1972. Cristallochimie des phengites dans les guartzites micacks
metamorphques du Permo-Triassic des Alpes ptmontaise. Sci. Gtol. Bull., 25: 201-229.
Dutta, P.K. and Suttner, L.J., 1986. Alluvial sandstone composition and paleoclimate, 11. authigenic
mineralogy. Jour. Sed. Petrol., 56: 346-358.
Dyer, K.R., 1972. Sedimentation in estuaries. In: R.S.K. Barnes and J. Green (Editors), The Estuarine
Environment. Applied science, London, 10-32.
Dymond, J., Biscaye, P.E. and Rex, R.W., 1974. Eolian origin of mica in Hawaiian soils. Geol. SOC.
Amer. Bull., 85: 37-40.
Dymond, J., Corliss, J.B., Cobler, R. and Muratli, C.M., 1980. Composition and origin of sediments
recovered by deep drilling of sediment mounds, Galapagos Spreading Center. In: B.R. Rosendahl, R.
Hekinian et al., Initial Reports of the Deep Sea Drilling Project, V. 54, Wash. (U.S. Gov. Printing
Office), 377-385.
Dymond, J. and Eklund, W., 1978. A microprobe study of metalliferous sediment components. Earth
Planet. Sci. Lett., 40: 243-251.
Dypvik, H., 1979. Mineralogy and geochemistry of the Mesozoic sediments of Andoya, northern Norway.
Sed. Geol., 24; 45-67.
Eargle, D.H., 1968. Nomenclature of formations of Claiborne Group, middle Eocene, Coastal Plain of
Texas. US. Geol. Survey Bull. 1251-D, 25 p.
Earley, J.W., Brindley, G.W., McVeagh, W.J. and Vanden Heuvel, R.C., 1956. A regularly interstratified
montmorillonite-chlorite. Amer. Mineral., 41: 258-267.
Eaton, G.P., 1963. Volcanic ash deposits as a guide to atmospheric circulation in the geologic past. Jour.
Geoph. Res., 68: 521-528.
Eaton, G.P., 1964. Windbourne volcanic ash: a possible index to polar wandering. Jour. Geol., 172: 1-35.
Eberl, D., 1978. The reaction of montmorillonite to mixed-layer clay: the effect of interlayer alkali and
alkaline earth cations. Geochim. Cosmochim. Acta, 42: 1-7.
738

Eberl, D.D., 1980. Alkali cation selectivity and fixation by clay minerals. Clays Clay Miner., 28:
161-1 72.
Eberl, D. and Hower, J., 1976. Kinetics of illite formation. Geol. Soc. Amer. Bull., 187: 1326-1330.
Eberl, D.D., Jones, B.F. and Khoury. H.N., 1982. Mixed-layer kerolite/stevensite from the Amargosa
Desert, Nevada. Clays Clay Miner., 30: 321-326.
Eberl, D.D., Srodoh, J., and Northrop, H.R., 1986. Potassium fixation in smectite by wetting and drying.
In: J.A. Davis and K.F. Hayes (Editors), Geochemical Processes at Mineral Surfaces. Amer. Chem.
Soc. Symposium Series 323, 296-326.
Eberl, D., Whitney, G. and Khoury, H., 1978. Hydrothermal reactivity of smectite. Amer. Miner., 63:
401 -409.
Echle, W., 1961. Mineralogische Untersuchungen au Sedimenten des Steinmergelkeupers und der Roten
Wand aus der Ungebung vonGottingen. Beitr. Miner. Petrogr., 8: 28-67.
Eckel, E.C., 1904. On the chemical composition of American shales and roofing slates. Jour. Geol., 12,
25-29.
Eckhardt, F.J., 1965. Uber den einfluss der temperatur auf den kritallographischen ordmungsgard von
kaolinit. Proc. Inter. Clay Conf., Stockholm 1963, 137-145.
Edzwald, J.K. and OMelia, C.R., 1975. Clay distribution in recent estuarine sediments. Clays Clay
Miner., 23: 39-44.
Egashira, K., Dixon, J.B. and Hossner, L.R., 1981. High charge smectite from lignite overburden of East
Texas. Proc. Inter. Clay Conf., Bologna and Paris, 335-345.
Eggleton. R.A. and Bailey, S.W., 1967. Structural aspects of dioctahedral chlorite. Amer. Miner., 52:
673-689.
Eggleton, R.A. and Boland, J.N., 1982. Weathering of enstatite to take through a sequence of transitional
phases. Clays Clay Miner., 30: 11-20.
Eggleton, R.A. and Buseck, P.R., 1980. High resolution electron microscopy of feldspar weathering.
Clays Clay Miner., 28; 173-178.
Ehlers, E.G. and Hoover, K.V., 1961. The clay mineralogy of the shaly portions of the Brassfield
Limestone. Ohio Jour. Sci., 61: 227-234.
Ehlmann, A.J., 1968. Clay mineralogy of weathered products and of river sediments, Puerto Rico. Jour.
Sed. Petrol, 38: 885-894.
Ehlmann, A.J., Hulings, N.C. and Glover, E.D., 1963. Stages of glauconite formation in modern
foraminifera1 sediments. Jour. Sed. Petrol, 33: 87-96.
Eicher, D.L. and McAlester, A.L.. 1980. Hisotry of the Earth. Prentice-Hall. Inc., N.J., 413 p.
Einsele, G . , Gieskes, J., Curray, J., Moore, D., Aguayo, E. Aubry, M.-P., Fornari, D., Guerrero. J..
Kastner, M., Kelts, K., Lyle, M., Matoba, M., Molina-Cruz, A., Niemitz, J., Rueda, J., Saunders, A,.
Schrader, H., Simoneit, B., and Vacquier, V., 1980. Intrustion of bsaltic sills into highly porous
sediments, and resulting hydrothermal activity. Nature, 283: 44-445.
Einsele, G . and vonRad, U.. 1979. Facies and paleoenviornment of Lower Cretaceous sediments at
DSDP Site 397 and in the Aauin Basin (Northwest Africa). IN: U. vonRad, W.B.F. Ryan et al.
(Editors), Initial Reports of the DSDP. 47, Part 1, Wash. (U.S. Gov. Printing Office), 559-577.
Einstein, H.a. and Krone, R.B., 1961. Esturial sediment trnasport patterns. Jour.Hydraulics Div., Proc.
Amer. Soc. Civil Eng., HY 2, 51-59.
Elderfield. H., 1976. Hydrogenous material in marine seiments: excluding manganese nodules. In: J.P.
Riley and R. Chester (Editors), Chemical Oceanography. Academic Press, London. 137-216.
Elizalde, Par G. and Steinberg, M., 1973. Mineralogie de la fraction argileuse de la serie du Gondwana
DUruguay. Bull. Groupe franc. Argiles, XXV, 29-36.
Ellis, B.D., 1972. Holocene sediments of the South Atlantic Ocean: The calcite compensation depth and
concentration of calcite, opal and quartz. M.S. Thesis, Oregon State Univ., Corvallis, OR., 62 p.
Elprince, M., Mashhady, A.S. and Aba-Husayn, M.M., 1979. The occurrence of pedogenic palygorskite
(attapulgite) in Saudi Arabia. Soil Sci., 128: 211-218.
Eisma, D. and van der Marel, H.W., 1971. Marine muds along the Guyana coast and their origin from
the Amazon Basin. Contr. Miner. Petrol., 31: 321-334.
Emelyanov, E.M. and Shimkus, K.M., 1971. Suspended matter in the Mediterranean Sea. In: D.J.
Stanley, (Editor), The Mediterranean Sea. Dowden, Hutchinsin and Ross, Inc., Stroudsburg, Pa.,
417-470.
739

Ermery, K.O. and Honjo, S., 1979. Surface suspended matter off Western Africa: relations of organic
matter, skeletal debris and detrital minerals. Sedimentology, 26: 775-794.
Emery, K.O. and Milliman, J.D., 1978. Suspended matter in surface waters: influence of river discharge
and of upwelling. Sedimentology, 25: 125-140.
Enos, P. and Freeman, T., 1979. Lower Cretaceous peritidal limestones at 2. 700 m depth, Blake nose,
Atlantic Ocean. Geol., 7: 83-87.
Enu, E.I., 1980. Lithostratigraphy and sedimentology of the Upper Cretaceous Formations in the Cham
area (Upper Benue Trough, Nigeria). Annales Museum d’Historie Nat. de Nice, 6; 60-81.
Epshteyn, O.G., 1978. Mesozoic-Cenozoic climates of northern Asia and glacial-marine deposits. Int.
Geol. Rev., 20: 49-58.
Epstein, A.G., Epstein, J.B. and Harris, L.D., 1976. Conodont color alteration-an index to organic
metamorphism. U.S. Geol. Survey Prof. Paper 995, 27 p.
Ernst, W.G., 1963. Significance of phengitic micas from low-grade schists. Amer. Miner., 48: 1357-1 373.
Eroshchev-Shak, V.A., 1970. Mixed-layer biotite-chlorite formed in the course of local epigenesis in the
weathering crust of a biotite gneiss. Sedimentology, 15: 115-121.
Eslinger, E.V. and Savin, S., 1973a. Mineralogy and oxygen isotope geochemistry of the hydrothermally
altered rocks of the Ohaki-Broadlands, New Zealand geothermal area. Amer. Jour. Sci.. 273: 240-267.
Eslinger, E.V. and Savin, S.M., 1973b. Oxygen isotope geothermometry of the burial metamorphism of
the Precambrian Belt. Bull. Geol. Soc.Am., 84: 2549-2560.
Eslinger, E.V., Savin, S.M. and Yeh, H.-W., 1979. Oxygen isotope geothermometry of diagenetically
altered shales. Soc. Econ. Paleo. Miner., Spec. Pub. 26, 113-124.
Eslinger, E.V. and Sellars, B., 1981. Evidence for the formation of illite from smectite during burial
metamorphism in the Belt Supergroup, Clark Fork, Idaho Jour. Sed. Petrol, 51: 203-216.
Esquavin, J., 1969. Influence de la composition chimique des illite sur leur cristallinite. Bull. Centre Rech.
Pau-S.N.P.A., 3: 147-154.
Esteoule-Choux, J., 1984. Palygorskite in the Tertiary deposits of the Armorican Massif. In: A. Singer
and E. Galan (Editors), Palygorskite-Sepiolite, Developments in Sedimentology, 37. Elsevier,
Amsterdam, 75-86.
Esteoule, J. and Esteoule, J., 1976. Kaolin deposits in France. The 7th Symp. on genesis of kaolin, Tokyo,
173-82.
Eswaran, J. and Barzanji, J.F., 1974. Evidence for the neoformation of attapulgite in some soils of Iraq.
Inter. Cong. Soil. Sci. Trans. loth, Moscow, 7: 154-160.
Eswaran, H. and Bin, W.C., 1978a. A study of a deep weathering profile on granite in Peninsular
Malaysia: 111. Alteration of feldspars. Soil Sci. Amer. Jour. 42: 154-158.
Eswaran, H. and Bin, W.C., 1978b. A study of a deep weathering profile on granite in peninsular
Malaysia: I1 mineralogy of the clay, silt, and sand fractions. Soil. Sci. SOC.Amer., 42: 149-153.
Eswaran, H. and Heng, Y.Y.,1976. The weathering of biotite in a profile on gneiss in Malaysia.
Geoderma, 16: 9-20.
Eswaran, H. and Sys, C., 1972. Clay mineralogy of soils on ultrabasic rocks from Sabah. Borneo. Proc.
Inter. Clay Conf., Madrid, 215-226.
Etheridge, M.A., Paterson, M.S. and Hobbs, B.E., 1974. Experimentally produced preferred orientation
in synthetic mica aggregates. Contrib. Miner. Petrol., 44:235-294.
Eugster, H.P. and Hardie, L.A., 1978. Saline lakes. In: A. Lerman (Editor), Lakes. Springer-Verlag,
Berlin, 237-293.
Evans, G., 1975a. Intertidal flat sediments and their environments of deposition in the Wash. Geol. SOC.
London, Quart, Jour., 121: 209-245.
Evans, G., 1975b. Intertidal flat deposits of the Wash, western margin of the North Sea. In: R.N.
Ginsburg (Editor), Tidal Deposits. Springer-Verlag, 13-20.
Evans, L.J. and Adams, W.A., 1975. Chlorite and illite in some lower Paleozoic mudstones of Mid-Wales.
Clay Miner., 10: 387-397.
Evernden, J.F.. Curtis, G.H., Obradovich, J. and Kistler. R., 1961. On the evaluation of glauconite and
illite for dating sedimentary rocks by the potassium-argon method. Geochim. Cosmochim. Acta. 23:
78-99.
Faas, R.W. and Crocket, D.S.1983. Clay fabric development in a deep-sea core: site 515 Deep Sea
740

Drilling Project Leg 72, In: Barker. P.R.. Carlson. R.L.. and Johnson, D.A., ed., Initial Re-DSDP.
V.LXXI1. Wash. U.S. Gov. Print. Office, 519-535.
Fahey. J.J., Ross, M. and Axelrod. J.M., 1960. Loughlinite a new hydrous sodium magnesium silicate.
Amer. Miner.45. 270-28.
Fanning, D.S. and Keramidas, V.Z., 1977. Micas. In: R.C.Dinauer (Editor), Minerals in Soil Environ-
ments, Soil Sci. Soc.Amer.: Madison. 195-258.
Faust, G.T. and Fahey, J.J.. 1962. The serpentine-group minerals. U.S. Geol. Survey Prof. Pap. 384-A:
92p.
Faust. G.T., Hathaway. J.C. and Millot, G.. l9?9. Restudy of stevensite. Amer. Miner.. 44: 342-370.
Faust, G.T. and Murata. K.J.. 1953. Stevensite. redefined as a member of the montmorillonite group.
Amer. Miner., 38: 973-987.
Feuillet, J.P., 1976. James River estuary. Virginia: Control of clay mineral distribution by estuarine
circulation. Geol. Soc. Amer., Abs. with Prog. NE/SE Sec.. 169
Feuillet, J.P. and Fleischer, P.. 1980. Estuarine circulation: controlling factor of clay mineral distribution
in James River estuary. Virginia Jour. Sed. Petrol., 50: 267-279.
Fisher, G.W.. Pettijohn, .F.J.. Reed, J.C.Jr., Weaver, K.N.. 1970. Studies of Appalachian Geology;
central and southern. Interscience Pub.: 460p.
Fisher, M.J. and Jeans. C.V., 1982. Clay minerals stratigraphy in the Permo-Triassic red bed sequences of
BNOC 72/10-A. western approaches, and the South Devon Coast. Clay Minerals: 17. 79-89.
Fiskell. J.G.A. and Perkins, H.F., 1959. Lakeland Series: Norfolk Series, In: Rich, C.I.. Seatz. L.F.. and
Kunze, G.W., ed.. Certain properties of selected southeastern United States soils and mineralogical
procedures for their study. Rept. Coop. Res. Under Southern Regional Project S-14, 67-96.
Filch. F.J., Hooker. P.J., Miller. J.A. and Brereton, N.R., 1978. Glauconite dating of Palaeogene Eocene
rocks from East Kent and the time-scale of Palaeogene volcanism in The North Atlantic region. Jour.
Geol.Soc. Lond.. 135: 499-512.
Flawn. P.T., 1953. Petrographic classification of argillaceous. sedimentary. and low-grade metamorphic
rocks in subsurface. Amer. Assoc. Pet. Geol. Bull., 37: 560-565.
Flawn, P.T.. Goldstein. A. Jr.. King, P.B. and Weaver, C.E.. 1961. The Ouachita System. Texas Bureau
Econ. Geol., Pub. No. 6120, 401 p.
Fleischer. P. 1972a. Mineralogy and sedimentation history. Santa Barbara Basin. California. Jour. Sed.
Petrol., 42: 49-58.
Fleischer. P., 1972b. Sepiolite associated with Miocene diatomite Santa Cruz Basin. California. Amer.
Miner., 57: 903-913.
Flemal, R.C., 1967. Sedimentology of the Sespe Formation, Southwestern California. Ph.D. Thesis.
Princeton U.. 230p.
Flexor. J.M., de Oliveira, J.J., Rapaire. J.L.and Siefferman, G.. 1975. La degradation des illites en
montmorillonite dans I'alios de podzols tropicaux humo-ferrugineux du recancavo bahianais et du
Par& Cah.Orstom, sir. PCdol. 13: 41-48.
Folk. R.L.. 1960. Petrography and origin of the Tuscarora. Rose Hill and Keefer Formation. Lower and
Middle Silurian of eastern West Virginia. Jour. Sed. Petrol., 30. 1-58.
Folk. R.L., 1974. Petrology of sedimentary rocks. Austin, Texas. Hemphill's. 182p.
Follett, E.A.C., McHardy, W.J., Mitchell. B.D. and Smith, B.F.L.. 1965. Chemical dissolution techniques
in the study of soil clays, I. Clay Miner. 6: 23-34.
Fontes. J.C.. Fritz. P.. Gauthier, J. and Kulbicki, G., 1967. Mineraux argileux. elements-traces et
compositions isotopiqees ' n O / ' 6 0 et 13C/'2C dans les formation gypsifkres de L'Eockne Superieur
el de L'Oligockne de Cormeilles-en-Parisis. Bull. Centre Rech. Pau-SNPA, 1: 315-366.
Forman. S.A. and Brydon, J.E., 1965. Clay mineralogy of Canadian soils. Geol. Pedol. Eng. Studies,
Royal Soc. Canada Spec. Pub No. 3, 14-146.
Foscolos, A.E. and Kodama. H., 1974. Diagenesis of clay minerals from Lower Cretaceous shales o f
northeastern British Columbia. Clays Clay Miner.. 22: 319-335.
Foscolos. A.E. and Powell, T. G., 1978. Mineralogical and geochemical transformation of clays during
burial-diagenesis (catagenesis): relation to oil generation. Proc. Inter. Clay Conf.. Oxford. 261 -270.
Foscolos. A.E. and Powell, T.G., 1979. Categenesis in shales and occurrence o f authigenic clays in
sandstones, North Sabine H-49 well. Canadian Arctic Islands. Can. Jour. Earth Sci., 16: 1309-1314.
741

Foscolos, A.E., T.G. Powell and Gunthier, P.R. 1976. The use of clay minerals and inorganic and organic
geochemical indicators for evaluating the degree of diagenesis and oil generating potential of shales.
Geochim Cosmochem. Acta. 40: 953-966.
Foster, M.D., 1962. Interpretation of the Composition and a classification of the Chlorites. U.S. Geol.
Sur. Prof. Paper 414-A: 33p.
Foster, M.D., 1963. Interpretation of the Composition of Vermiculites and Hydrobiotites: Clays Clay
Miner., 10: 70-89.
Foth, H.D., 1984. Fundamentals of Soil Science. John Wiley and Sons, 435 p.
Foth, H.D. and Turk, L.M., 1972. Fundamentals of Soil Science. John Wiley and Sons, N.Y., 454 p.
Fournier, R.O., 1965. Montmorillonite pseudomorphic after plagioclase in a porpyry copper deposit.
Amer. Miner, 50: 771-777.
Frakes, L.A., 1979. Climates Throughout Geologic Time. Elsevier, Amsterdam, 310p.
Francis, E.H., 1961. Thin beds of graded kaolinized tuff and tuffaceous siltstones in the Carboniferous of
Fife. Great Britain Geol. Surv. Bull. 17, 191-215.
Frankel, L., 1977. Microorganisms induced weathering of biotite and hornblende grains in estuarine
sands. Jour. Sed. Petrol., 47: 849-854.
Frank-Kamonetzky, V.A., Kator, N.V., Carlo, E.A., and Klotchkova, G.N., 1971. Structural transforma-
tion of some clay minerals under pressure in hydrothermal conditions. Miner. Soc. Japan, Spec. Paper
1, 88-97.
Franks, P.C., 1966. Petrology and stratigraphy of the Kiowa and Dakota Formations (basal cretaceous).
north-central Kansas. Kansas U., PhD. dissert., 218p.
Franks, S.G. and Forester, R.W., 1984. Relationships among secondary porosity, pore-fluid chemistry
and carbon dioxide Texas Gulf, Coast. In: D.A. McDonald and R.C. Surdam (Editors), Clastic
Diagenesis, Amer. Assoc. Pet. Geol., Tulsa, 63-79.
Fraser, G.S., Harvey, R.D. and Baxter, J.W., 1973. Paleoenvironmental significance of corrensite in
Middle Mississippian rocks in southern Illinois. Bull. Geol. Soc. Amer, Abs with Prog. 15: 315.
Freas, D.H., 1962. Occurrence, mineralogy, and origin of the lower Golden Valley kaolinitic clay deposits
near Dickinson, North Dakota. Geol. SOC.Amer. Bull., 73: 1341-1364.
Freed, R.I., 1980. Shale mineralogy of the No. 1 Pleasant Bayou geothermal test well: a progress test. In:
Proc. Fourth Geopressured-Geothermal Energy Conf., Austin, Texas, University of Texas, 153-165.
Frey, M., 1969. A mixed layer paragonite/phengite of low-grade metamorphic origin Contr. Mineral.
Petrol., 24: 63-65.
Frey, M., 1970. The step from diagneesis to metamorpohism in pelitic rocks during Alpine orogenesis.
Sedimentology, 15: 261-279.
Frey, M. 1974. Alpine metamorphism of pelitic and marly rocks of the Central Alps. Schweiz. Miner.
Petrog. Mitteilungen, 54: 490-506.
Frey, M., 1978. Progressive low-grade metamorphism of a black shale formation, Central Swiss Alps,
with special reference to pyrophyllite and margarite bearing assemblages. Jour. Petrol., 19: 95-1 35.
Frey, M., 1987. Low Temperature Metamorphism. Blackie and Son. 288 p.
Frey, M. and Hunziker, J.C., 1973. Progressive niedriggradige metamorphase glaukonitfuhrender hori-
zonte in den helvetischen Alpen der Ostschweiz. Contr. Miner. Petrol., 39: 185-218.
Frey, M. and Niggli, E. 1971. Illit-kristallinitat, mineralfazien und inkohlungagrad. Schweiz. Mineral.
Petrogr. Mitt., 58: 229-234.
Frey, M., Teichmiiller, M., Teichmiiller, R., Mullis, J., Kiinzi, B., Breitschmid, A., Gruner. U. and
Schwizer, B., 1980. Very low- grade metamorphism in external parts of the Central Alps: illite
crystallinity, coal rank and fluid inclusion data. Eclogae Geol. Helv., 73: 173-203.
Frey, R.W. and Basan, P.B., 1978. Coastal Salt Marshes In: Davis, Jr., R.A., (Editor), Coastal
Sedimentary Environments, Springer-Verlag, N.Y., 101-170.
Frezon, S.E. and Schultz, L.G., 1961. Possible bentonite beds in Arkansas and Oklahoma. U.S. Geol.
Surv. Prof. Pop. 424C. 82-84.
Friend, P.F., 1966. Clay fractions and colours of some Devonian red beds in the Catskill Mountains,
U.S.A.. Q. Jour. Geol. Soc. Lond., 122: 273-92.
Fritz, S.J. and Mohr, D.W., 1984. Chemical alteration in the micro weathering environment within a
spheroidally-weathered anorthosite boulder. Geochim Cosmochim Acta, 48: 2527-2535.
742

Fry, J.C., Willman. H.B. and Glass, H.D., 1960. Gumbotil, accretion-gley and the weathering profile. 111.
State Geol. Sur. Circ. 295, 39p.
Frye. J.D.. Glass. H.D.. Leonard, A.B. and Coleman. D.D.. 1974. Caliche and clay mineral zonation of
Ogallala Formation, central-eastern New Mexico. New Mexico Bur. Mines Miner. Res, Cir. 144. 16p.
Fiichtbauer, H.. 1967. Influence of different types of diagenesis on sandstone porosity. Proc. World
Petroleum Cong., 71h, 12: 353-369.
Fiichtbauer, H. and Goldschmidt, H., 1959. Die tonminerale der Zechsteinformation. Beitr. Z. Miner. u.
Petrog., 6: 320-345.
Fiichtbauer, H. and Goldschmidt. H.. 1963. Beobachtungen zur Tonmineral-diagenese. Proc. Inter. Clay
Conf., 1st. Stockholm. 99-111.
Fiichtbauer. H. and Muller. G.. 1970. Sediment Petrologie (Teil I 1 - Sediment Sedimentgesteine). F..
Schweizerbart'sche Verlagsbuchhendlung. Stuttgart. 709p.
Fullager, P.D. and Bottino, M.L., 1969. Rubiduim - Strontuim age study of Middle Devonian
K-bentonite. Southeastern Geol., 10: 247-256.
Fysch, S.A.. Cashion, J.D. and Clark, P.E., 1983. Mossbauer effect studies of iron in kaolin I . structural
iron. Clays Clay Miner.. 31: 285-292.
Gac. J.Y.. Droubi. A,. Fritz, B. and Tardy, Y.. 1977. Geochemical behaviour of silica and magnesium
during the evaporation of waters in Chad. Chem. Geol.. 19: 215-228.
Galan. E.. Bell, J.M., Lalglesia, A. and Robertson, R.H.S.. 1975. The Ciceres palygorskite deposith.
Spain. Proc. Internat. Clay Conf. 1975, Applied Publishing Ltd.. Wilmette. 11.. 81-94.
Galan. E. and Castillo. A,, 1984. Sepiolite-palygorskite in Spanish Tertiary basins: genetical patterns in
Continental environments. In: A. Singer and E. Galan (Editor). Palygorskite-Sepiolite. Developments
in Sedimentology 37, Elsevier, Amsterdam. 87-124.
Galan, E. and Ferrero, A.. 1982. Palygorskite-sepiolite clays of Lebrija, southern Spain. Clays Clay
Miner., 30: 191-199.
Galliher, E.W., 1935. Geology of glauconite. Bull. Amer. Assoc. Petrol. Geol., 19: 1569-1601.
Galloway, W.E., 1974. Deposition and diagenetic alteration of sandstone in northeast Pacific arc-related
basins: Implication for graywacke genesis. Geol. Soc.Amer.Bull., 85: 379-390.
Garcia Palacios. M.C.. 1977. Mineralogia d e arcillas del Terciario de la cuenca alta del Tajo: un ejemplo
de secuencia magnesiana. Estudios geol., 33: 473-484.
Garcia Palacios, M.C., Lucas, J., De La PeRa, J.A. and Marfil. R. 1977. As Cuenca Triasica de la rama
Castellana da la Cordillera Iberica. 1. Petrografia y mineralogia. Cuadernos Geologia Ibtrica. 4:
341 -354.
Gardner, L.R., Kheoruenromne, I. and Chen, H.S., 1978. lsovolumetric geochemical investigation of a
buried granite sapiolite near Columbia, S.C., USA. Geochim. Cosmochim. Acta. 42: 417-424.
Garrels, R.M. and Howard. P.. 1959. Reaction of feldspar and mica with water at low temperature and
pressure. Proc. 6th Natl. Conf. Clays and Clay Minerals, Pergamon Press. 62-88.
Garrels. R.M. and Mackenzie, F.T., 1971. Evolution of Sedimentary Rocks. W . W. Norton and Co.. New
York, 397p.
Gast. R.G., 1969. Standard free energies of exchanges of alkali metal cations on Wyoming bentonite. Soil
Sci. Soc. Arner. Proc., 33: 37-41.
Gates. G.R., 1959. Clay mineral composition of borate deposits and associated strata at Boron.
California. Science, 130: 102.
Gaudette, H.E., 1965. lllite from Fond du Lac County. Wisconsin. Amer. Miner., 50: 411-417.
Gaultier. J.P. and Mamy. J.. 1977. A study concerning the factors influencing the structural evolution of
montmorillonite and and its reversibility. Proc. Eur. Clay Conf., 3rd, 70-72.
Ghent. E.D. and Miller. B.E., 1974. Zeolite and clay-carbonate assemblages in the Blairmore Group
(Creteceous) Southern Alberta Foothills. Canada. Contr. Miner. Petrol., 44: 31 3-329.
Gibbs. R.J.. 1967. The geochemistry of the Amazon River system: Part 1. The factors that control the
salinity and the com-position and concentrtation of the suspended solids. Geol. Soc. Amer. Bull.. 78:
1203- 1232.
Gibbs. R.J.. 1977. Clay mineral segregation in the marine environment. Jour. Sed. Petrol.. 47: 237-243.
Gibbs. R.J., 1983a. Coagulation rates of clay minerals and material sediments. Jour. Sed. Petrol.. 53:
1 193- 1203.
743

Gibbs, R.J., 1983b. Effect of natural organic coatings on the coagulation of particles. Env. Sci. Technol.,
17: 237-240.
Gibbs. R.J., 1984. Settling velocity. diameter and density for flocs of illite, kaolinite and montmorillonite.
Jour. Sed. Petrol., 55: 65-68.
Gibbs, R.J., Konwar, L. and Terchunian, A,, 1983. Size of flocs suspended in Delaware Bay. Can. Jour.
Fish. Aquatic Sci., 40: 102-104.
Giese. R.F. Jr.. 1971. Hydroxyl orientation in muscovite as indicated by electrostatic energy calculations.
Science, 172: 263-264.
Gieskes, J.M. and Lawrence. J.R., 1981. Alteration of volcanic matter in deep sea sediments: evidence
from the chemical composition of interstitial waters from deep sea drilling cores. Geochem. Cosmo-
chem. Acta. 45: 1687-1703.
Gile, L.H., Hawley, J.W. and Grossman, R.B., 1970. Distribution and genesis of soils and geomorphic
surfaces in a desert region of southern New Mexico. Soil Sci. SOC.Amer.Guide-book, soil-geomor-
phology field conferences, Aug. 21-22, 29-30, 1970, 156p.
Giletti. B.J., 1974. Studies in diffusion I: argon in phlogopite mica. In: A.W. Hofmann. B.J. Gilette. H.S.
Yoder, Jr. and R. A. Yund (Editors), Geochemical Transport and Kinetics, Carnegie Inst. Wash.
Publ. 634, 107-115.
Gilkes, R.J., 1978. On the clay mineralogy of upper Eocene and Oligocene sediments in the Hampshire
Basin. Proc. Geol. Ass., 89: 43-56.
Gilkes, R.J., Suddhiprakarn, A. and Armitage, T.M., 1980. Scanning electron microscope morphology of
deeply weathered granite. Clays Clay Miner., 28: 29-34.
Gill, J.R., Cobban, W.A. and Schultz, L.G., 1972. Stratigraphy and composition of the Sharon Springs
Member of the Pierre Shale in western Kansas. U.S. Geol. Sur. Prof. Paper 728, Sop.
Gill, W. D., Khalaf, F.I. and Massoud, M.S., 1977. Clay mineralogy as an index of the degree of
metamorphism of the carbonate and terrigenous rocks in the South Wales coalfield. Sedimentology.
24: 675-691.
Gillery, F.H., 1959. The x-ray study of synthetic Mg-AI serpentines and chlorites. Amer. Miner.. 44:
143-152.
Gipson, M., Jr., 1965. Application of the electron microscope to the study of particle orientation and
fissility in shale. Jour. Sed. Petrol., 35: 408-414.
Gjems, 0.. 1963. A swelling dioctahedral clay mineral of a vermiculite-smectite type in the weathering
horizons of podzols. Clay Miner., 5: 183-193.
Gjems, 0.. 1969. X-ray analyses of clay minerals in brown earth profiles on two different moraine types
in Hardanger, West Norway. Proc. Inter. Clay Conf. Tokyo, 553-558.
Glaccum, R.A., 1978. The mineralogical and elemental composition of mineral aerosols over the tropical
North Atlantic: the influence of Saharan dust. Thesis, U. Miami, Miami, Fla., 160p.
Glass, H. D., 1958. Clay mineralogy of Pennsylvanian sediments in southern Illinois. Clays Clay Miner.,
5: 227-241.
Glass, H.D., Potter, P.E., Siever. R., 1956. Clay mineralogy of some basal Pennsylvanian sandstones, clay
and shales. Amer. Assoc . Petrol. Geol. Bull., 40: 750-754.
Glasman, J.R., 1982. Alteration of andesite in wet, unstable soils of Oregon’s Western Cascades. Clays
Clay Miner., 30: 253-263.
Gluskoter, H.J., 1967. Clay minerals in Illinois coals. Jour. Sed. Petrol., 37: 205-214.
Goh, T.B. and Huang, P.M. 1986. Influence of citric and tannic acids on hydroxy-Al interlayering in
montmorillonite. Clays Clay Miner., 34: 37-44.
Goldberg, E.D. and Griffin, J.J.. 1964. Sedimentation rates of mineralogy in the South Atlantic. Jour.
Geophys. Res., 69: 4293-4309.
Goldberg, E.D. and Griffin, J.J., 1970. The sediments of the northern Indian Ocean. Deep-sea Res., 17:
513-537.
Goldich, S.S., Baadsgaard, Edwards, G., and Weaver, C.E., 1959. Investigations in the radioactive-dating
of sediments. Amer. Assoc. Pet. Geol. Bull., 43: 654-662.
Govett. G.J.S., 1966. Origin of banded iron formations. Geol. Soc. Am. Bull.. 77: 1191-1212.
Gorbunova, Z.N. and Shirshov, P.P., 1976. Clay sized minerals from cores of the southeast Pacific Ocean,
Deep Sea Drilling Pro-ject, Leg 35. In Hollister, C.D.. Craddock, C., et al., 1976 Initial Reports of the
Deep Sea Drilling Project, v. 35, Wash. (U.S. Gov. Printing Office), 479-487.
744

Gradusov. B.P., 1974. A tentative study of clay mineral distribution in soils of the world. Gerderma. 12:
49-55.
Grady. W.C.. 1984. Petrographic evidence for the time and mode of emplacement of minerals in coal.
Geol. Soc. Amer. Bull., Prog. with Abs, S.E. and N.C. Sections, 141.
Graff-Peterson, 1961. Lermineralogien: de limniske jura-sedimenter p i Bornholm. Kobenhavn, 149p.
Graham, S.A., Dickinson, W.R., Ingersoll, R.V., 1975. Himalayan-Bengal model for flysch dispersal in
the Appalachian Ouachita system. Geol. Soc. Amer. Bull., 86: 273-280.
Greenberg, S.A. and Price, E.W., 1957. The solubility of silica in solutions of electrolytes. Jour. Phys.
Chem., 61: 1539-1541.
Greene-Kelly, R., 1955. Dehydration of the montmorillonite minerals. Miner. Mag.. 30: 604-61 5.
Greenland, D.J., 1971. Interaction between humic and fulvic acids and clays. Soil Sci., 111: 34-41.
Griffin, G.M., 1962. Regional clay-mineral facies-products of weathering intensity and current distribu-
tion in the northeastern Gulf of Mexico. Geol. Soc. Amer. Bull., 73: 737-768.
Griffin, G.M., 1963. Occurrence of talc in clay fractions from beach sands of the Gulf of Mexico. Jour.
Sed. Petrol., 33: 231-233.
Griffin, G.M. and Ingram, R.L., 1955. Clay minerals of the Neuse River Estuary. Jour. Sed. Petrol.. 25:
194-200.
Griffin, G.M. and Parrott, B.S., 1964. Development of clay mineral zones during deltaic migration. Bull.
Amer. Assoc. Petrol. Geol., 48: 57-69.
Griffin, J.J. and Goldberg, E.D., 1963. Clay-mineral distributions in the Pacific Ocean. In: M.N. Hill
(Editor), The Sea. Interscience Pub., N.Y., 728-741.
Griffin, J.J. and Goldberg, E.D., 1969. Recent sediments of the Caribbean Sea. In: A.R. McBirney
(Editor), Tectonic Relations of Northern Central America and the Western Caribbean: The Bonacca
Expedition, Amer. Assoc. Petrol. Geol., Mem. 11, 258-268.
Griffin, J.J., Windom, H. and Goldberg, E.D., 1968. The distribution of clay minerals in the world ocean.
Deep Sea Res., 15: 433-459.
Grim, R. E., 1953. Clay Mineralogy, McGraw-Hill, New York, 384p.
Grim. R.E., 1968. Clay Mineralogy, McGraw-Hill Book Co.. New York, 596p.
Grim, R.E., Bradley, W.F. and White, W.A., 1957. Petrology of the Paleozoic shales of Illinois.
III.Geol.Survey, Rpt. of Invest. 203, 35p.
Grim, R.E., Bray, R.H. and Bradley, W.F., 1937. The mica in argillaceous sediments. Amer.Miner.. 22:
813-829.
Grim. R.E. and Giiven, N., 1978. Bentonites. Elsevier Sci.Pub. Co.. Amsterdam, 256p.
Grim, R.E. and Johns, W.D.. 1954. Clay mineral investigation of sediments in the northern Gulf of
Mexico. Clays Clay Miner., 2: 81-103.
Grim, R.E. and Kulbicki, G., 1961. Montmorillonite: High temperature reactions and classification.
Amer. Miner., 46: 1329-69.
Grimm. W.E., 1962. Idiomorphic quartz crystals as marker minerals for saliferous facies. Erdoel Kohle,
15: 880-887.
Groot. J.J. and Glass, H.D., 1960. Some aspects of the mineralogy of the northern Atlantic Coastal Plain.
Clays Clay Miner., 7: 271-284.
Gross. D.L., Lineback, J.A., Shimp. N.F. and White, W.A.. 1972. Composition of Pleistocene sediments
in southern Lake Michigan, U.S.A. 24th Int. Geol Cong., 8: 215-222.
Grout, F.F.. 1925. Relation of texture and composition of clays. Bull.Geol.Soc Amer., 36: 393-416.
Gruner, J.W., 1932. The crystal structure of kaolinite. Z. Krist., 83: 75-88.
Gruner, J.W., 1934. The structure of vermiculite and their collapse by dehydration. Amer. Miner.. 19:
557-575.
Gruner. J.W., 1936. The structure and chemical composition of Greenalite. Amer. Miner., 21: 449-55.
Gruner. J.W., 1944. The kaolinite structure of amesite and additional data on chlorites. Amer.Miner.. 29:
422-30.
Guggenheim, S.. Bailey, S.W., Eggleton. R.A. and Wilkes, P., 1982. Structural aspects of greenalite and
related minerals. Can. Miner., 20: 1-18.
Guthrie, J.M., Houseknecht, D.W. and Johns, W.D.. 1986. Relationships among vitrinite reflectance.
illite crystallinity, and organic geochemistry in Carboniferous strata, Ouachita Mountains, Oklahoma
and Arkansas. Amer. Assoc. Petrol. Geol. Bull., 70: 26-33.
745

Gutschick, R.C. and Sandberg. C.A., 1983. Mississippian continental margins of the conterminous
United States, In: Stanley, D.J. and Moore, G.T., (Editors), The Shelfbreak: Critical Interface on
Continental Margins, Soc. &on. Paleo. Mineral., Sp. Pub. No. 33, 79-96.
Giiven, N., 1970. Compositional and structural relationships between phengites and illites. Clays Clay
Miner., 18: 233-235.
Giiven, N., Hower, W.F. and Davies, D.K., 1980. Nature of authigenic illites in sandstone reservoirs.
Jour. Sed. Petrol., 50: 761-766.
Guven, N. and Kerr, P.F., 1966. Selected Great Basin playa clays. Amer. Miner., 51: 1056-1067.
Hagemann, F., 1957. On the petrography of the Silurian shales from Hadeland, Norway. Norsk
Geologisk Tidskrift, 37: 229-246.
Hagemann, F. and Spjeldnaes, N., 1955. The Middle Ordovician of the Oslo region, Norway. Norsk
Geologisk Tidskrift, 35: 29-52.
Hall, A.M., 1985. The clay mineralogy of the Lower San Andres Formation, Palo Duro Basin, Texas.
Master Thesis, Georgia Inst. Tech., Atlanta, Ga., 136p.
Hall, A. and Weaver, C.E., 1985. The clay mineralogy of the Lower San Andres Formation, Palo Duro
Basin, Texas. Battelle Mem. Inst., Report No. 15, 136p.
Hallam, A. and Sellwood, B. W., 1968. Origin of fuller’s earth in the Mesozoic of southern England.
Nature, 220: 1193-1195.
Hamilton, R., 1964. Microscopic studies on laterite formation. In: A. Jongerius (Editor), Soil Micromor-
phology, Elsevier Publ. Co., 269-276.
Hancock. N.J. and Taylor, A.M., 1978. Clay mineral diagenesis and oil migration in the Middle Jurassic
Brent Sand Formation. Jour. Geol. Soc. Lond., 135: 69-72.
Hann, H.H. Stumm, W., 1970. The role of coagulation in natural waters. Amer. Jour. Sci., 268: 354-368.
Harder, H., 1972. The role of magnesium in the formation of smectite minerals. Chem. Geol., 10: 31-39.
Harder, H., 1974. Mite mineral synthesis at surface temperatures. Chem.Geol., 14: 241 -253.
Harder, H., 1977. Clay mineral formation under lateritic weathering conditions. Clay Miner., 12:
281-288.
Harder, H., 1980. Syntheses of glauconite at surface temperature. Clays Clay Miner., 28: 217-222.
Harris, L.D. and Milici, R.C., 1977. Characteristics of thin-skinned type of deformatin in the southern
Appalachians, and potential hydrocarbon traps. U.S. Geol. Sur. Prof. Paper 1018, 40p.
Harris, W., 1976. Rb-Sr glauconite isochron, Maestrichtian Unite of Peedee Formation (Upper Creta-
ceous), North Carolina. Geology, 4: 761-762.
Harris, W. and Baum, G.E.. 1977. Foraminifera and Rb-Sr glauconite ages of a Paleocene Beaufort
Formation outcrop in North Carolina. Geol. Soc. Amer. Bull., 88: 869-872.
Harris, W. and Bottino, M.L., 1974. Rb-Sr study of Cretaceous lobate glauconite pellets, North Carolina.
Geol. Soc. Amer. Bull., 85: 1475-1478.
Harrison, J.L. and Droste, J.B., 1958. Division of Mississippian rocks in Indiana by clay mineral
variaton. Bull. Geol. SOC.Amer., 69, Abs., 1556.
Harrison, J.L. and Droste, J.B., 1960. Clay partings in gypsum deposits in southwestern Indiana. Clays
Clay Miner., 5: 195-199.
Harrison, S.C., 1971. The sediment and sedimentary processes of the Holocene tidal flat complex,
Delmarva Peninsula, Virginia. Johns Hopkins U., PhD. dissert., 54-153.
Hart, S.R. and Staudigel, H., 1979.0cean crust - sea water interaction: Sites 417 and 418. In Donnelly,
T., Francheteau, J., Bryan, W., Robinson, P., Flower, M., Salisbury, M., et al., Initial Reports of the
Deep Sea Drilling Project, 51, 52, 53 Part 2: Wash (US. Gov. Printing Office), 1169-1179.
Harvey, R.D. and Beck, C.W., 1960. Hydrothermal regularly interstratified chlorite-vermiculite and
tobermorite in alteration zones at Goldfield, Nevada. Clays Clay Miner., 9: 343-354.
Harvey, R.D., White, W.A., Cluff, R.M., Frost, J.K. and DuMontelle, P.B., 1977. Petrology of New
Albany Shale Group (Upper Devonian and Kinderhookian). In: G.L. Schott et al. (Editors), The
Illinois Basin, a preliminary report, Proc. First East. Gas Shale Symp., Morgantown, W. Va., Dept.
Energy. 328-354.
Harward, M.E., Carstea, D.D. and Sayegh, A.H., 1969. Properties of vermiculite and smectites:
Expansion and collapse. Clays Clay Miner., 16: 437-447.
Hassouba, H. and Shaw., H.F., 1980. The occurrance of palygorskite in Quaternary sediments of the
coastal plain of north-west Egypt. Clay Miner., 15: 77-83.
746

Hatcher. R.D.. Jr., 1978. Tectonics of the western Piedmont and Blue Ridge, Southern Appalachians:
Review and speculation. Amer. Jour. Sci., 278: 276-304.
Hathaway. J.C., 1955. Studies of some vermiculite-type clay minerals. Clays Clay. Miner.. 3: 74-86.
Hathaway, J.C., 1972. Regional clay mineral facies in estuaries and continental margin of the United
States east coast. Geol. SOC.Amer. Mem. 133, 293-316.
Hathaway, J.C., McFarlin, P.F. and Ross, D.A., 1970. Mineralogy and origin of sediments from drill
holes on the continental margin of Florida. U.S. Geol. Sur. Prof. Paper 581-E, 26p.
Hathaway. J.C. and Sachs, P.L., 1965. Sepiolite and clinoptilolite from the Middle Atlantic ridge. Amer.
Miner., 50: 852-867.
Hauf, P.L. and McKee, .E.D... 1979. The use of corrensite as an indicator of depositional environment in
the Supai Group of the Grand Canyon region, Arizona. Clays Clay Miner., 28th Conf.. abs.. 41.
Haven, D.S. and Morales-Alamo, R., 1968. Occurrence and transport of fecal pellets in suspension in a
tidal estuary. Sed. Geol.. 2: 141-151.
Haven. D.S. and Morales-Alamo. R., 1972. Biodeposition as a factor in sedimentation of fine suspended
solids in estuaries. Geol. Soc. Amer. Memoir 133, 124-130.
Hawkins. G.P.. Dodge, C.F. and Butler, J.C., 1974. Clay mineralogy of the Lewisville Member o f the
Cretaceous Woodbine Formation in the Arlington area, Tarrant County, Texas. Gulf Coast Assoc.
Geol. SOC..Trans., 24: 283-290.
Hawkins, D.B. and Roy, R.. 1963. Experimental hydrothermal studies on rock alteration and clay
mineral formation. Geochim. Cosmochim. Acta, 27: 1047-1054.
Hay. R.L. and Guldman, S.G.. 1986. Silicate diagenesis in sediments of Searles Lake, California. Clay
Miner. SOC.Prog. with Abs.. Jackson. MS, 15.
Hay, R.L. and Moiola, R.J., 1963. Authigenic silica minerals in Searles Lake, California. Sedimentology,
2: 312-332.
Hayes, J.B.. 1961. Clay mineralogy ofMississippian strata of southeast Iowa. Clays Clay Min. 10:
413-425.
Hayes, J.B., 1970. Polytypism of chlorite in sedimentary rocks. Clays Clay Miner., 18: 285-306.
Hayes, J.B., 1973. Clay petrology of mudstones, Leg 18, Deep Sea Drilling Project, in Kulm. L.D..
vonHuene, R. et al., 1973, Initial Reports of the Deep Sea Drilling Project, V. 18. Wash. (U.S. Gov.
Print. Office), 903-924.
Heath, G.R.. 1969. Mineralogy of Cenozoic deep-sea sediments from Equatorial Pacific Ocean. Geol.
Soc. Amer. Bull., 80: 1997-2018.
Heath, G.R. and Dymond, J.. 1977. Genesis and transformation of metalliferous sediments from the East
Pacific Rise, Bauer Deep. and Central Basin, northwest Nazca plate. Geol. SOC. Amer. Bull.. 88:
723-733.
Heath, G.R., Moore, T.C., Jr. and Roberts, G.L., 1974. Mineralogy of surface sediments from the
Panama Basin, eastern Equatorial Pacific. Jour. Geol., 82: 145-160.
Heezen, B.C., Nesterhoff, W.D., Oberlin, A. and Sabatier, G., 1965. Decouverte d’attapulgite dans les
sediments profonds du Golf d’Aden et de la Mer Rouge. C. r. hebd. Skanc. Acad. Sci., Paris, 260:
2819-2821.
Heim. D., 1957. Uber die mineralischen. nich karbonitischen Bestandteile des Cenoman und Turon der
mitteldeutschen Kreidemulden und ihre Verteilung. Heidelberg. Beitr.. 5: 302-330.
Hein, F.J. and Longstaffe, F.J., 1983.Geotechnical. sedimentological and mineralogical investigations in
Artic Fjords. In: J.P.M. Syvitsky and C.P. Blakeney (Editors). Sedimentology of Artic Fjords
Experiment: HU 82-031 Data Report, VI, Canadian Data Report of Hydrography and Ocean
Science No.12, 157p.
Hein, J.R. and Scholl, D.W., 1978. Diagenesis and distribution of late Cenozoic volcanic sediment in the
southern Bering Sea. Geol. SOC.Amer. Bull., 89: 197-210.
Hein. J.R., Yeh. , H.W. and Alexander, E., 1979. Origin of iron-rich montmorillonite from the manganese
nodule belt of the north equatorial Pacific. Clays Clay Miner, 27: 185-194.
Hekinian, R. and Hoffert, M., 1975. Rate of palagonization and manganese coating on basalt rocks from
the Rift Valley in the Atlantic Ocean near 36” 50”. Marine Geol., 19: 91-109.
Hekinian, R., Rosendahl, B.R., Cronan, D.S., Dmitriev. Y.. Fodor. R.V., Goll. R.M.. Hoffert. M..
Humphris, S.E., Mattey, D.P., Natland. J., Petersen. N., Roggenthen, W., Schrader. E.L., Srivastava,
747

R.K and Warren, N.. 1978. Hydrothermal deposits and associated basement rocks from the Galapa-
gos spreading center. Oceanol. Acta I: 373-482.
Heling, D., 1974. Diagenetic alteration of smectite in argillaceous sediments of the Rhinegraben (SW
Germany). Sedimentology, 21 : 463-472.
Heling, D.. 1978. Diagenesis of illite in argillaceous sediments of the Rhinegraben. Clay Miner.. 13:
211-220.
Heller-Kallai, L., Nathan, Y. and Zak. I., 1973. Clay mineralogy of Triassic sediments in southern Israel
and Sinai. Sedimentology, 20: 513-521.
Helmold. K.P. and Van de Kamp, P.C.. 1984. Diagenetic mineralogy and controls on albitization and
laumontite formation in Paleogene arkoses, Santa Ynez Mountains, California. In: D.A. McDonald
and R.C. Surdam, (Editors). Clastic Diagenesis Amer. Assoc. Pet. Geol. Mem. 37, Tulsa, 239-270.
Hendricks, S.B., 1939. Random structure of layer minerals as illustrated by cronstedite ( 2 F e 0 . Fe,O,.
SiO,.ZH,O). Possible iron content of kaolinite. Amer. Miner., 24: 529-39.
Hendricks, S.B. and Ross, C.S., 1941. Chemical composition and genesis of glauconite and celadonite.
Amer. Miner.. 26: 683-708.
Henley. J.J., 1959. Some mineralogical equilibria in the system K,O-AI,O,-SIO, -H,O. Amer. Jour.
Sci., 251: 241-270.
Henmi, T. and Wada, K.. 1976. Morphology and composition of allophane. Amer. Miner., 61: 379-390.
Henry, V.J.Jr. and Bader, R.G., 1961. Recent sedimentation and related oceanographic factors in the
west Mississippi Delta area. Texas A and M College. Dept. Oceanog. and Meteorol. Ref.61-6T, Proc..
111-129.
Henson, M.R., 1973. Clay minerals from the lower New Red Sandstone of South Devon. Proc. Geol.
ASSW., 84: 429-445.
Hepburn, J.C. and Rehmer, J.A., 1981. The digenetic to metamorphic transition in the Narragansett and
Norfolk Basins. Massachusetts and Rhode Island. In: J.C. Boothroyd and O.D. Hermes (Editors),
Guidebook to Geological Field Studies in Rhode Island and Adjacent Areas, 73rd New England
1ntercoll.Geol.Conf.. Dept. Geol., U. Rhode Island, 47-65.
Herbillon. A.J., Frankart, R. and Vielvoye, L., 1981. An occurrence of interstratified kaolinite-smectite in
a red-black soil toposequence. Clay Miner.. 16: 195-201.
Herbillon, A.J. and Makumbi, M.N.. 1975. Weathering of chlorite in a soil derived from a chloritoschist
under humid tropical conditions. Geoderma, 13: 89-104.
Herbillon, A.J., Mestdagh. M.M.. Vielvoye, L. and Derouane. E.G., 1976. Iron in kaolinite with special
reference to kaolinite from tropical soils. Clay Miner. 11: 201-219.
Heron, S.D. Jr., 1960. Clay minerals of the outcropping basal Cretaceous beds between the Cape Fear
River, North Carolina and Lynches River, South Carolina. Clays Clay Miner., 7: 148-161.
Heron, S.D., Jr., Robinson. G.C. and Johnson, H.S.. Jr., 1965. Clays and opal-bearing claystone of the
South Carolina Coastal Plain. Div. Geol., S.C. State Dev. Board, Bull. 31, 66p.
Heron, S.D., Jr. and Wheeler, W.H.. 1964. The Cretaceous Formations along the Cape Fear River, North
Carolina. Fifth Ann. Field Excursion, Atlantic Coastal Plain Geol.Assoc.. 55p.
HCroux, Y., Chagnon, A. and Bertrand, R., 1979. Compilation and correlation of major thermal
maturation indicators. Amer. Assoc. Petrol. Geol. Bull., 63: 2128-2144.
Herzog, L., Pinson, W.H.Jr. and Cormier, R.F., 1958. Sediment age determination by Rb/Sr analysis of
glauconite. Amer.Assoc. Pet. Geol. Bull., 42: 71 3-733.
Hey, M.H. 1954. A new review of the chlorites. Miner. Mag., 30: 277-92.
High, L.R. Jr. and Picard, M.D., 1965. Sedimentary petrology and origin of analcime-rich Pop0 Agie
Member, Chugwater (Triassic) Formation, west-central Wyoming. Jour. Sed. Petrol.. 35: 49-70.
Hiltabrand, F.. Ferrell. R.E. Jr.. and Billings, G.K. 1973. Experimental diagenesis of Gulf Coast
argillaceous sediments. Bull. Amer. Assoc. Petroleum Geol., 57: 338-348.
Hinch. H.H., 1978. The nature of shales and the dynamics of hydrocarbon expulsion in the Gulf Coast
Tertiary section. Amer. Assoc. Pet. Geol., Course Note Series No. 8. E-1 E-31.
Hinckley, D.N., 1963. Variability in “crystallinity” values among the kaolin deposits of the coastal plain
of Georgia and South Carolina. Clays. Clay Min., 13: 229-235.
Ho, C and Coleman, J.M.. 1969. Consolidation and cementation of recent sediments in the Atchafalaya
Basin. Geol. Soc. Amer. Bull., 80: 183-192.
748

Hobbs, B.E., Means, W.D. and Williams, P.E..1976. An Outline of Structural Geology Wiley, New York,
N.Y.. 571p.
Hodge, T., Turchenek. L.W. and Oades, J.M., 1984. Occurence of palygorskite in ground-water rendzinas
(Petrocalcic calciaquolls) in south-east South Australia. In: A. Singer and E. Galan (Editors),
Palygorskite-Sepiolite, Devlop. Sed. 37, Elsevier, 199-210.
Hoeve, J., Rawsthorn, K and Quirk, D., 1981. Clay minerals as a clue to uranium mineralization in the
Athabasca Basin. Uranium Symp., CIM Geol. Div.. Regina, Sask., 10-20
Hoeve, J.. Rawsthorn, K., and Quirk, D., 1981. Clay minerals as a clue to uranium mineralization in the
Athabasca Basin. Uranium Symp., Regina, Sask.
Hoffert, M., Perseil, A,, Hekinian. R., Choukroune, P.. Needham. H.D., Francheteau. J. and Le Pichon,
X. 1978. Hydrothermal deposits sampled by diving saucer in transform fault “A” near 37 N on the
Mid-Atlantic Ridge, FAMOUS area. Oceanol Acta I : 73-86.
Hoffert, M., Person, A,, Courtois, C., Karpoff, A.M. and Trauth, D., 1980. Sedimentology, mineralogy
and geochemistry of hydro-thermal deposits from holes 424, 424A. 424B and 424C (Galapagos
Spreading Center). Rosendahl, B.R., Hakinian, R.. et al.. Initial Reports of the Deep Sea Drilling
Project, V.54, Wash. ( U S . Gov. Printing Office), 339-376.
Hoffman, J. and Hower. J., 1979. Clay mineral assemblages as low grade metamorphic geothermometerb:
Application to the thrust faulted disturbed belt of Montana, U.S.A.. In: Scholle. P.A. and Schluger,
P.R. (Editors). Soc. Econ. Paleo. Miner., Special Pub. 26: 55-79.
Hofmann, U., 1956. Intercrystalline swilling. cation exchange. and anion exchange of minerals of the
montmorillonite group and of kaolinite. Clays Clay Miner. Nat. Acad. Sci-Nat. Res. Coun., Washing-
ton Pub. 456, 273-287.
Holdoway, K., 1978. Deposition of evaporites and red beds of the Nippewalla group. Permian, western
Kansas. Kansas Geol. Surv. Bull. 215, 43p.
Holeman, J.N.. 1968. The sediment yield of major rivers of the world. Water Resources Res.. 4: 737-747.
Hollister. C.D.. Ewing, J.I., et al.. 1972. Initial Reports o f the Deep Sea Drilling Proj.. XI. Washington.
1077~.
Holser. W.T.. 1979. Mineralogy of evaporites. In: Burns. R.G. (Editor), Marine Minerals, Miner. Soc.
Amer.. Short Course Notes V6, Wash., D.C., 211-286.
Honnorez, J.. 1981. The aging of the oceanic crust at low temperatures. In: Emiliani, C. (Editor), The
Sea, 7. Wiley-Interscience, N.Y., 525-587.
Hood. A, Gutjahr. C.C.M and Heacock, R.L., 1975. Organic metamorphism and the generation of
petroleum. Amer. Assoc. Petrol. Geol. Bull.. 59: 986-996.
Hooks, W.C. and Ingram, R.L., 1955. The clay minerals and the iron oxide minerals of the Triassic “red
beds” of the Durham Basin, North Carolina. Amer. Jour. Sci., 253: 19-25.
Hoskins, J.S.. Nielson, M.A. and Carthew, A.R., 1957. A study of clay mineralogy and particle size.
Australian Jour. Agr. Res.. 8: 45-74.
Hosterman. J.W., Patterson, S.H., Sweeney, J.W. and Hartwell, J.W., 1968. Clay. Mineral Resources of
the Appalachian Region, U.S. Geol. Survey Prof. Paper 580, 167-188.
Hosterman, J.W. and Whitlow, S.I., 1983. Clay mineralogy of Devonian shales in the Appalachian Basin.
U.S. Geol. Survey Prof. Paper 1298, 31p.
Howard, J.J.. 1981. Lithium and potassiium saturation of illite/smectite clays from interlaminated shales
and sandstones. Clays Clay Miner., 29: 136-142.
Hower, J., 1961. Some factors concerning the nature and origin of glauconite. Amer. Miner., 46:
3 1 3 - 334.
Hower. J. and Altaner, S.P., 1983. The petrologic significance of illite/smectite. Clay Miner. Soc. Prog.
with Abs.. Buffalo, 4 0 .
Hower, J., Eslinger, E.V., Hower, M.E. and Perry, G.A.. 1975. Mechanism of burial metamorphism of
argillaceous sediment: 1 . Mineralogical and chemical evidence. Geol. Soc. Amer. Bull.. 87: 727-757.
Hower, J., Hurley. P.M.. Pinson, W.H. and Fairbairn, H.W.. 1963. The dependence of K-Ar age on the
mineralogy of various particle size ranges in a shale. Geochem. Cosmochem. Acta. 27: 405-410.
Hower. J. and Mowatt, T.C., 1966. The mineralogy of illites and mixed-layer illite-montmorillonites.
Amer. Miner., 51: 825-854.
Hsu, P.H., 1977. Aluminum hydroxides and oxhydroxides. In: Dinauer. R.C. (Editor). Minerals in Soil
Environments, Soil Sci. Soc. Amer., Madison. 99-143.
749

Hsu, P.H. and Bates, T.F., 1964. Fixation of hydroxy-aluminium polymers by vermiculite: Soil Sci. SOC.
Amer. Proc., 28: 763-768.
Hsu, P.H. and Rich, C.I., 1960. Aluminum fixation in a synthetic cation exchanger. Soil Sci. SOC.Amer.
Proc., 24: 21-25.
Huang, W.H. and Keller, W.D., 1970. Dissolution of rock-forming silicate minerals in organic acid:
simulated first-stage weathering of fresh mineral surfaces. Amer. Miner., 55: 2076-2094.
Huang, W.H. and Keller, W.D., 1972. Geochemical mechanics for the dissolution, transport and
deposition of aluminum in the zone of weathering. Clays Clay Miner., 20: 69-74.
Huang, P.M. and Lee, S.Y., 1969. Effect of drainage on weathering transformation of mineral colloids of
some Canadian Prairie soils. Proc. Inter. Clay Conf., Tokyo, 541-552.
Huck, G. and Kanveil, J., 1955. Physikalisch-chemische Probleme der Inkohlung. Brennstoff-Chemie, 36:
1-11.
Huff, W.D. and Turkmenoglu, A.D., 1981. Chemical characteristics and origin of Ordovician K-be-
ntonites along the Cincinnati Arch. Clays Clay Miner., 29: 113-123.
Huggett, J.M., 1986. Back-scatter electron imaging in an SEM of phyllosilicates in sandstones and
mudstones. Clay Miner. SOC.Prog. with Abs., Baton Rouge, 62.
Huggins, C.W., Denny, M.V. and Shell, H.R., 1962. Properties of palygorskite, an abestiform mineral.
U S . Bur. Mines, Rpt. Invest. 6071, 8 p.
Hughes, R.E. and White, W.A., 1969. A flint clay in Sangamon County, Illinois. Proc. Internatl. Clay
Conf., Tokyo, 291-303.
Humphris, S.E., Nelson, W.G. and Thompson, R.N.,1980. Basalt weathering on the East Pacific Rise
and the Galapagos Spreading Center, Deep Sea Drilling Project Leg 54. Rosendahl, B.R., Hekinian,
R., et al., Initial Reports of the Deep Sea Drilling Project, 54: Wash. (US.Gov. Printing Office),
773-787.
Huneke, J.C., 1976. Diffusion artifacts in dating by stepwise thermal release of rare gases. Earth Planet.
Sci. Lett., 28: 407-417.
Hunt, C.B., 1972. Geology of Soils. W.H.Freeman and Co., San Francisco, 344p.
Hunter, K.A., 1983. On the estuarine mixing of dissolved substances in relation to colloid stability and
surface properties. Geochim. Cosmochim. Acta, 47: 467-473.
Hunter, K.A. and Liss, P.S., 1979. The surface charge of suspended particles in estuarine and coastal
waters. Nature, 282: 823-825.
Hunziker, J.C., 1979. Potassium-argon dating. In: Jager, E. and Hunziker, J.C. (Editors), Lectures in
Isotope Geology, Springer-Verlag, Berlin, 52-76.
Hurley, P.M., 1966. K-Ar dating of sediments. In: Schaeffer, O.A. and Zahringen, J. (Editors), Potassium
Argon Dating, Springer-Verlag, 134-150.
Hurley, P.M., Cornier, R.F., Hower, J., Fairbairn, H.W. and Pinson, W.H.Jr., 1960. Reliability of
glauconite for age measurement by K-Ar and Rb-Sr methods. Amer. Assoc. Pet. Geol. Bull., 44:
1793-1808.
Hurley, P.M., Heezen, B.C., Pinson, W.H. and Fairbairn, H.W., 1963. K-Ar age values in pelagic
sediments of the North Atlantic. Geochim. Cosmochim. Acta, 27: 393-399.
Hurst, A., 1981. The clay mineralogy of Jurassic shales from Brora, N.E. Scotland. Proc. Internat. Clay
Conf., Bologna, 677-684.
Hurst, A,, 1984. Mineral diagenesis in sandstones. Clay Miner. SOC.Prog. with Abs., Baton Rouge, 64.
Hurst, A. and Irwin, H., 1982. Geological modelling of clay diagenesis in sandstones. Clay Miner., 17:
5-22.
Hurst, V.J. and Schlee, J.S., 1962. Ocoee metasediments north central Georgia - southeast Tennessee.
Guidebook No. 3, Dept. Mines, Min. and Geol., Atlanta, Ga., 28p.
Hutchinson, G.E., 1957. A Treatise on Limnology, V.I. John Wiley and Sons, Inc., N.Y., 1015p.
Hyne, N.J., Laidig, L.W. and Cooper, W.A., 1979. Prodelta sedimen-tation on a lacustrine delta by clay
mineral flocculation. Jour. Sed. Petrol., 49: 1209-1216.
Ichikawa, A. and Shimoda, S., 1976. Tosudite from the Hokuno Mine, Hokuno Gifu Prefecture, Japan.
Clays Clay Miner., 24: 142-148.
Iglesia Fernandez, A. and Martin Vivaldi, J.L., 1978. A contribution to the synthesis of kaolinite. Proc.
Int. Clay Conf., Madrid, 173-198.
750

Iijima, A. and Hay, R.L., 1968. Analcime composition in tuffs of the Green River Formation of
Wyoming. Amer. Miner., 53: 184-200.
Iijima, A. and Matsumoto, R., 1982. Berthierine and chamosite in coal measures of Japan. Clays Clay
Miner., 30: 264-274.
Iijima, A. and Utada, M., 1971. Present-day zeolite diagenesis of the Neogene geosynclinal deposits in
the Niigata oil field, Japan. Adv. Chem. Series, 101, Molecular Sieve Zeolite, I: 342-349.
Ildefonse, P., Copin, E. and Velde, B., 1979. A soil vermiculite formed from a meta-gabbro, Loire-Atlan-
tique, France. Clay Miner., 14: 201-209.
Imam, M.B. and Shaw, H.F., 1985. The diagenesis of Neogene clastic sediments from the Bengal Basin,
Bangladesh. Jour. Sed. Petrol., 55: 665-671.
Ingram, R.L., 1953. Fissility of mudrocks. Geol. Soc. Amer. Bull., 64: 869-878.
Iniguez, A.M. and Zalba, P.E., 1979. Carboniferous clays of the Paganzo Group, La Rioja, Republica
Argentina. In: E.S. Belt and R.W. Macqueen (Editors), N e u v i h e Congrts International de Stratigra-
phie et de Gciologie du Carboniftre, 3, Washington and Champaign-Urbana, 674-678.
Ireland, B.J., Curtis, C.D. and Whitman, J.A., 1983. Compositional variation within some glauconites
and illites and implications for their stability and origin. Sedimentology, 30: 769-786.
Ismail, F.T., 1969. Role of ferrous iron oxidation in the alteration of biotite and its effect on the type of
clay minerals formed in soils of arid and humid regions. Amer. Miner., 54: 1460-1466.
Ismail, F.T., 1975. Biotite weathering and clay formation in and and humid regions, California. Soil Sci.,
109: 257-261.
Isphording, W.C., 1984. The clays of Yucatan, Mexico: A contrast in genesis. In: A. Singer and E. Galan
(Editors), Palygorskite-Sepiolite, Elsevier, 59-73.
Ivarson, K.C., Ross, G.J. and Miles, N.M., 1980. The microbiological formation of basic ferric sulfates 3.
Influence of clay minerals on crystallization. Can. Jour. Soil Sci., 60: 137-140.
Jackson, M.L., 1959. Frequency distribution of clay minerals in major great soil groups as related to the
factors of soil formation. Clays Clay Min., 6: 133-143.
Jackson, M.L., 1960. Structural role of hydronium in layer silicates during soil genesis. Inter. Cong. Soil
Sci. Trans., 7th, Madison, 11: 445-455.
Jackson, M.L., 1962. Interlayering of expansible layer silicates in soils by chemical weathering. Clays
Clay Miner., 11: 29-46.
Jackson, M.L., 1963. Interlayering of expansible layer silicates in soils by chemical weathering. Clays
Clay Miner., 11: 29-46.
Jackson, M.L, 1964. Chemical composition of soils. In: F.E. Bear (Editor), Chemistry of the Soil,
Reinbold Pub. Corp., New York, 71-141.
Jackson, M.L., 1965. Clay transformation in soil genesis during the Quaternary. Soil Sci., 99: 15-22.
Jackson, T.A. and Keller, W.D., 1970. A comparative study of the role of lichens and “inorganic”
processes in the chemical weathering of recent Hawaiian lava flows. Amer. Jour. Sci., 31: 80-86
Jackson, M.L. and Sridhar, K., 1974. Scanning electron microscopic and x-ray diffraction study of
natural weathering of phlogopite through vermiculite to sponite. Soil Sci. Amer. Proc., 38: 843-847.
Jacobs, M.B., 1974. Clay mineral changes in Antarctic deep-sea sediments and Cenozoic climatic events.
Jour. Sed. Petrol., 44: 1079-1086.
Jacobs, M.B. and Ewing, M., 1965. Mineralogy of particulate matter suspended in sea water. Science,
149: 179-180.
Jacobs, M.B. and Ewing, M., 1969. Mineral sources and transport in waters of the Gulf of Mexico and
Caribbean Sea. Science, 163: 805-809.
Jacobs, M.B., Thorndike, E.M. and Ewing, M., 1973. A comparison of suspended particulate matter from
nepheloid and clear water. Marine Geol., 14: 117-128.
Jagodzinski, H., 1949. Eindimensionale Ferlordnung in Kristallen und ihr Einfluss auf die Rontgenin-
terferenzen. I. Berechnung des Fehlordnungsgrades aus der Rontgenintensitaten. Acta Cristallogr., Z:
201-207.
Jakobsson, S. P. and Moore, J. G., 1986. Hydrothermal minerals and alteration rates of Surtsey volcano,
Iceland. Geol. Soc. Amer. Bull., 97: 648-659.
James, H. L., 1954. Sedimentary facies of iron-formations. Econ. Geol., 49: 235-283.
Jansa, L.F., Enos, P., Tucholke, B.E., Gradstein, F.M., and Sheridan, R. E., 1979. Mesozoic-Cenozoic
751

sedimentary formations of thc North American Basin; Western North Atlantic. In: M. Talwani, W.
Hay and W.B.F. Ryan, Deep Sea Drilling Results in the Atlantic Ocean: Continental Margins and
Paleoenvironment. Amer. Geoph. Union, Wash. D.C., 1-57.
Jansa. L.F., Gardner. J.V. and Dean, W.E., 1977. Mesozoic sequences of the central North Atlantic. In:
Y. Lancelot and E. Siebold et al., Initial Reports of the DSDP, 41, Wash. (U.S. Gov. Printing Office),
991 -1010.
Jansa, L.F. and Pe-Piper, G., 1985. Early Cretaceous volcanism on the northeastern American margin
and implications for plate tectonics. Geol. Soc. Amer. Bull., 96: 83-91.
Jarman, C.B., 1973. Clay mineralogy and sedimentary petrology of the Cretaceous Hudspeth Formation,
Mitchell, Oregon. Oregon State U., Ph.D. Thesis, 173 p.
Jeans, C.V., 1968. The origin of the montmorillonite of the European chalk with special reference to the
Lower Chalk of England. Clay Minerals, 7: 311-329.
Jeans, C.V., 1978. The origin of the Triassic clay asemblages of Europe with special reference to the
Keuper Marl and Raetic of parts of England. Philo. Trans. Royal SOC.London, 289: 549-639.
Jeans, C.V., Merriman, J.G.. Mitchell, J.G. and Bland, D.J., 1982. Volcanic clays in the Cretaceous
southern England and northern Ireland. Clay Miner., 17: 105-156.
Jenny, H., 1941. Factors of Soil Formation. McGraw-Hill, Inc., New York, 281 p.
Jepson, W.B. and Rowse, J.B., 1975. The composition of kaolinite - an electron microscope microprobe
study. Clays Clay Miner., 23: 310-317.
Johns, W.D. and Grim, R.E., 1958. Clay mineral composition of recent sediments from the Mississippi
River delta. Jour. Sed. Petrol., 28: 186-199.
Johns, W.D. and Kurzweil, H., 1979. Quantitative estimation of illite-smectite mixed phases formed
during burial diagenesis. TMPM Tschermaks Min. Petr. Mitt., 26: 203-215.
Johns, W.D. and Murray, H.H., 1959. Empirical crystallinity index for kaolinite. Geol. SOC.Amer. Abs.,
Pittsburg, 110.
Johnson, L.J., 1964. Occurrrence of regularly interstratified chlorite-vermiculite as a weathering product
of chlorite in a soil. Amer. Miner., 49: 556-572.
Johnson, L.J., 1970. Clay minerals in Pennsylvania soils relation to lithology of the parent rock and other
factors - 1. Clays Clay Miner., 18: 247-260.
Johnson, N.M., Likens, G.E., Bormann, F.H. and Pierce, R.S., 1968. Rate of chemical weathering of
silicate minerals in New Hampshire. Geochim. Cosmochim. Acta, 32: 531-545.
Johnson, R.G., 1974. Particulate matter at the sediment-water interface in coastal environments. Jour.
Mar. Res., 32: 313-330.
Jones, B.F. and Bowser, C.J., 1978. The mineralogy and related chemistry of lake sediments. In: A.
Lermian (Editor), Lakes. Springer-Verlag, 180-235.
Jones, B.F. and Spencer, R.J., 1985. Clay minerals of the Great Salt Lake Basin. Int. Clay Conf., 1985,
Denver, Abs., 114.
Jones, B.F. and Weir, A.H., 1983. Clay minerals of Lake Abert, an alkaline, saline lake. Clays Clay
Miner., 31: 161-172.
Jones, L.H., Milne, A.A. and Attiwel, P.M., 1964. Dioctahedral vermiculite and chlorite in highly
weathered red l o a m in Victoria, Australia. Soil Sci. Soc. Amer. Proc., 28: 108-113.
Jones, R.L. and Blatt, H., 1984. Mineral dispersal patterns in the Pierre Shale. Jour. Sed. Petrol., 54:
17-28.
Jordan, R.R. and Adams, J.K., 1962. Early Tertiary beontonite from the subsurface of central Delaware.
Geol. SOC.Amer. Bull., 173: 395-398.
Juster, T.C., Brown, P.E. and Bailey, S.W., 1987. NH,-bearing illite in very low grade metamorphic rocks
associated with coal, northeastern Pennsylvania. Amer. Miner., 72: 555-565.
Juteau, T., Noack, Y., Whitechurch, H. and Courtois, C., 1979. Mineralogy and geochemistry of
alteration products in Hole 417A and 417D basement samples (Deep Sea Drilling Project Leg 51). In:
T. Donnelly, J. Francheteau, W. Bryan, P. Robinson, M. Flower, M. Salisbury et al., Initial Reports
of the Deep Sea Drilling Project 51,52,53, Part 2: Wash. (US.Gov. Printing Office), 1273-1292.
Kaiser, W.R., 1984. Predicting reservoir quality and diagenetic history in the Frio Formation (Oligocene)
of Texas. In: D.A. McDonald and R.C. Surdam (Editors), Clastic Diagenesis. Amer. Assoc. Pet.
Geol., Tulsa, 195-215.
752

Kaiser, W.R. and Richman, D.L., 1981. Predicting diagenetic history and reservoir quality in the Frio
Formation of Brazoria County, Texas and Pleasant Bayou test wells. In: D. G . Bebout and A. L.
Bachman (Editors), Proc. Fifth Conf., Geopressured Geothermal Energy, 67-74.
Kane, J.E., 1967. Organic aggregates in surface waters of the Ligurian Sea. Limn. Oceanog.. 12: 287-294.
Kantor, W. and Schwertmann, U., 1974. Mineralogy and genesis of clays in red-black soil toposequences
on basic igneous rocks in Kenya. Jour. Soil. Sci.. 25: 67-78.
Kantorwicz. J.. 1984. The nature, origin and distribution of authigenic clay minerals from Middle
Jurassic Ravenscar and Brent Group sandstones. Clay Miner., 19: 359-375.
Karlsson, W.. Vollset, J., Bjorlykke, K. and Jorgensen, P.. 1979. Changes in mineralogical composition of
Tertiary sediments from North Sea Wells. Developments in Sedimentology 27. Proc. Internat. Clay
Conf., Oxford 1978, Elsevier, 281-290.
Karpova, G.V., 1969. Clay mineral post-sedimentary ranks in terrigenous rocks. Sedimentology. 13:
5-20.
Kanveil, J., 1956. Die metamorphose der kohlan vom standpunkt der physikalischen chemie. Zeitschr.
deutsch. geol. Ges., Bd 107.
Kastner. M., 1976. Diagenesis of basal sediments and basalts of Sites 322 and 323. Leg 25, Bellingshau-
sen Abyssal Plain. In: C.D. Hollister, C. Cradddock et al.. Initial Reprots of the Deep Sea Drilling
Project, 35. Wash. ( U S . Gov. Printing Office), 513-519.
Kastner, M., 1981. Authigenic silicates in deep-sea sediments: Formation and diagenesis. In: Emiliani, C.
(Editor). The Sea 5 , John Wiley&Sons, 915-980.
Kastner, M., 1982. Evidence for two distinct hydrothermal systems in the Guaymas Basin. In: J.R.
Curray and D.G. Moore et al., Initial Reports DSDP, 64, Pt. 2; Wash. ( U S . Govt. Printing Office).
1143-1 157.
Kautz, K. and Porada, H.. 1976. Sepiolite formation in a pan of the Kalahari, South West Africa. N . Jb.
Miner. Mh., 12: 545-559.
Kay, G.M., 1935. Distribution of Ordovician altered volcanic materials and related clays. Geol. Soc.
Amer. Bull., 46: 225-244.
Keene, J.B. and Kastner, M., 1974. Clays and formation of deep-sea chert. Nature. 249: 754-755.
Keller. G.H., Lambert, D.N. and Bennett, R.H., 1979. Geotechnical properties of Continental Slope
deposits-Cape Halteras T o Hydrographer Canyon. In: Doyle, L.J. and Pilkey, O.H. (Editor). Geology
of Continental Slopes. SOC.Econ. Paleo. Miner. Spec. Paper No. 27, 131-151.
Keller. W.D.. 1958. Glauconitic mica in the Morrison Formation in Colorado. Clays Clay Miner.. 5 :
120-128.
Keller. W.D.. 1962a. The Principles of Chemical Weathering. Lucas Brothers Publishers. Columbia. Mo.,
111 p.
Keller, W.D., 1962b. Clay minerals in the Morrison Formation of the Colorado Plateau. U.S. Geol. Sur.
Bull.. 1150. 90 p.
Keller. W.D., 1964. The origin of high-alumina clay minerals - a review. Clays Clay Miner., 12: 129-151.
Keller, W.D.. 1968. Flint clay and flint-clay facies. Clays Clay Miner., 16: 113-128.
Keller, W.D.. 1977. Scan electron micrographs of kaolins collected from diverse environments of origin -
IV. Georgia kaolin and kaolinizing source rocks. Clays Clay Miner., 25: 311-345.
Keller. W.D.. 1978a. Progress and problems in rock weathering related to stone decay. In: E. M. Winkler
(Editor), Engineering Geology Case Histories Number 11. Geol. Soc. Amer.. 37-46.
Keller. W.D.. 1978b. Kaolinization of feldspars as displayed in scanning electron micrographs. Geol.. 6:
184-188.
Keller, W.D., Hanson, R.F., Huang. W.H. and Cervantes, A,. 1971. Sequential active alteration of
rhyolitic volcanic rock to endellite and a precursor of it at a spring in Michorcan. Mexico. Clays Clay
Miner., 19: 121-127.
Keller, W.D., Pickett. E.E. and Reesman, A.L.. 1966. Elevated dehydroxylation temperature o f the
Keokuk geode kaolinite - a possible reference mineral. Proc. Internat. Clay Conf., Jerusalem. 1:
75-85.
Keller, W.D., Reynolds, R.C., and Atsuyuki, I.. 1986. Morphology of clay minerals in the smectite-to-il-
lite conversion servies. Clays Clay Miner.. 34: 187-197.
Keller. W.D.. Westcott. J.F., and Bledsoe. A. 0.. 1953. The origin of Missouri fire clays. Clays Clay
Miner.. 2: 7-46.
753

Keller, W.D., Westcott, J.F., and Bledsoe. A.O., 1954. The origin of Missouri fire clays. Clays Clay
Miner. Natl. Acad. Sci.-Natl. Res. Coun. Pub. 327, 7-76.
Kelley, J. T., 1983. Composition and origin of the inorganic fraction of southern New Jersey coastal mud
deposits. Geol. Soc.Amer. Bull., 94: 689-699.
Kelts, K., 1982. Petrology of hydrothermally metamorphosed sediments at Deep Sea Drilling Site 477,
southern Guaymas Basin Rift, Gulf of California. In: J. R. Curray and D. G. Moore el al., Init.
Repts. DSDP, 64, Pt. 2: Wash. (U.S. Govt. Printing Office), 1123-1136.
Kempe, D.R.C., 1974. The petrology of the basalts, leg 26. In: T.A. Davies, B.P. Luyendyk et al., Initial
Reports of the Deep Sea Drilling Project 26, Wash. (U.S. Gov. Printing Office), 465-503.
Kennedy, V.C., 1964. Sediment transported by Georgia streams. U.S.G.S. Water Supply Paper 1668, 54 p.
Kennedy, V.C., 1965. Mineralogy and cation-exchange capacity of sediments from selected streams. U.S.
Geol. Survey Prof. Paper 433-D, 28 p.
Kent, D.B. and Kastner, M., 1985. Mg2+ removal in the system Mg2+-amorphous SO,-H,O by
absorption and Mg-hydroxysilicate precipitation. Geochim. Cosmochim. Acta. 49: 1123-1 136.
Keppens, E. and Pasteels, P., 1982. A comparison of rubidium - strontium and potassium - argon
apparent ages on glauconies. In: G.S. Odin (Editor), Numerical Dating in Stratigraphy. John Wiley
and Sons, 225-244.
Kerr, P. F., 1950. Analytical data on reference clay materials. Preliminary report No. 7. Reference Clay
Min. Amer. Petroleum Inst., Res. Prop. 49, Columbia U., NY., 160p.
Kharaka, Y.K., Hull, R.W. and Carothers, W.D., 1985. Water-rock interactions in sedimenatary basins.
In: Relationship of Organic Matter and Mineral Diagenesis, Short Course No. 17, Soc. Econ. Paleo.
Miner., Tulsa, 79-176.
Khoury, H.N., Eberl, D.D. and Jones, B.F., 1982. Origin of magnesium clays from the Amargosa Desert,
Nevada. Clays Clay Miner., 30: 327-336.
Kiersch, G.A. and Keller. W.D., 1955. Bleaching clay deposits, Sanders - Defiance Plateau District,
Navajo County, Arizona. Econ. Geol., 50: 469-494.
Kimbara, K., 1975. Regularly interstratified clay minerals of chlorite and saponite (“corrensite”) in the
Miocene Green Tuff Formation in Japan. Bull. Geol. Surv. Jap., 26: 669-679.
Kimberley, M.M., 1978. Paleoenvironmental classification of iron formations. Econ. Geol., 73: 215-229.
Kirkland, D.L. and Hajek, B.F.. 1972. Formula derivation of Al-interlayered vermiculite in selected soil
clays. Soil Sci., 114: 317-322.
Kisch, H.J., 1966. Zeolite facies and regional rank of bituminous coal. Geological Mag., 103: 414-422.
Kisch, H.J., 1968a. Coal-rank and burial-metamorphic mineral facies. Organic Geochem., Pergamon
Press, 407-425.
Kisch, H.J., 1968b. Coal rank and lowest-grade regional metamorphism in the southern Bowen Basin,
Queensland, Australia. Geologie En Mijnbouw, 47: 28-36.
Kisch, H.J., 1974. Anthracite and meta-anthracite coal ranks associated with “anchimetamorphism” and
“ very-low-stage” metamorphism. Koninkl. Nederl. Akadem. Wetenschappen Proceed. Series B, 77:

81 -1 18.
Kisch, H.J., 1980a. Illite crystallinity and coal rank associated with lowest-grade metamorphism of the
Taveyanne greywacke in the Helvetic zone of the Swiss Alps. Eclogae Ged. Helv., 73: 753-777.
Kisch, H.J., 1980b. Incipient metamorphism of Cambro-Silurian clastic rocks from the Jamtland
Supergroup, central Scandinavian Caledonides, western Sweden: illite crystallinity and ‘‘ vitrinite”
reflectance. Jour. Geol. Soc. London, 137: 271-288.
Kisch, H.J., 1983. Mineralogy and petrology of burial diagenesis (burial metamorphism) and incipient
metamorphism in clastic rocks. In: G. Larsen and G.V. Chilingar (Editors), Diagenesis in Sediments
and Sedimentary Rocks, 2. Elsevier, 289-494.
Kisch, H.J., 1987. Correlation between indicators of very low-grade metamorphism, In: Frey, M.
(Editor). Low Temperature Metamorphism. Blackie, 227-300.
Kitagawa, Y., 1971. The “unite particle” of allophane. Amer. Miner., 56: 465-475.
Klein, G. DeV., 1962. Sedimentary structures in the Keuper Marl (Upper Triassic). Geol. Mag. XCIX:
137-144.
Klimentidis, R.E. and Mackinnon, I.D.R., 1986. High-resolution imaging of ordered mixed-layer clays.
Clays Clay Miner., 34: 155-164.
754

Klingebiel, A. and Latouche. C., 1964. Aspects gknkraux d e la skdimentation argileuse dans les faciks
littoraux du P a l t o g h e Nord-Aquitain. Develop. Sedimentol., 1: 200-206.
Knebel, H.J., Conomos, T.J. and Commeau, J.A., 1977. Clay-mineral variability in the suspended
sediments of the San Francisco Bay system, California. Jour. Sed. Petrol., 47: 229-236.
Knebel, H.J., Kelly, J.C. and Whetten, J.T., 1968. Clay minerals of the Columbia River: a qualitative.
quantitative, and statistical evaluation. Jour. Sed. Petrol., 38: 600-611.
Knox, R.W.OB., 1982. Clay mineral trends in cored Lower and Middle Jurassic sediments of the
Winterborne k n g s t o n Borehole, Dorset. In: G.H. Rhys, G.K. Lott and M.A. Calver (Editors). the
Winteborne Kingston Borehole, Dorset, England. Rep. Inst. Geol. Sci., 81 : 91 -96.
Knox, R.W. OB., 1985. Stratigraphic significance of volcanic ash in Paleocene and Eocene sediments at
Sites 549 and 550. In: P.C. Graciansky and C.W. de Poag et al. (Editors). Init. Reprts. DSDP. 80:
Wash. (U.S. Govt. Printing Office), 845-849.
Kodama, H., 1966. The nature of the component layers of rectorite. Amer. Miner., 15: 1035-1055.
Kodama. H., and Brydon, J.E., 1968. A study of clay minerals in podzol soils in New Brunswick. eastern
Canada. Clay Miner., 7: 295-309.
Kodama, H. and Dean, R.S., 1980. Mite from Eldorado, Saskatchewan. Can. Miner., 18: 109-118.
Kodama, H. and Schnitzer, M., 1969. Thermal analysis of a fulvic acid-montmorillonite complex. Proc.
Inter. Clay Conf., Tokyo, 1: 765-774.
Kodama. H., Schnitzer, M. and Jaakkimainen, M., 1983. Chlorite and biotite weathering by fulvic acid
solutions in closed and open systems. Can. Jour. Soil. Sci.. 63: 619-629.
Kohyama, N., Fukushima, K. and Fukama, A,, 1978. Observation of the hydrated form of tubular
halloysite by an electron microscope equipped with an environmental cell. Clays Clay Miner.. 26:
25-40.
Kolla, V., Kostecki, J.A., Robinson, F. and Biscaye, P.E., 1981. Distribution of clay minerals and quartz
in surface sediments of the Arabian Sea. Jour. Sed. Petrol., 51: 563-569.
Kolla, V., Nadler, L. and Bonatti. E., 1980. Clay mineral distributions in surface sediments of the
Philippine Sea. Oceanol. Acta, 3: 245-250.
Konta, J., 1985. Mineralogy and chemical maturity of suspended matter in major rivers sampled under
the SCOPE/UNEP Project. SCOPE/UNEP Sonderband Heft 58, Hamburg, 569-592.
Konta. J., Borovec, Z., Srbmek, J. and Tolar, VI., 1970. Changes of primary biotite and muscovite during
kaolinization of granite, Carlsbad Area, Czechoslovakia. Proc. Conf. Clay Miner. Petrol., 5th. Praha,
27-43.
Kopp, O.C. and Fallis, J.M., 1974. Corrensite in the Willington Formation, Lyons, Kansas. Amer.
Miner., 59: 623-624.
Kossovskaya, A.G., 1969. Specific features of the alteration of clay minerals under different facies-climatic
conditions. Internatl. Clay Conf.. Tokyo, 1: 339-347.
Kossovskaya, A.G. and Drits, V.A., 1970. The variability of micaceous minerals in sedimentary rocks.
Sedimentology 15: 83-101.
Kossovskaya, A.G., Gushchina, E.B., Drits, V.A., Dmitrik, A.L., Lomova, O.S., and Serebrennikova,
N.D., 1975. Mineralogy and origin of the Mesozoic-Cenozoic deposits of the Atlantic Ocean based on
data from cruise 2 of the Glomar Challenger. Lithol. Miner. Resources, 6: 675-694.
Kossovskaya, A.G. and Shutov, V.D., 1963. Facies of regional epi- and metagenesis, Izy. Abad. Nauk.
S.S.S.R., Ser. Geol. 28(7); 3-18. (Translated in: Intern. Geol. Res., U.S.A., 7: 1157-1167.)
Kossovskaya, A.G., Shutov, V.D. and Alexandrova, V.A., 1964. Dependence of the mineral composiiton
of the clays in the coal-bearing formations on the sedimentary conditions. Cong. Inter. Strat. Geol.
Carbonifere, 5th. Paris 1963, Compt. Rend., 2: 515-529.
Kraeuter, J.N., 1976. Biodeposition by salt-marsh invertebrates. Mar. Biol., 35: 215-223.
Kranck, K., 1973. Flocculation of suspended sediment in the sea. Nature, 246: 348-350.
Kranck, K., 1975. Sediment deposition from flocculated suspensions. Sedimentology, 22: 1 1 1-123.
Kraus, I. and Zuberec, J., 1976. Bentonites and kaolins at the SW margin of the KremnickC Pohorie Mts.
in the neovolcanic region of central Slovakia. Seventh Conf. Clay Miner. Petrol., Karlovy Vary -
439-449.
Krauskopf, K.B., 1967. Introduction to Geochemistry. McGraw-Hill, New York. 721 p.
Kristmannsdottir, H., 1978. Alteration of basaltic rocks by hydrothermal activity at 100-300°C. Proc.
Internatl. Clay Conf., Oxford, 359-367.
755

Krumm, H., 1969. A scheme of clay mineral stability in sediments. based on clay mineral distribution in
Triassic sediments of Europe. Proc. internatl. Clay Conf.. Tokyo, 1: 313-324.
Krylov, A.Ya. and Logvinenko. N.V.. 1979. Radiometric age of glauconite in Holocene Ocean sediments.
Doklady Akad. Nauk. SSSR, 249: 158-160.
Krylov. A.Ya. and Silin, Yu.1.. 1963. Argon dating in marine geology and paleogeography. In: A.P.
Vinogradov (Editor). Chemistry of the Earth's Crust, 1: 413-424.
Kuhler, B., 1964. Les argiles, indicateurs de mttamorphisme. Rev. Inst. Franc. Petrole, 19: 1093-1112.
Kuhler, B., 1967. La cristallinite de Mite et les zones tout h fait suptrieures du mttmorphisme. In: Etages
Tectoniques, A ,la Baconnitre, Neuchitel, Suisse, 105-121.
Kubler, B., 1968. Evaluation quantiative du metmorphisme par la cristallinite de l'illite. Bull. Centre.
Rech. Pau-SNPA, 2: 385-397.
Kuhler, B. 1970. Crystallinity of illite. Detection of metamorphism in some frontal parts of the Alps.
Fortschr. Miner. 47 (Beih. 1): 39-40.
Kuhler, B. 1973. La corrensite, indicateur possible de milieux de sedimentation et du degri d e
transformation d'un sediment. Bull. Centre Rech. Pau-S.N.P.A. 7: 543-556.
Kubler, B., Pittion, J.L., Htroux, Y., Charollais. J. and Weidmann, M.. 1979. Sur le pouvoir rtflecteur d e
la vitrinite dans quelques roches du Jura, de la Molasse et des Nappes prealpines. helvetiques et
penniques. Eclogae Geol. Helv. 72: 347-373.
Kulbicki, G., 1953. Constitution et gentse du sediments argileux si derolithiques et lacustres du Nord d e
I'Aquitaine. These Sci., Toulouse et Sci. Terre, 4: 5-101.
Kunze, G.W., Knowles, L.L. and Kitano, Y., 1968. The distribution and mineralogy of clay minerals in
the Taku Estuary of southeastern Alaska. Marine. Geol.. 6: 439-448.
Kunze. G.W. and Templin, E.H., 1956. Houston black clay, the type Grumusol: 11. Mineralogical and
chemical characterization. Soil Sci. Soc. Proc.. 20: 91 -96.
Kurnosov, V.B., Kholodkevich, I.V., Ludmila. P., Kokorina. L.P., Nicolai, V., Kotov. N.V.. Oleg. V. and
Chudaev, B., 1981. The origin of clay minerals in the oceanic crust revealed by natural and
experimental data. Proc. Int. Clay Conf., Bologna, Pavia, 547-556.
Lagache, M., Yart, J.W.. and Sahatier, 1963. On the formation of kaolin by alteration of albite at 200°C
in the presence of CO,. Acad. Sci. Paris Compt. Rend., 256: 2501-2503.
Lagaly, G. and Weiss, A., 1975. The layer charge of smectitic layer silicates. Proc. Inter. Clay Conf.,
Mexico City, 157-172.
Lahann, R.W. and Roberson, H.E., 1980. Dissolution of silica from montmorillonite: effect of solution
chemistry. Geochim. Cosmochim Acta, 44: 1937-1943.
Lahav. N.. Shani. U. and Shahtai. J.. 1978. Cross-linked smectites. 1. synthesis and properties of
hydroxy-aluminum-montmorillonite. Clays Clay Miner., 26: 107-1 15.
La Iglesia, A. and Galan, E., 1975. Halloysite-kaolinite transformation at room temperature. Clays Clay
Miner., 23: 109-113.
La Iglesia, A. and Martin Vivaldi, J.L., 1973. A contribution to the synthesis of kaolinite. Proc. Inter.
Clay Conf., Madrid, 173-185.
Lambert-Aikhionhare, D.O. and Shaw, H.F.. 1982. Significance of clays in the petroleum geology of the
Niger Delta. Clay Minerals. 17: 91-103.
Lancelot, Y.,Hathaway, J.C., and Hollister, C.D.. 1972. Lithology of sediments from the western North
Atlantic, Leg XI, Deep Sea Drilling Project. In: C.D. Hollister. J.I. Ewing et al., (Editors), Initial
Reports DSDP, V. XI. Wash. ( U S Gov. Printing Office). 901-950.
Land, L.S. and Dutton, S.P.. 1978. Cementation of a Pennsylvanian deltaic sandstone isotopic data. Jour.
Sed. Petrol., 48: 1167-1176.
Laplante, R.E., 1972. Petroleum generation in Gulf Coast Tertiary sediments. Amer. Assoc. Petrol. Geol..
Ahs. 56: 635.
Lashkov, E.M.. Lashkova, L.N. and Kotovich. V.A.. 1970. Diffractometric studies of a clay fraction in
Ordovician deposits in Silurian boundary layers in the southern Baltic Sea region. Liet. TSR Aukst.
Mokyklu Mokslo.. Geogr. Geol., 7: 207-212.
Latouche, C.. 1978. Clay minerals as indicators of the Cenozoic evolution of the North Atlantic Ocean.
Proc. Inter. Clay Conf., Oxford, 271-279.
Latouche, C. and Viguier. C.. 1976. L'evolution du cortege des mineraux argileux dans la sedimentation
756

marine Neogene du bassin occidental du Guadalquivir (Espagne du S.O.). Acta Geol. HispLnica. 9:
8-14.
Lawrence, J.R., 1979. Temperatures of formation of calcite in the basalts from Deep Sea Drilling Project
Holes 417A and 417D. In: T. Donnelly, J. Francheteau, W. Bryan, P. Robinson, M. Flower and M.
Salisbury et al., Initial Reports of the Deep Sea Drilling Project. 51, 52, 52 Part 2: Wash. (U.S. Gov.
Printing Office), 1183-1184.
Leclue, G.G., 1962. Distribution of clay minerals in the Paleozoic rocks of southwestern Montana. U.
Indiana, M.S. Thesis, 19 p.
Ledbetter, M.T., 1985. Tephrochronology of marine tephra adjacent to Central America. Geol. Soc.
Amer. Bull., 96: 77-82.
Lee, J.H., Ahn, J.O. and Peacor, D.R., 1985. Textures in layered silicates: progressive changes through
diagenesis and low-temperature metamorphism. Jour. Sed. Petrol., 55: 532-540.
Lee, J.H., Peacor, D.R., Lewis, D.D. and Wintsch, R.P., 1984. Chlorite-illite/smectite interlayered and
interstratified crystals: a TEM/STEM study. Contrib. Miner. Petrol., 88: 372-385.
Lee, J.H., Peacor, D.R., Lewis, D.D. and Wintsch, R.P., 1986. Evidence for syntectonic crystallization for
the mudstone to slate transition at Lehigh Gap, Pennsylvania, U.S.A. Jour. Stru. Geol., 8: 767-780.
Lee, M., Aronson, J.L. and Savin, S.M., 1984. Diagenesis of the Permian Rotliegendes Sandstone:
petrologic, oxygen isotopes. Clay Miner. SOC.Prog. with Abs., Baton Rouge. 77.
Lee, M.J. and Chaudhuri, S., 1976. Clay mineral studies of the Lower Permian Havensville Shale in
Kansas and Oklahoma. Clays Clay Miner., 24: 239-245.
Lee, S.Y., Jackson, M.L. and Brown, J.L., 1975. Micaceous occlusions in kaolinite observed by
ultramicrotomy and high resolution electron microscopy. Clays Clay Miner., 23: 125-129.
Lebauer, L.R., 1964. Petrology of the middle Cambrian Wolsey Shale of Southwestern Montana. Jour.
Sed. Petrol., 34: 503-511.
Lelong, F., 1969. Nature et genhse des produits d’altiration de roches cristallines sous climat tropical
humide (Guyane franpise). Sciences de la Terre, Mimoire 14, 188 p.
Leonard, R.A. and Weed, S.B., 1967. Influence of exchange ions on the b-dimension of dioctahedral
vermiculite. Clays Clay Miner., 15: 149-161.
Leopold, L.B., Holman, M.G. and Miller, J.P., 1964. Fluvial Processes in Geomorphology. H.H.
Freeman, San Francisco, 522 p.
Lerman, A,, Mackenzie, F.T. and Bricker, O.P., 1975. Rates of dissolution of aluminosilicates in sea
water. Earth Planet. Sci-Lett., 25: 82-88.
Lever, A. and McCave. I.M.. 1983. Eolian components in Cretaceous and Tertiary North Atlantic
sediments. Jour. Sed. Petrol., 53: 811-832.
Lewis, A.G. and Syvitski, J.P.M., 1983. The interaction of plankton and suspended sediments in fjords.
Sed. Geol.. 36: 81-92.
Lewis, C.F.M., 1966. Sedimentation studies of unconsolidated deposits in the Lake Erie Basin. Ph.D.
Thesis, U. Toronto, Toronto, Ont., 134 p.
Lewis, S.R., 1974. Significance of the vertical and lateral changes in the clay mineralogy of the
Dunbarton Triassic Basin, South Carolina. M.S. Thesis, U. North Carolina, Chapel Hill, NC. 34 p.
Liebling, R.S. and Scherp, H.S., 1976. Chlorite and mica as indicators of depositional environment and
provenance. Geol. Soc. Amer. Bull., 87: 513-514.
Liewig, N., Caron, J-M., and Clauer, N., 1981. Geochemical and K-Ar isotopic behavior of Alpine sheet
silicates during polyphased deformation. Tectonophysics, 78: 273-290.
Linares, J. and Huertas, F., 1971. Kaolinite synthesis at room temperature. Science. 171: 896-897.
Lindsey, D.A., 1970. Glacial sedimentology of the Precambrian Gowganda Formation, Ontario, Canada.
Geol. SOC.Am. Bull.. 80: 1685-1702.
Lion, L.W., Altmann, R.S. and Leckie, J.O., 1982. Trace-metal adsorption characteristics of estuarine
particulate matter: evaluation of contributions of Fe/Mn oxide and organic surface coatings.
Environ. Sci. Tech., 16: 660-666.
Lippmann. F., 1954. Uber einen Keuperton von Zaisers weiher bei Maulbronn. Heidelb. Beitr. Miner.,
Petrog. 4: 130-134.
Lippmann, F., 1956. Clay minerals from the Rot member of the Triassic near Gottingen, Germany. Jour.
Sed. Petrol., 26: 125-139.
757

Lippmann, F., 1976. Corrensite. a swelling clay mineral, and its influence on floor heave in tunnels in the
Keuper Formation. Bull. Internatl. Assoc. Eng. Geol., 13: 65-70.
Lisitzin, A.P.. 1972. Sedimentation in the World Ocean. SOC.Econ. Paleo. Mineral., Sp. Pub. No. 17, 218
P.
Lomova, 03..1975. The palygorskite clays of the Eastern Atlantic and their genetic association with
alkaline volcanism. Lithol. Miner. Resources. 6, No. 4: 10-27.
Lomova. 0,s..1979. Palygorskite and sepiolite as indicators of geological environments. Dokl. Akad.
Nauk SSSR, Geol. Inst., Tr.. 336 p.
Long, G., Neglia, S. and Favretto, L., 1964. Geochemical contribution to research for the reconstruction
of the paleogeography of a sedimentary basin. In: U. Columbo. and G.D. Hobsin (Editors), Adv. Org.
Geochem. McMillan, N.Y.. 239-256.
Longchambon, H., and Morgues, F., 1972. Sur le gisement de magnCsite de salinelles (Card). Bull. Soc.
Franq. Miner.. 50: 66-74.
Longstaffe. F.J., 1984a. The role of meteoric water in diagenesis of shallow sandstones: stable isotope
studies of the Milk River aquifer and gas pool, southeastern Alberta. In: D.A. McDonald and R.C.
Surdam (Editors), Clastic Diagenesis. Amer. Assoc. Pet. Geol., Tulsa, 81-98.
Longstaffe, F.J.. 1984b. Oxygen-isotope studies of diagenesis in the basal Belly River sandstone
Pembina-I-pool, Alberta. Clay Miner. Soc. Prog. with Abs., Baton Rouge, 83.
Lonsdale, P.. 1977. Deep-tow observations at the mounds abyssal hydrothermal field, Galapagos Rift.
Earth Planet. Sci. Lett., 36: 92-110.
Lopatin, N.V., 1971. Temperature and geologic time as factors in coalification. Akdemiya nauk, SSSR
ser. geologicheskaya, Izvestiya, No. 3: 95-106.
Lopatin, N.V.and Bostick. N.H., 1973. Nature of organic matter in recent and fossil sediments. Nauka
Press, Moscow. 79-90. (Illinois State Geol. Survey, Reprint Series, 1974q.)
Lopez-Aguayo, F. and Martin-Vivaldi, J.L., 1972. The Spanish Wealdian Clay: Its geological and
economical importance. Proc. Internatl. Clay Conf. Madrid. 749-761.
Louail, J., Esteoule, J. and Esteoule-Choux, J. 1979. The origin of clay minerals in Cenomanian littoral
deposits around the Armorican Massif. Proc. Inter. Clay Conf., Oxford, 291-300.
Loughman, F.C., 1957. A technique for the isolation of montmorillonite and halloysite, Amer. Miner. 42:
393-398.
Loughnan, F.C., 1969. Chemical Weathering of the Silicate Minerals. Elsevier, 154 p.
Love, S.K., 1936. Suspended matter in several small streams. Trans. Amer. Geophys. Un., pt. 2. 447-452.
Lovering, T.S., 1959. Significance of accumulator plants in rock weathering. Geol. SOC.Amer. Bull.. 70:
781-800.
Lucas, J., 1962. La transformation du minkraux argileux dans la stdimentation. Etudes sur les argiles du
Trias. MCm. Serv. Carte Gtol. Alcase Lorraine. 23. 202 p.
Lucas, J. and Ataman, G., 1968. Mineralogical and geochemical study of clay mineral transformations in
the sedimentary Triassic Jura Basin (France). Clays Clay Miner., 16: 365-372.
Lucas, J., Chaabani. F. and PrCvBt, L.. 1979. Phosphorites et Cvaporites: deux formations d e milieux
sidimentaires voisine tstudiCs dans la coupe du P a l t o g h e de Foum Selja (Metlaoui, Tunisie). Sci.
GCol. B ~ l l . ,32: 7-19.
Lucas, J. PrCvBt, L. and Trompette, R., 1980. Petrology, mineralogy and geochemistry of the late
Precambrian phosphate deposits of Upper Volta (W. Africa). Jour. Geol. SOC.,137: 787-792.
Lynch, L. and Reynolds, R.C., Jr., 1984. The stoichrometry of the smectite-illite reaction. Clay Miner.
SOC.Prog. with Abs.. Baton Rouge, 84.
Lynn. W.C. and Whittig, L.D., 1966. Alteration and formation of clay minerals during cat clay
development. Clays Clay Miner., 14: 241-248.
MacBustin, R. and Bayliss, P., 1979. Clay mineralogy of the Eureka Sound and Beaufort Formations,
Axel Heiberg and West Central Ellesmere Islands, Eastern Canadian Arctic Archipelago. Bull. Can
Petrol. Geol., 27: 446-452.
Macchi, L., Levison, A,, Curtis, C.D., Woodward, K., and Hughes, C.R., 1986. Chemistry, morphology,
and distribution of illites from Morecambe G a s Field. Irish Sea, Offshore United Kingdom. Amer.
Assoc. Pet. Geol. Progr. with abs., Atlanta. 615.
MacEwan. D.M.C., Ruiz, A.R. and Brown, G.. 1961. lnterstratified clay minerals. In: G . Brown (Editor).
The X-Ray Identification and Crystal Structures of Clay Minerals. Miner. SOC.London, 393-445.
758

MacEwan, D.M.C. and Wilson. M.J.. 1980. Interlayer and intercalation complexes of clay minerals. In:
G.W. Brindley and G . Brown (Editors), Crystal Structures of Clay Minerals and their X-ray
Identification. Miner. Soc. London. 197-248.
Mackenzie. F.T.. and Carrels. R.M., 1965. Silicates: reaction with seawater. Science, 150: 57-58.
Mackenzie. F.T. and Carrels, R.M., 1966. Chemical mass balance between rivers and oceans. Amer. Jour.
Sci.. 264: 507-525.
Mackenzie, R.C., 1957. Saponite from Allt Ribhein, Fiskavaig Bay. Skye. Miner. Mag. 31, 672-680.
Mackenzie, R.C., 1962. De Natura Lutorum. Proc. Clays Clay Miner., 11: 11-28.
Mackenzie, R.C., Wilson, M.J., and Mashhady, AS., 1985. Origin of palygorskite in some soils of the
Arabian Peninsula. In: A. Singer and E. Galen (Editors). Palygorskite-Sepiolite, Devlop. Sed. 37.
Elsevier, 177-186.
Mackin, J.E., 1986. Control of dissolved A1 distributions in marine sediments by clay reconstitution
reactions: experimental evidence leading to a unified theory. Geochim. Cosmochim. Acta, 50:
207-214.
Mackin. J.E. and Aller, R.C.. 1984. Dissolved Al in sediments and waters of the East China Sea:
implications for authigenic mineral formation. Geochim. Cosmochim. Acta., 48: 281 -297.
MacNeill. S.. 1978. A chemical investigation of a chlorite intergrade mineral in the Keuper Marl. Clay
Miner., 13: 357-365.
Mahjoory, R.A., 1975. Clay mineralogy, physical. and chemical properties of some soils in arid regions of
Iran. Soil Sci. Soc. Amer. Proc.. 39: 1157-1164.
Maisano. M.D., 1975. Petrology and depositional dynamics: fluvial Fort Union Group (Paleocene) silts,
Williston Basin, southwestern North Dakota. M.S. Thesis, Arizona State U.. Tempe, 184 p.
Maksimovii, Z.. 1966. P-kerolite-pirnelite series from GoleS Mountain, Yugoslavia. Proc. Int. Clay Conf..
Jerasalem, 97-105.
Makumbi, L. and Herbillon, A.J.. 1972. Vermiculitization expirimentale d’un chlorite. Bull. Groupe
Franq. Argiles. 24: 153-165.
Malachoff, A,, McMurtry, G.M.. Wiltshire, J.C. and Yeh, H.-W.. 1982. Geology and chemistry of
hydrothermal deposits from active submarine volcano Loihi, Hawaii. Nature 298: 234-239.
Malcolm. R.L., Nettleton, W.D. and McCracken. R.J., 1969. Pedogenic formation of montmorillonite
from 2 : 1-2: 2 intergrade clay mineral. Clays Clay Miner., 16: 405-414.
Malden. P.J. and Meads, R.E., 1967. Substitution by iron in kaolinite. Nature, 215: 844-846.
Malla. P.B. and Douglas, L.A.. 1987. Problems in identification of montmorillonite and beidellik. Clays
Clay Miner., 35: 232-236.
Manghnani. M.H., and Hower. J.. 1986. Glauconites: cation exchange capacities and infrared spectra.
Amer. Miner., 49: 586-598.
Manheim, F.T.. Hathaway, J.C. and Uchuppi. E., 1972. Suspended matter in surface waters of the
northern Gulf of Mexico. Limnol. Oceanogr., 17: 17-27.
Manker. J.P. and Griffin, G.M., 1971. Souce and mixing of insoluble clay minerals in a shallow water
carbonate environment - Florida Bay. Jour. Sed. Petrol., 41: 302-306.
Mankin. C.J. and Dodd, C.G., 1963. Proposed reference illite from the Ouachita Mountains of
southeastern Oklahoma. Proc. Clays Clay Miner., 10: 373-379.
Manley. F.H., 1973. Oceanward shoal areas and associated clay minerals in Recent marine carbonates.
Florida Keys. Geol. Soc. Amer. Prog. with Abs., S.E. Section. 416.
Manley. F.H.. Ogren. D.E. and Webb, L.C.. 1975. Mottled Upper Ordovician carbonates in northwest
Georgia. Jour. Sed. Petrol., 45: 615-617.
Marshall. C.E., 1955. Multifunctional ionization as illustrated by the clay minerals. Clays Clay Miner.. 2:
364-385.
Martin. F.M.. Garcia, J.G., and Lopez Azcona. M.C.. 1976. Estudio de las arcillas de las rocas
carhonhticas Jurhsicas y Crethcicas comprendidas entre las provincias de Guadalajara y Cuenca.
Estudios Geol.. 32: 204-213.
Martin. 1.. Brell, J.M. and Gal6n. E., 1976. Mineralogia de 10s materiales Terciarios del area de Alcall de
Nenares (Depresi6n del Tajo). Estudios Geol., 32: 105-1 13.
Martin. J.M.. Jednatok. J. and Pravdic, V.. 1971. The physico-chemical aspects of trace element behavior
in estuarine environments. Thalassia Jugosl.. 7: 619-637.
759

Martin, R.T., 1962. Adsorbed water on clay: a review. Clays Clay Miner. 9: 28-70.
Martin-Vivaldi, J.L., and Cano-Ruiz, J., 1955. Contribution to the study of sepiolite: 11. Some considera-
tions regarding the mineralogical formula. Clays Clay Miner. Nat. Acad. Sci. Nat. Res. Council Pub.
456: 166-172.
Martin-Vivaldi, J.L. and MacEwan, D.M.C., 1960. Corrensite and swelling chlorite. Clay Miner. Bull., 3:
177-183.
Masood, H., White, J.L. and Zinsmeister, W.J., 1986. Clay mineralogy and depositional environments of
Eureka Sound Formation, Canadian Arctic Archipelago. Clays Clay Miner. Prog. with Abs., Jackson,
MS, 29.
Masui, J. and Shoji, S., 1969. Crystalline clay minerals in volcanic ash soils of Japan. Proc. Inter. Clay
Conf., Tokyo, 383-392.
Mathieson, A. McL., 1958. Mg-Vermiculite: A refinement and re-examination of the crustal structure of
the 14.36 phase. Amer. Miner., 43: 216-227.
Mathieson, A,, McL., and Walker, G.F., 1954. Structure of Mg-Vermiculite. Amer. Miner., 39: 231-255.
Matti, J.C., Zemmels, I. and Cook, H.E., 1973. X-ray mineralogy of sediments from the far western
Pacific, Leg 20, DSDP. In: B.C. Heezen, I.D. MacGregor et al., Initial Reports of the DSDP, 20,
Wash. ( U S . Gov. Printing Office), 323-334.
Matti, J.C., Zemmels, I. and Cook, H.E., 1974a. X-ray mineralogy data, western Indian Ocean Leg 24,
Deep Sea Drilling Project. In: R.L. Fisher, E.T. Bunce et al., Initial Reports of the DSDP 24, Wash.
( U S Gov. Printing Office), 811-826.
Matti, J.C., Zemmels, I. and cook, H.E., 1974b. X-ray mineralogy data. In: R.B. Whitmarsh, O.E. Weser,
D.A. Ross et al.. Initial Reports of the DSDP, 23, Wash. ( U S . Gov.Printing Office), 1131-1136.
Maxwell, D.T. and Hower, J., 1967. High-grade diagenesis and low-grade metamorphism of illite in the
Precambrian Belt series. Amer. Miner., 52: 843-857.
McAllister, R.F., Jr., 1964. Clay minerals from West Mississippi Delta marine sediments. In: R.L. Miller
(Editor), Papers in Marine Geol. Shepard Commen. Vol., McMillan, 457-473.
McBride, M.B., 1976. Origin and position of exchange sites in kaolinite: Am ESR study. Clays Clay
Miner., 24: 88-92.
McCave. I.N., 1972. Transport and escape of fine-grained sediment from shelf areas. In: D.J.P. Swift,
D.B. Duane and O.H. Pilkey (Editors), Shelf Sediment Transport: Process and Pattern. Dowden,
Hutchinson and Ross, Inc. Stroudsburg, Pa., 225-248.
McCave, I.N., 1975. Vertical flux of particles in the ocean. Deep-sea Res., 22: 491-502.
McCave, I.N., 1983. Particulate size spectra, behavior, and origin of nepheloid layers over the Nova
Scotian continental rise. Jour. Geoph. Res., 88: 7647-7666.
McCrone, A.W., 1967. The Hudson River Estuary: sedimentary and geochemical properties between
Kingston and Haverstraw, New York. Jour. Sed. Petrol., 37: 475-486.
McCulloh, T.H., Cashman, S.M. and Stewart, R.J., 1978. Diagenetic baselines for interpretive reconstruc-
tion of maximum burial depths and paleotemperatures in clastic sedimentary rocks. In: D.F. Oltz
(Editor), Low Temperature Metamorphism of Kerogen and Clay Minerals. Pacific Sec. SOC.Econ.
Paleo. Miner., Los Angeles, 18-46.
McDowell, S.D. and Elders, W.A., 1980. Authigenic layer silicate minerals in borehole Elmore 1, Salton
Sea geothermal field, California, USA. Contrib. Miner. Petrol., 74: 293-310.
McDuff, R.E., 1981. Major cation gradients in DSDP interstitial waters: the role of diffusive exchange
between seawater and upper oceanic crust. Geochim. Cosmochim. Acta, 45: 1705-1713.
McHardy, W.J., Wilson, M.J., and Tait, J.M., 1982. Electron microscope and X-ray diffraction studies of
filamentous illitic clay from sandstones of the Magnus Field. Clay Miner., 17: 23-39.
McKeague, J.A. and Brydon, J.E., 1970. Mineralogical properties of ten reddish brown soils from the
Atlantic provinces in relation to parent materials and pedogenesis. Can. Jour. Soil Sci., 50: 47-55.
McKee, E.D. and Weir, G.W., 1953. Terminology for stratification and cross-stratification in sedimen-
tary rocks. Geol. Soc. Amer. Bull., 64: 381-389.
McKyes, E., Sethi, A. and Yong, R.N., 1974. Amorphous coatings on particles of sensitive clay soils.
Clays Clay Miner., 22: 427-433.
McLean, S.A., Allen, B.L. and Craig, J.R., 1972. The occurrence of sepiolite and attapulgite on the
southern High Plains. Clays Clay Miner., 20: 143-149.
760

McMurtry, G.M.. Wang, C.-H.. and Yeh. H.-W.. 1983. Chemical and isotopic investigations into the
origin of clay minerals from the Galapagos hydrothermal mounds field. Geochim. Cosmochim. Acta.
47: 475-489.
McMurtry, G.M. and Yeh, H.-W.. 1981. Hydrothermal clay mineral formation of East Pacific Rise and
Bauer Basin sediments. Chem. Geol., 32: 189-205.
McPherson. H.J., 1971. Dissolved, suspended and bed load movement patterns in Two O'clock Creek.
Rocky Mountains, Canada, Summer, 1969. Jour. Hydrology, 12: 221-233.
McRae. S.G.. 1972. Glauconite. Earth-Sci. Rev.. 8: 397-440.
Meade, R.H., 1966. Factors influencing the early steps of compaction of clays and sands A review.-

Jour. Sed. Petrol., 36: 1085-1101.


Meade, R.H., 1969. Landward transport of bottom sediments in estuaries of the Atlantic Coastal Plain.
Jour. Sed. Petrol., 39: 222-234.
Meade. R.H.. 1972. Transport and deposition of sediments in estuaries. Geol. Soc. Amer. Memoir 133:
91-120.
Meads. R.E. and MaId.cn, P.J., 1975. Electron spin resonance in natural kaolinites containing Fe3+ and
other transition metal ions. Clay Miner. 10: 313-345.
Mehmel, M.. 1935. Uber die struktur halloysit und metahalloysit. Z. Kristallogr. Kristallgeom. 90: 35-43.
Melson, W.G., Thompson, G. and van Andel, Tj.H., 1968. Volcanism and metamorphism in the
Mid-Atlantic Ridge, 22"N latitude. Jour. Geoph. Res., 73: 5925-5941.
Melson, W.G. and van Andel. Tj.H., 1966. Metamorphism of the Mid-Atlantic Ridge. 22"N latitude.
Mar. Geol., 4: 165-186.
Melson, W.G. and Thompson, G., 1973. Glassy abyssal basalts, Atlantic seafloor near St. Paul's Rocks:
petrography and composition of secondary clay minerals. Geol. Soc. Amer. Bull.. 84: 703-716.
Mkndez, P.A. and GalHn, E.. 1976. Estudio mineralbgico de la formacibn Voznuevo, entre Santiago de las
Villas y La Vecilla del Rio Curueiio (provincia de Lebn). Estudio Geol.. 32: 349-370.
Merino, E. and Ransom, B., 1982. Free energies of formation of illite solid solutions and their
compositional dependence. Clays Clay Miner., 30: 29-39.
Mermut. A.R.. Ghabru, S.K. and St.Amaud. R.J., 1987. Natural occurrence of iron hydroxy interlayered
vermiculite derived from biotite. Clay Miner. Soc. Prog. with Abs., 95.
Merrill, W.M. and Winar, R.M., 1958. Molas and associated formations in San Juan Basin - Needle
Mountain area southwestern Colorado. Bull. Amer. Assoc. Petrol. Geol., 42: 2107-21 32.
Mestdagh, M.M., Vielvoye, L. and Herbillon, A.J., 1980. Iron in kaolinite: I I The relationship between
kaolinite crystallinity and iron content. Clay Miner.. 15: 1-13.
Meunier, A. and Velde. B., 1976. Mineral reactions at grain contacts in early stages of granite weathering.
Clay Miner., 1 1 : 235-240.
Meunier, A. and Velde. B., 1979. Weathering mineral facies in altered granites: The importance of local
small-scale equilibria. Miner. Mag., 43: 261-268.
Middleton. M.F.. 1982. Tectonic history from vitrinite reflectance. Royal Astro. Soc. Geoph. Jour.. 68:
121-132.
Midgley, H.G., 1959. A sepiolite from Mullion, Cornwall. Clay Min. Bull.. 4: 88-93.
Miller. R., Cousins, P., Dillon, W., Folger. D. Knebel, H., 1978. Middle Atlantic outer Continental Shelf
environmental studies. Vol. 111. Geologic Studies. Bureau Land Manag. Rpt. BLM/ST-78/29.
281-299.
Milliman, J.D. and Meade, R.H., 1983. World-wide delivery of river sediment to the oceans. Jour. Geol..
91: 1-21.
Milliman, J.D., Summerhayes, C.P. and Barretto. H.T., 1975. Oceanography and suspended matter of the
Amazon River February-March 1973. Jour. Sed. Petrol.. 45: 187-206.
Millot. G . , 1949. Relations entre la constitution et la genise des roches skdimentaires argileuses. Gkologie
Appliquke et Prospection Miniire, Nancy, France, 2: 347p.
Millot, G.. 1970. Geology of Clays. Springer-Verlag. N.Y.. 429 p.
Millot. G.. and Camez. T., 1963. Genesis of vermiculite and mixed-layered vermiculite in the evolution of
the soils of France. Clays Clay Miner., Proc., 10: 90-95.
Millot, G.. Paquet, H. and Ruellen. A.. 1969. Neoformation de I'attapulgite dans les sols a carapaces
calcaires de la basse Moulouya (Maroc oriental). C.R. Acad. Sci.. Paris. 268: 2771-2774.
761

Milne, I.H. and Earley, J.W., 1958. Effect of source and environment on clay minerals. Bull. Amer.
Assoc. Petrol. Geol., 42: 328-338.
Milne, I.H. and Shott. W.L.. 1958. Clay mineralogy of recent sediments form the Mississippi Sound area.
Clays Clay Miner., 5: 253-265.
Mingarro Martin, P.F. and Lbpez De Azcona, M.C.. 1975. Estudio de la fraccion arcilloso contenida en
las rocas carboniticas del Cretacico Superior d e la provincia de Segovia. Estudios Geol., 31: 531-542.
Mintz, L.W.. 1981. Historical Geology. Charles E. Merrill Pub. Co., Columbus, OH, 611 p.
Mitchell, W.A., 1955. Review of the mineralogy of Scottish soil clays. Jour. Soil Sci.. 16: 94-98.
Miyashiro. A., Shido, F. and Ewing. M., 1971. Metamorphism in the Mid-Atlantic Ridge near 24 and
30"N. Phil. Trans. Roy. Soc. Lond., Series A268, 589-603.
Moberly, R. Jr., Kimura, H.S. and McCoy. F.W. Jr.. 1968. Authigenic marine phyllosilicates near Hawaii.
Geol. Soc. Amer. Bull., 79: 1449-1460.
Mohr, E.C.J. and Van Baren, F.A., 1954. Tropical Soils. lnterscience Pub. Ltd., London, 498 p.
Moncure, G.K., Lahann, R.W. and Siebert, R.M., 1984. Origin of secondary porosity and cement
distribution in a sandstone/shale sequence from the Frio Formation (Oligocene). In: D.A. McDonald
and R.C. Surdam (Editors), Clastic Diagenesis. Amer. Assoc. Pet. Geol., Tulsa, 151-161.
Moncure, G.K. and Force, L.M.. 1976. Potomac Group clays. Geol. Soc. Amer. Prog. with Abs., 8:
232-233.
Mongiorgi, R. and Morandi, N., 1970. A1 saponite e strati misti clorite-Al saponite melle idrotermaliti di
una breccia a contatto coi diabasi di Rossena nell'Appennino reggiano. Miner. Petrog. Acta. 16:
139-154.
Montague, C.L., 1984. Influence of biota on erodibility of sediments. In: A.J. Mehta (Editor). Lecture
Notes on Coastal and Estuarine Studies. 14 Estuarine cohesive Sediment Dynamics. Springer-Verlag.
251-269.
Moon, C.F., and Hurst, C.W., 1984. Fabric of muds and shales: an overview. In: D.A.V. Stow and
D.J.W. Piper (Editors), Fine-grained sediments: deep-water processes and facies. Blackwell Sci. Pub..
Oxford, 579-593.
Moore, H.B., 1936. The muds of the Clyde Sea area. 111 chemical and physical conditions; rate of
sedimentation; and infauna. Mar. Biol. Assoc. U.K.J., 17: 325-358.
Moore, J.E., 1961. Petrography of northeastern Lake Michigan bottom sediments. Jour. Sed. Petrol., 31:
402-436.
Moore, J.G.. 1966. Rate of palagonitization of submarine basalt adjacent to Hawaii. U.S. Geol. Survey
Prof. Paper 550-D, 163-171.
Moort, J.C. van, 1971. A comparative study of the diagenetic alteration of clay minerals in Mesozoic
shales from Papua, New Guinea, and in Tertiary shales from Louisiana, U.S.A. Clays Clay Miner..
19: 1-20.
Morad, S.. 1984. Diagenetic matrix in Proterozoic graywackes from Sweden. Jour. Sed. Petrol., 54:
1157-1 168.
Morey, G.B., 1967. Stratigraphy and sedimentology of the Middle Cambrian Rove Formation in
northeastern Minnesota. Jour. Sed. Petrol., 37: 1154-1 162.
Moriarty, K.C., 1977. Clay minerals in southeast Indian Ocean sediments, transport mechanisms and
depositional environments. Marine Geol., 25: 149-174.
Morris, A.W., Howland, R.J.M. and Bale, A.J., 1986. Dissolved aluminum in the Tamar Estuary.
southwest England. Geochim. Cosmochim. Acta. 50: 189-197.
Morris, R.C., 1974. Sedimentary and tectonic history of the Ouachita Mountains. In: W.R. Dickinson
(Editor), Tectonics and Sedimentation. Soc. Econ. Paleo. Miner.. Spec. Pub. 22: 120-142.
Morton, J.P., 1985. Rb-Sr evidence for punctuated illite/smectite diagenesis in the Oligocene Frio
Formation, Texas Gulf Coast. Geol. Soc.Amer. Bull., 96: 114-122.
Morton, J.P. and Long. L.E., 1980. Rb-Sr dating of Paleozoic glauconite from the Llano region, central
Texas. Geochim. Cosmochim. Acta 44: 663-672.
Morton, J.P. and Long, L.E., 1984. Rb-Sr ages of glauconite recrystallization; dating times of regional
emergence above sea level. Jour. Sed. Petrol., 54: 495-506.
Morton, R.W., 1972. Spatial and temporal distribution of suspended sediment in Narragansett Bay and
Rhode Island Sound. Geol. Soc.Amer. Memoir 133: 131-141.
762

Mossler, J.H. and Hayes, J.B., 1966. Ordovician potassium bentonites of Iowa. Jour. Sed. Petrol.. 36:
414-427.
Mottl, M.J., 1983. Metabasalts, axial hot springs, and the structure of hydrothermal systems at mid-ocean
ridges. Geol. Soc. Amer. Bull., 94: 161-180.
Mottl, M.J. and Holland, H.D., 1978. Chemical exchange during hydrothermal alteration of basalt by
seawater - 1. Experimental results for major and minor components of seawater. Geochim. Cosmo-
chem. Acta. 42: 1103-1115.
Muffler, L.J.P. and White, D.E., 1969. Active metamorphism of Upper Cenezoic sediments in the Salton
Sea geothermal field and the Salton Trough, southeastern California. Geol. Soc. Amer. Bull.. 80:
157-182.
Muller. A.. Parting, H. and Thorez, J., 1974. Caracttres sidimentologiques et miniralogiques des couches
de passage du Trias au Lias sur la bordure nord-est du Bassin de Paris. Ann. Soc. Giol. Belgique, 96:
671-707.
Muller. G. and Forstner, U.. 1973. Recent iron ore formation in Lake Malawi, Africa. Mineral.
Deposita., 8: 278-290.
Muller, G . and Quakernaat, J., 1969. Diffractometric clay mineral analysis of Recent sediments of Lake
Constance (Central Europe). Contr. Miner. Petrol., 22: 268-275.
Muller, J.-P. and Bocquier, G., 1985. Mineralogy and microstructural aspects of pedogenic alteration in a
lateritic formation of Cameroon. Int. Clay Conf.. Denver, Abs., 161.
Murray, H.H., 1954. Genesis of clay minerals in some Pennsylvanian shales of Indiana and Illinois. Clays
and Clay Min. Proc. 2nd Natl. Conf., Pub. 327, Natl. Acad. Sci. Natl. Res. council, 47-67.
-

Murray, H.H. and Duncan. J., 1977. Comparative mineralogy of the Greyson (Precambrian) and
Woolsey (Cambrian) shales near Cardwell, Montana. Geol. Soc. Amer. Prog. with Abs., 635.
Murray, H.H. and Leininger, R.K.. 1956. Effect of weathering on clay minerals. Clays Clay Miner.. 4:
340-347.
Murray, H.H. and Lyons, S.C., 1956. Degree of crystal perfection of kaolinite. Clays Clay Minerals Proc.
Fourth Conf. Natl. acad. Sci., Publ. 456: 31-40.
Murray, H.H. and Partridge, P., 1981. Genesis of Rio Jari kaolin. Proc. Int. Clay Conf., Bologna,
Elsevier, 279-291,
Murray, H.M. and Patterson, S.H., 1975. Kaolin, ball-clay, and fire-clay deposits in the United States -

Their ages and origins. Proc. Internatl. Clay Conf., Mexico, 511-520.
Nadeau, P.H. and Bain, D.C., 1986. Composition of some smectites and diagenetic illitic clays and
implications for their origin. Clays Clay Miner., 34: 455-464.
Nadeau, P.H., Farmer, V.C., McHardy, W.J. and Bain, D.C., 1985. Compositional variations of the
Unterrupsroth beidellite. Amer. Miner.. 70: 1004-1010.
Nadeau. P.H., and Reynolds, R.C. Jr.. 1981a. Burial and contact metamorphism in the Mancos Shale.
Clays Clay Miner., 29: 249-259.
Nadeau. P.N. and Reynolds, R.C., Jr.. 1981b. Volcanic components in pelitic sediments. Nature, 294:
72-74.
Nadeau, P.H., Tait, J.M., McHardy, W.J., and Wilson, M.J., 1984. Interstratified XRD characteristics of
physical mixtures of elementary clay particles. Clay Miner.. 19: 67-76.
Nagy, B.. and Bradley, W.F., 1955. Structure of sepiolite. Amer. Miner.. 40: 885-892.
Naidu, A.S. and Mowatt, T.C.. 1983. Sources and dispersal patterns of clay minerals in surface sediments
from the continental-shelf areas off Alaska. Geol. Soc. Amer. Bull., 94: 841-854.
Nanz, R.H., 1953. Chemical composition of Precambrian slates with notes on the geochemical evolution
of lutites. Jour. Geol.. 61: 51-64.
Narkis. N., Rebhum, M. and Sperber, H., 1968. Flocculation of clay suspensions in the presence o f humic
and fulvic acid. Israel Jour. Chem., 6: 298-305.
Natland, J.H. and Mahoney. J.J., 1981. Alteration in igneous rocks at Deep Sea Drilling Project Sites 458
and 459, Mariana Fore-Arc region: relationship to basement structure. In: D.M. Hussong. S. Uyeda
et al. (Editors), Initial Reports of the DSDP, 60, Wash. (U.S. Gov. Printing Office), 769-788.
Neacsu. G . and Neacsu, V., 1980. Fireclay and kaolin deposits in Romania. Acta Miner. Petrog.. 24:
39-45.
Neiheisel. J.. 1973. Long range spoil disposal study, Part 111, Sub-Study 2, Nature, source and cause of
the shoal, Appendix A. U S . Army Eng. Dist. Philadelphia, 140 p.
763

Neiheisel, J. and Weaver, C.E., 1967. Transport and deposition of clay minerals, southeastern United
States. Jour. Sed. Petrol., 37, 1084-1116.
Nelson, B.W., 1959. New bentonite zone from the Pennsylvanian of southwestern Virginia. Geol. SOC.
Amer. Bull., 70, 1651.
Nelson, B.W., 1960. Clay mineralogy of the bottom sediments, Rappahammock River, Virginia. Clays
Clay Miner., 7: 135-147.
Nelson, B.W., 1971. Mineralogical differentiation of sediments dispersed from the Po Delta. In: D.J.
Stanley (Editor), The Mediterranean Sea. Dowden, Hutchinson and Ross, Inc., Stroudsburg, Pa.,
441-453.
Nelson, B.W., 1973. Mineralogy of the Maccrady Formation near Saltville, Virginia. Amer. Jour. Sci.,
273-A: 539-565.
Nelson, B.W. and Roy, R., 1958. Synthesis of the chlorites and their structural and chemical constitution.
Amer. Miner. 43: 707-725.
Nelson, C.H., 1967. Sediments of Crater Lake, Oregon. Geol. SOC.Amer. Bull., 78: 833-848.
Nelson, D.D., 1973. Late Pleistocene and Holocene clay mineralogy and sedimentation in the Pamlico
Sound region, North Carolina. U. South Carolina, Ph.D. Dissertation, 103 p.
Nelson, W.A., 1922. Volcanic ash beds in the Ordovician of Tennessee, Kentucky and Alabama. Geol.
SOC.Amer. Bull., 33: 605-615.
Nelson, W.G. and Geoffrey, T., 1973. Glassy abyssal basalts, Atlantic sea floor near St. Paul’s Rocks:
Petrography and composition of secondary clay minerals. Geol. Soc.Amer. Bull., 84: 703-716.
Nesbitt, H.W., Markovris, G. and Price, R.C., 1980. Chemical processes affecting alkalis and alkaline
earths during chemical weathering. Geochim. Cosmochim. Acta, 44: 1659-1666.
Nesbitt, H.W. and Young, G.M., 1982. Early Proterozoic climates and plate motions inferred from major
element chemistry of lutites. Nature, 299: 715-717.
Nesbitt, H.W. and Young, G.M., 1984. Prediction of some weathering trends of plutonic and volcanic
rocks based on thermodynamic and kinetic considerations. Geochim. Cosmochim. Acta, 48:
1523-1534.
Nesteroff, W.D., 1973. Distribution of the fine-grained sediment component in the Mediterranean. In:
W.B.F. Ryan, K.J. Hsii et al., Initial Reports of the DSDP, 13, Wash. (U.S. Gov. Printing Office),
666-671.
Nettleton, W.D., Flack, K.W., and Nelson,, R.E., 1970. Pedogenic weathering of tonalite in southern
California. Geoderma, 4: 387-402.
Nettleton, W.D., Lynn, W.C., Klameth, L.C. and Baird, S.L., 1985. Palygorskite in arid soils of USA and
Jordan. Inter. Clay Conf. Prog. with abs., Denver, 166.
Nettleton, W.D., Nelson, R.E. and Flach, K.W., 1973. Formation of mica in surface horizons of dryland
soils. Soil Sci. Soc. Amer. Proc., 37: 473-478.
Newnham, R.E., 1961. A refinement of the dickite structure and some remarks on polymorphism in
kaolin minerals. Mineral. Mag., 32: 683-704.
Nichols, M.M., 1972. Sediments of the James River Estuary, Virginia. Geol. Soc. Amer. Memoir 133:
169-212.
Niederbudde, E.A. and Fischer, W.R., 1980. Clay mineral transformation in soils as influenced by
potassium release from biotite. Soil Sci., 130: 225-231.
Nielsen, O.B., 1974. Sedimentation and diagenesis of Lower Eocene sediments at O h , Denmark. Sed.
Geol., 12: 25-44.
Niem, A.R., 1977. Mississippian pyroclastic flow and ash-fall deposits in the deep-marine Ouachita
flysch basin, Oklahoma and Arkansas. Geol. SOC.Amer. Bull., 88: 49-61.
Nishiyama, T. and Shimoda, S., 1981. Ca-bearing rectorite from Tooho Mine, Japan. Clays Clay Miner.,
29: 226-240.
Norrish, K., 1954. Swelling of montmorillonite. Trans. Faraday SOC.,18: 120-134.
Norrish, K., 1972. Factors in the weathering of mica to vermiculite. Proc. Inter. Clay Conf., Madrid,
417-432.
Norrish, K. and Pickering, J.G., 1983. Soils: an Australian Viewpoint. Div. Soils, CSIRO, Melbourn
Academic Press, London, 281-308.
vick, B.E. and Martin, R.T., 1983. Solvation methods for expanded layers. Clays Clay Miner., 31:
235-238.
764

Novak. R.J., Motto, H.L. and Douglas, L.A., 1971. The effect of time and particle size on mineral
alteration in several Quaternary soils in New Jersey and Pennsylvania, U.S.A. In: D.H. Yaalon
(Editor), Paleopedology. Israel University Press, Jerusalem, 21 1-224.
Nuhfer, E.B., Vinopal, R.J. and Klanderman, E.S., 1979. X-radiograph atlas of lithotypes and other
structures in the Devonian shale sequence of West Virginia and Virginia. Natl. Tech. Inf. Service,
Springfield, Va., , METC/CR-79/27, 45 p.
OBrien, N.R., 1970. The fabric of shale - an electron-microscope study. Sedimentology. 15: 229-246.
O’Brien, N.R., 1971. Fabric of kaolinite and illite floccules. Clays Clay Miner., 19: 353-359.
OBrien, N.R. and Burrell, D.C., 1970. Mineralogy and distribution of clay size sediment in Glacier Bay.
Alaska. Jour. Sed. Petrol., 40: 650-655.
O’Brien. N.R., Nakozawa, K. and Takuhashi. S., 1980. Use of clay fabric to distinguish turbiditic and
hemipelagic siltstones and silts. Sedimentology, 27: 47-61.
Odin. G.S.. 1982. Effect of pressure and temperature on clay mineral potassium-argon ages. In: G.S.
Odin (Editor), Numerical Dating in Stratigraphy. John Wiley and Sons, 307-319.
Odin, G.S., 1984. “Marine herthierine”. further data and questions on its mineralogy and destiny. Clay
Miner. Soc. Meet., Prog. with Abs.. 91.
Odin. G.S., 1985. La “verdine” faci6s granulaire vert, marin et cAtier, distinct de la glauconie:
distribution actuella et composition. C.R. Acad. Sc. Paris, 301, Serie 11, No. 2: 105-108.
Odin, G.S.. Curry, D. and Hunziker, J.C., 1978. Radiometric dates from Northwestern Europe glauconites
and the Palaeogene time-scale. Jour. Geol. Soc. Lond., 135: 481-497.
Odin. G.S. and Dodson, M.H.. 1982. Zero isotopic age of glauconies. In: G.S. Odin (Editor), Numerical
Dating in Stratigraphy. John Wiley and Sons, 277-305.
Odin, G.S. and Mather, A., 1981. De glauconiarum origine. Sedimentology. 28: 611-641.
Odin. G.S. and Stephan, J.F., 1981. The occurrence of deep water glaucony from the eastern Pacific: The
result of in situ genesis or subsidence. In: J.S. Watkins, J.C. Moore et al., Initial Reports DSDP. 66,
Wash. ( U S . Govt. Printing Office), 419-428.
Odom. 1.E.. 1967. Clay fabric and its relation to structural properties in mid-continent Pennsylvanian
sediments. Jour. Sed. Petrol., 37: 610-623.
Odom, I.E., 1976. Microstructure, mineralogy and chemistry of Cambrian glauconite pellets and
glauconite. central U.S.A. Clays Clay Miner., 24: 232-238.
Odom. I.E.. Willand. T.N. and Lassin, R.J., 1979. Paragenesis of diagenetic minerals in the St. Peter
Sandstone (Ordovician) Wisconsin and Illinois. In: P.A. Scholle and P.R. Schluger (Editors), Aspects
of Diageneis. SEPM Spec. Pub. No. 26. Tulsa, 425-443.
Oinuma. K. and Kobayashi, K., 1966. Quantitative studies of clay minerals in some recent marine
sediments and sedimentary rocks from Japan. Clays Clay Miner., 14: 209-219.
Ojanuga. A.G., 1973. Weathering of biotite in soils of a humid tropical climate. Soil Sci. Soc. Amer. Proc.
37: 644-646.
Okada, H. and Tomita, K.. 1973. Clay mineralogy of deep-sea sediments in the northwestern Pacific,
DSDP, Leg 20. In: B.C. Heezen, I.D. MacGregor et al., Initial Reports of the DSDP, 20. Wash. (U.S.
Govt. Printing Office), 335-343.
Okusami, T.A., 1985. Environmental factors and properties of soils with mixed-layer clay mineralogy in
the tropics. Int. Clay Conf. Denver, Ahs.. 172.
Olorunfemi. B.N.. Fyfe, W.S., and Kronberg, B., 1983. Clay diagenesis as a function of marine.
non-marine Niger Delta. Clays Clay Miner., 20th Ann. Meet., Abs.. 44.
Ordonez. S. and Aguayo, L., 1982. Mudstone associated with sodic salts deposits. Madrid Tertiary Basin
(central Spain). Eleventh Internatl. Congress Sed., Hamilton, Canada, Abs., 178.
Ortega-Huertas, M., Rodriguez-Gallego, M. and Lbpez-Aguayo, F., 1979. Mineralogia d e la fraccibn fina
de la “Block Formation”. Depresibn de Granada. Estudios Geol.. 35: 541-548.
Oveharenko, F.D., 1984. Palygorskite and sepiolite deposits in the U.S.S.R. and their uses. In: A. Singer
and E. Galan (Editors), Palygorskite-Sepiolite. Elsevier, 233-241.
Ovejero, R. and Bossi, G., 1974. Asociaciones mineralogicas de las argillas de la cuenca de Ischigualasto-
Ischichuga. Parte 111: Perfil Zanja de la Viuda. Noveno Cong. Geol. Argentino. S.C. De Bariloche. V.
197-208.
Oyawaya. M.O. and Hirst, D.M., 1964. Occurrence of a montmorillonite mineral in the Ingerian Younger
Granites at Ropp, Plateau Province. northern Nigeria. Clay Miner., 5: 427-433.
765

Padan, A., 1984. Clay mineralogy of the bedded salt deposits in the Paradox Basin, Gibson Well No. 1,
Utah. Ph.D. Thesis, Ga. Inst. Tech., Atlanta, Ga., 272 p.
Padan, A. and Weaver, C.E., 1985. Pennsylvanian clay-salt association in Paradox Basin, Utah. In press.
Padan, A., Weaver, C.E. and Wampler, J.M., 1984. The formation of evaporitic clay minerals, Paradox
Formation, Utah. Clay Miner. Conf., Baton, Rouge, Prog. with Abs., 94.
Paerl, H.W., 1973. Detritus in Lake Tahoe: Structural modification by attached microflora. Science, 180:
496-498.
Paerl. H.W., 1975. Microbial attachment to particles in marine and fresh water ecosystems. Microbial
Ecol., 2: 73-83.
Palacio, M.R., 1967. The white clays of Lara State. Bol. Info. Assoc. Venez. Geol., Miner. Petrol., 10:
141-175.
Papke, K.G., 1969. Montmorillonite deposits in Nevada. Clays Clay Miner., 17: 211-222.
Papke, K.G., 1972. A sepiolite-rich playa deposit in southern Nevada. Clays Clay Miner., 20: 211-215.
Paquet, H., 1970. Evolution gtochimique des mintraux argileux dans les alttrations et les sols des climats
mtditerranbens tropicaux A saisons contrasttes. Mtm. Serv. Carte Gtol. DAlsace Lorraine, No. 30,
210 p.
Paquet, H., Duplay, J. and Nahon, D., 1982. Variations in the composition of phyllosilicates monopar-
ticles in a weathering profile of ultrabasic rocks. Proc. Inter. Clay Conf., Bologna and Pavia, 595-603.
Paquet, H. and Millot, G., 1972. Geochemical evolution of clay minerals in the weathered products and
soils of Mediterranean climates. Proc. Inter. Clay Conf., Madrid, 199-206.
Parachoniak, W. and Srodoh, J., 1973. the formation of kaolinite, montmorillonite and mixed-layer
montmorillonite-illites during the alteration of Carboniferous tuff (The Upper Silesian Coal Basin).
Mineral. Polonicia, 4: 37-52.
Parham, W.E., 1963. Lateral clay-mineral variations in certain Pennsylvania underclays. Clays Clay
Miner. 12 Conf., 581-612.
Parham, W.E., 1966. Lateral variations of clay mineral assemblages in modern and ancient sediments.
Proc. Inter. Clay Conf., Jerusalem, 1, 135-146.
Parham, W.E. and Hogberg, R.K., 1964. Kaolin clay resources of the Minnesota River valley Brown,
Redwood and Renville Counmties, a preliminary report. Minn. Geol. Sur., Rpt. Invest. 3, 43 p.
Parkin, D.W. and Padgham, R.C., 1975. Further studies on trade winds during the glacial cycles. Proc.
Royal Soc. Lond., A346: 245-260.
Parfenova, E.I. and Yarilova, E.A., 1962. Mineralogical Investigations in Soil Science. Moscow Acad. Sci.
U.S.S.R. (translation by A. Gourevitch and N. Kaner, Jerualem, Israel Program for Scientific
Translations, 1965). 177 p.
Parrish, J.T., Ziegler, A.M. and Scotese, C.R., 1982. Rainfall patterns and the distribution of coals and
evaporites in the Mesozoic and Cenezoic. Paleogeo., Paleoclim., Palaeoecol., 40: 67-101.
Parry, W.T. and Reeves, C.C., 1968. Sepiolite from Pluvial Mound Lake, Lynn and Terry Counties.
Texas. Amer. Miner., 53: 984-993.
Pastouret, L., Chamley, H., Delibrias, G., Duplessy, J.C. and Thiede, J., 1978. Late Quaternary climatic
changes in Western Tropical Africa deduced from deep-sea sedimentation off the Niger Delta.
Oceanologica Acta, 1: 217-232.
Patterson, E.M. and Gillette, D.A., 1977. Commonalities in measured size distributions for aerosols
having a soil-derived component. Jour. Geoph. Res., 82: 2074-2082.
Patterson, S.H., 1967. Bauxite reserves and potential aluminum resources of the world. U.S. Geol. Survey
Bull. 1228, 176 p.
Patterson, S.H. and Buie, B.F., 1974. Field conference on kaolin and fuller’s earth, Nov. 14-16, 1974. Ga.
Geol. Survey Guidebook 14, 53 p.
Patterson, S.H. and Hosterman, J.W., 1960. Geology of the clay deposits in the Olive Hill district,
Kentucky. Clays Clay Miner., 7: 178-195.
Pearson, M.J., Watkins, D. and Small, J.S., 1982. Clay diagenesis and organic maturation in northern
North Sea sediments. Proc. Inter. Clay Conf., Bologna and Pavia, 665-675.
Pedio, G., Carmouze, J.P. and Velde, B., 1978. Peloidal nontronite formation in recent sediments of Lake
Chad. Chem. Geol., 23: 139-149.
Pelzer, E.E., 1966. Mineralogy, geochemistry and stratigraphy of the Besa River Shale, British Columbia.
Bull. Can. Petrol. Geol., 14: 273-321.
766

Perrin, R.M.S., 1971. The Clay Mineralogy of British Sediments. London Miner. SOC., London, 247 p.
Perry, E.A., Jr., 1969. Burial diagenesis in Gulf Coast pelitic sediments. Ph.d. Dissertation. Case Western
Reserve U., Cleveland, Ohio. 125p.
Perry, E.A., Jr., 1974. Diagenesis and the K-Ar dating of shales and clay minerals. Geol. SOC.Amer. Bull..
85: 827-830.
Perry, E.A., Jr., 1976. Chemical and mineralogical studies, Sites 322 and 325. In: C.D. Hollister, C.
Craddock et al. (Editors), Initial Reports of the Deep sea Drilling Project, V. 35, Wash. (U.S. Gov.
Printing Office), 465-470.
Perry, E.A., Jr. and Hower, J., 1970. Burial diagenesis of Gulf Coast pelitic sediments. Clays Clay Miner..
18: 165-177.
Perry, E.A.. Jr. and Hower. J., 1972. Late-stage dehydration in deeply buried pelitic sediments. Bull.
Amer. Assoc. Petroleum Geol., 56: 2013-2021.
Persoz. F., 1982. Inventaire mintralogique, diagenkse des argiles et mintralostratigraphie des shies
jurassiques et critacies infkrieures du plateau suisse et de la bordure sud-est du Jura entre les lacs
d'Annecy et de Constance. Mattriaux pour la Carte Gtol. Suisse, Nouvelle sirie. 155 livraison, 52 p.
Peters, Tj. and Hofmann, B., 1984. Hydrothermal clay mineral formation in a biotite-granite in northern
Switzerland. Clay Miner., 19: 579-590.
Peterson, M.N.A., 1961. Expandable chloritic clay minerals from Upper Mississippian carbonate rocks of
the Cumberland Plateau in Tennessee. Amer. Miner., 46: 1245-1269.
Peterson, M.N.A., 1962. The mineralogy and petrology of Upper Mississippian carbonate rocks of the
Cumberland Plateau in Tennessee. Jour. Geol., 70: 1-31.
Peterson, M.N.A., Edger, N.T., Von der Borch, C.C. and Rex, R.W., 1970. Cruise summary and
discussion. In: M.N.A. Peterson et al.. Initial Report of the DSDP, 2, Wash. (U.S. Gov. Printing
Office), 413-427.
Peterson, M.N.A. and Griffin, J.J., 1964. Volcanism and clay minerals in the southeastern Pacific. Jour.
Marine Res., 22: 13-21.
Petree, D.H., Jr., 1972. The influence of grain size on the clay mineral composition of sediments in the
Neuse River Estuary, North Carolina. M.S. Thesis, U. N.Carolina, Chapel Hil, 30 p.
Petrov, V.P., 1958. Genetic types of white clays in the U.S.S.R. and laws governing their distribution.
Clay Min. Bull., 1 : 287-296.
Petruk, W., 1964. Determination of the heavy atom content in chlorite by means of the X-ray
diffractometer. Amer. Miner., 49: 61-71.
Pettijohn. F.J., 1975. Sedimentary Rocks. 3rd Ed. Harper and Row, NY, 628 p.
Pettijohn, F.J., Potter, P.E. and Siever, R., 1973. Sand and Sandstone. Springer-Verlag, NY, 618 p.
Pevear, D.R., 1972. Source of recent nearshore marine clays, southern United States. In: B.W. Nelson
(Editor), Environmental Framework of Coastal Plain Estuaries. Geol. SOC. Amer. Memori 133:
317-336.
Pevear, D.R., Williams, V.E. and Mustoe, G.E., 1980. Kaolinite. smectite, and K-rectorite in bentonites:
relation to coal rank at Tulameen, British Columbia. Clays Clay Miner., 28: 241-254.
Philippi, G., 1965. On the depth, time and mechanism of petroleum generation. Geochim. Cosmochim.
Acta, 29: 1021-1049.
Phillips, T.L., Goodwin, S., Peppers, R.A., Dimichele, W.A., 1983. Plant ecology of Pennsylvanian coal
swamps as an interregional and stratigraphic indicator of changes in climate. Geol. Soc. Amer. Bull.,
Prog. with Abs., 661.
Picard, M.D., 1953. Marlstone - a misnomer as used in the Unita Basin. Utah. Bull. Amer. Assoc. Petrol.
Geol., 37: 1075-1077.
Picard, M.D., 1971. Classification of fine-grained sedimentary rocks. Jour. Sed. Petrol., 41: 179- 195.
Picard, M.D. and High, L.R., Jr., 1972. Criteria for recognizing lacustrine rocks. In: J.K.Rigby and W.K.
Hamblin (Editors), Recognition of Ancient Sedimentary Environments. Soc. Econ. Paleo. and Miner.
Sp. Pub. No. 16: 108-159.
Pierce, J.W.. 1976. Suspended sediment transport at the shelf break and over the outer margin. In: D.J.
Stanley and D.J.P. Swift (Editors), Marine Sediment transport and Environmental Management. John
Wiley, NY, 612 p.
Pierce, J.W., Nelson, D.D. and Colquhoun, D.J., 1972. Mineralogy of suspended sediment o f f the
767

southeastern United States. In: D.J.P. Swift, D.B. Duane and O.H. Pilkey (Editors), Shelf/Sediment
Transport: Process and Pattern. Dowden, Hutchinson and Ross, Inc., Stroudsburg, Pa., 281-305.
Pinsak, A.P. and Murray, H.H., 1960. Regional clay mineral patterns in the Gulf of Mexico. Clays Clay
Miner. 7, 162-177.
Piper, D.J. and Slatt, R.M., 1977. Late Quaternary clay-mineral distribution on the eastern continental
margin of Canada. Geol. SOC.Amer. Bull., 88: 267-272.
PlanGon, A. and Tchoubar, C., 1977. Determination of structural defects in phyllosilicates by X-ray
powder diffraction - 11. Nature and proportion of defects in natural kaolinites. Clays Clay Miner.,
25: 436-450.
Pollard, C.O., Jr., 1971. Semidisplacive mechanism for the diagenetic alteration of montmorillonite layers
to illite layers: Appendix to Weaver, C.E. and Beck, K.C., Clay water diagenesis during burial, how
mud becomes gneiss. Geol. SOC.Amer. Spec. Paper. 134: 79-93.
Pollastro, R.M., 1985. Mineralogical and morphological evidence for the formation of illite at the expense
of illite/smectite. Clays Clay Miner., 33: 265-274.
Pollastro, R.M., Pillmore, C.L., Tschudy, R.H., Orth, C.J. and Gilmore, J.S., 1983. Clay petrology of the
conformable Cretaceous/Tertiary boundary interval, Raton basin, New Mexico and Colorado. Clay
Miner. Soc., Prog. with Abs., 83-84.
Poncelet, G. M. and Brindley, G.W., 1967. Experimental formation of kaolintie from montmorillonite at
low-temperatures. Amer. Miner., 52: 1161-1173.
Ponder, H. and Keller, W.D., 1960. Geology, mineralogy, and genesis of selected fireclays from Latah
County, Idaho. Clays Clay Miner., 8: 44-62.
Poppe, L.J., Hathaway, J.C. and Parmenter, C.M., 1983. Talc in the suspended matter of the northwest-
ern Atlantic. Clays Clay Miner., 31: 60-64.
Porrenga, D.H., 1966. Clay minerals in recent sediments of the Niger Delta. Clays Clay Miner., 14:
221-233.
Porrenga, D.H., 1968. Non-marine glauconitic illite in the Lower Oligocene of Aardenburg, Belguim.
Clay Miner. Bull., 17: 321-429.
Porthault, B., 1979. Profile gbochimique de la plate-forme Urgovienne au Bassin Vocontien (sud-est de la
france). Gtobios., MCmoire Special, 347-359.
Porthault, B. and Tixier, M., 1975. Evolution sedimentologique et gtochimique du Jurassic Infkrieur Du
Bassin Guercif (Marc Oriental). Congres Internat. de Sed., IX: 336-341.
Post, J.L, 1978. Sepiolite deposits of the Las Vegas, Nevada area. Clays Clay Miner., 26: 58-64.
Post, J.L. and Janke, N.C, 1974. Properties of 'swelling' chlorite in some Mesozoic formations of
California. Clays Clay Miner., 22: 67-77.
Potter, P.E. and Glass, H.D., 1958. Petrology and sedimentation of the Pennsylvanian sediments in
Southern Illinois: A vertical profile. Ill. Geol. Surv. Rpt. Invest. 204, 60 p.
Potter, P.E., Heling, D., Shimp, N.F.and VanWie, W., 1975. Clay mineralogy of modern alluvial muds of
the Mississippi River Basin. Bull. Centre Rech. Pau-SNPA, 9: 353-389.
Potter, P.E., Maynard, J.B. and Pryor, W.A., 1980. Sedimentology of Shale. Springerr-Velag. Heidelberg,
306 p.
Powell, T.G., Foscolos, A.E., Gunther, P.R., and Snowdon, L.R., 1978. Diagenesis of organic matter and
fine clay minerals: a comparative study. Geochim. Cosmochim. Acta, 42: 1121-1197.
Power, P.E., 1969. Clay mineralogy and paleoclimatic significance of some red regoliths and associated
rocks in western Colorado. Jour. Sed. Petrol., 39: 876-890.
Powers, M.C., 1959. Adjustment of clays to chemical changes and the concept of the equivalue level.
Clays Clay Miner., 6: 309-326.
PravdiC, V., 1970. Surface charge characterization of sea sediments. Limnol. Oceanogr., 15: 230-233.
Price, N.B. and Duff, P.M.D., 1969. Mineralogy and chemistry of tonsteins from Carboniferous
sequences in Great Britain. Sedimentology, 13: 45-69.
Prospero, J.M., 1981. Eolian transport to the world ocean. In: C. Emiliani (Editor), The Oceanic
Lithosphere, The Sea, 7. John Wiley and Sons, N.Y., v, 801-874.
Prospero, J.M. and Bonatti, E., 1969. Continental dust in the atmosphere of the eastern Equatorial
Pacific. Jour. Geoph. Res., 74: 3362-3371.
Prospero, J.M., Bonatti, E., Schubert, C., and Carlson, T.N., 1970. Dust in the Carribbean atmosphere
traced to an African dust storm. Earth. Planet. Sci. Lett., 9: 287-293.
768

Proust, D., 1981. Supergene alteration of hornblende in an amphibolite from Massif Central, France.
Proc. Inter. Clay Conf., Bologna and Pavia, 357-364.
Proust. D., Eymery, J.-P., and Beaufort, D., 1986. Supergene vermiculitization of a mgnesium chlorite:
iron and magnesium removal processes. Clays Clay Miner., 34: 572-580.
Pryor. W.A., 1975. Biogenic sedimentation and alteration of argillaceous sediments in shallow marine
environments. Geol. Soc. Amer. Bull., 86: 124441254,
Pryor, W.A. and Glass, H.D., 1961. Cretaceous-Tertiary Clay Minerals of the Upper Mississippi
Embayment. Jour. Sed. Petrol., 31: 38-51.
Purton, M.J. and Yowell, R.F., 1969. An X-ray investigation of some argillaceous rocks from the Skipton
Anticline, Yorkshire. Clay Miner., 10, 29-37.
Pusey. W.C., 1973a. How to evaluate potential gas oil source rocks. World Oil, 176. no 5 : 71-75.
Pusey. W.C., 1973b. Paleotemperatures in the Gulf Coast using the ESR-kerogen method. Trans. Gulf
Coast Assoc. Geol. Soc.. 23: 195-202.
Quaide. W.L, 1956. Petrography and clay mineralogy of Pliocene sedimentary rocks from the Ventura
Basin, California. Ph.D. Thesis, U. Calif., Berkely, 120 p.
Quakernaat, J., 1968. X-ray analyses of clay minerals in some Recent fulviatile sediments along the coasts
of central Italy. Ph.D. Dissertation, Amsterdam, 105 p.
Quinn, A.W., and Glass, H.W.. 1958. Rank of coal and metamorphic grade of rocks of the Narragansett
basin of Rhode Island. Econ. Geol., 53: 563-576.
Quiqley, R.M. and Thompson, C.D., 1966. The fabric of anisotropically consoliditated sensitive marine
clay. Can. Geotech. Jour., 3(2): 62-73.
Raam, A,, 1968. Petrology and diagenesis of Broughton sandstone (Permian) Kiama District, New South
Wales. Jour. Sed. Petrol., 38: 319-331.
Rabenhorst, M.C., Fanning, D.S., and Foss, J.E.. 1982. Regularly interstratified chlorite/vermiculite in
soils over meta-igneous rnafic rocks in Maryland. Clays Clay Minerals, 30: 156-158.
Radoslovich, E.W., 1958. Clay mineralogy of some Australian red-brown earths. Jour. Soil Sci.. 9:
242-250.
Radoslovich, E.W.. 1960. The structure of muscovite, KAI,(Si,, Al)O,,,(OH),. Acta Cryst., 13: 919-930.
Radoslovich, E.W., 1962. The cell dimensions and symmetry of layer-lattice silicates. 11. Regression
relations. Amer. Miner., 47: 617-636.
Radoslovich, E.W., 1963. The call dimensions and symmetry of layer-lattice silicates. IV. Interatomic
forces. Amer. Miner. 48: 76-99.
Ramamohana. R., 1977. Distribution of elements between coexisting phengite and chlorite from the
greenschist facies of the Tennant Creek area, central Australia. Lithos 10: 103-112.
Ramon, K.V. and Jackson, M.L., 1966. Layer charge relations in clay minerals of micaceous soils and
sediments. Clays Clay Miner., 14; 53-68.
Ramon, K.V. and Jackson, M.L.. 1964. Vermiculite surface morphology. Clays Clay Miner., 12: 424-429.
Ramos, A.N. and Formoso, M.L.L., 1976. Clay mineralogy of the sedimentary rocks of the Panama
Basin. Brazil. Revista Brasileira de Geosciencias, 6: 15-42.
Ramseyer, K. and Boles, J.R., 1986. Mixed-layer illite/smectite minerals in Tertiary sandstones and
shales, San Joaquin Basin, California. Clays Clay Miner., 34: 115-124.
Rao, C.G. and Bhattacharya, N., 1973. Corrensite from the Sirban Limestone of Riasi Jamma and
Kashmir State, India. Jour. Geol. Soc. India, 14: 193-196.
Rao. C.R. and Gluskater, H.J., 1973. Occurrence and distribution of minerals in Illinois coals. 111. State
Geol. Survey Cir. 476, 56 p.
Rashid, M.A., Buckley, D.E. and Robertson, K.R., 1972. Interaction of a marine humic acid with clay
minerals and a natural sediment. Geoderma, 8: 11-27.
Rateev, M.A., 1964. Regularities in the distribution and genesis of clay minerals in recent and old marine
basins. U.S.S.R. Acad. Sci., Geol. Inst., Trans. 12, 288 p.
Rateev. M.A., EmelLanov, E.M. and Kheirov, M.B., 1966. Conditions for the formation of clay minerals
in contemporaneous sediments of the Mediterranean Sea. Lithology Mineral Resources, 5 : 41 8-431.
Rateev, M.A., Gorbunova, Z.N., Lisitzyn. A.P., and Nosov, G.L., 1969. The distribution of clay minerals
in the oceans. Sedimentology. 13: 21-43.
Rateev. M.A. and Gradusov. B.P., 1970. A structural series of mixed-layer formations from the
Ordovician-Silurian metabentonites of the Baltic area. Dokl. Akad. Nauk SSSR. 194: 180-183.
769

Rateev, M.A., Gradusov, B.P. and Kheirov, M.B.. 1969. Potassium rectorite from the Upper Carbonifer-
ous of the Samarskaya Luka (Samara Bend of the Volga). Dokl. Akad. Nauk SSSR, 185: 116-119.
Rateev, M.A. and Kotel'nikes, D.D., 1956. New occurrences of a-sepiolite in the Carboniferous of the
Russian Platform. Dokl. Akad. Nauk SSSR 109: 191-194.
Rateev, M.A. and Timofeev, P.P., 1979. Palygorskite and sepiolite associated with arid facies in
Carboniferous rocks of the Russian Platform. In: E.S. Belt and R.W. Macqueen (Editors), N e u v i h e
Congr&s International de Stratigraphie et de Gtologie du Carbonifere, Washington and Champaign-
Urbana, 3: 669-673.
Rateev, M.A., Timofeev, P.R. and Rengarten, N.W.. 1980. Minerals of the clay fraction in Pliocene-
Quaternary sediments of the east Equatorial Pacific. In: B.R. Rosendahl, R. Hekinian, et al.. Initial
Reports of the Deep Sea Drilling Project 54, Wash. (U.S. Gov. Print. Office), 307-318.
Rateev, M.A., Voznesenskays, T.A. Gradusov, B.P.. 1980. Clay formation in the early geosynclinal
Skamara Basin of the southern Ural Paleozoic. Litol. Polezn, Iskop, 6: 101-115.
Rateev, M.A.. Gradusov, B.P. and Kheisov, M.B., 1969. Potassium rectorite from the Upper Carbonifer-
ous of the Samarskaya Luka (Samara Bend of the Volga). Dokl. Akad. Nauk. S.S.S.R., 185: 116-119.
Raup, O.B., 1966. Clay mineralogy of Pennsylvanian redbeds and associated rocks flanking Ancestral
Front Range of Central Colorado. Amer. Assoc. Petrol. Geol. Bull., 50: 251-268.
Rayner, J.H. and Brown, G., 1973. The structure of talc. Clays Clay Miner., 21: 103-114.
Reiche. P., 1945. A survey of weathering processes and products. U. New Mex. Pub. in Geology, No. 3,
95 p.
Reichenbach, H.G. and Rich, C.I., 1975. Fine-grained micas in soils. In: J.E. Gieseking (Editor), Soil
Components, V.2 Inorganic Components. Springer-Verlag, Berlin, 59-96.
Reineck, H.E. and Singh, I.B., 1973. Depositional Sedimentary Environments-with Reference to Terrige-
nous Clastics. Springer-Verlag. 439 p.
Rengasamy, P., 1976. Substitution of iron and titanium in kaolinite. Clays Clay Miner., 24: 265-266.
Rengasamy, P., Krishna Murti, G.S.R. and Sarma, V.A.K., 1975. Isomorphous substitution of iron for
aluminum in some soil kaolinites. Clays Clay Miner, 23: 211-214.
Renton, J.J., Cecil, C.B., Stanton, R. and Dulong, F., 1980. Compositional relationships of plants and
peats from modern peat swamps in support of a chemical coal model. In: Carboniferous Coal Short
Course and Guidebook, Geol. Geog. Dept. W. Virginia U., Morgantown, W. Va., 57-101.
Rettke, R.C.. 1981. Probably burial diagenetic and provenance effects on Dakota Group clay mineralogy,
Denver Basin. Jour. Sed. Petrol., 51: 541-551.
Rex, R.W., 1966. Authigenic kaolinite and mica as evidence for phase equilibria at low temperatures.
Clays Clay Miner., 13: 95-104.
Rex, R.W., 1967. Authigenic silicates formed from basaltic glass by more than 60 million years contact
with sea water, Sylvania Guyot, Marshall Islands. Clays Clay Miner., 5: 195-203.
Rex, R.W., 1970. X-ray mineralogy studies - Leg 2. In: M.N.A. Peterson, et al., Initial Reports of the
DSDP, 2, Wash. (U.S. Gov. Printing Office), 329-346.
Rex, R.W. and Goldberg, E.D., 1958. Quartz content of pelagic sediments of the Pacific Ocean. Tellus,
10: 153-159.
Rex, R.W. and Martin, B.D., 1966. Clay mineral formation in sea water by submarine weathering of
K-feldspar. Clays Clay Miner, 14: 235-240.
Reynolds, R.C., 1963. Potassium-rubidium ratios and polymorphism in illite and microclines from the
clay size fractions of Proterozoic carbonate rocks. Geochim. Cosmochim. Acta. 27: 1097-1 112..
Reynolds, R.C., 1965a. The concentration of boron in Precambrian seas. Geochim. Cocmochim. Acta,
29: 1-16.
Reynolds, R.C., 1965b. Geochemical behavior of boron during the metamorphism of carbonate rocks.
Geochim. Cosmochim. Acta, 29: 1101-1114.
Reynolds, R.C., 1967. Interstratified clay systems: Calculations of the total one-dimensional diffraction.
Amer. Miner. 52: 661-662.
Reynolds, R.C., 1971. Clay mineral formation in an alpine environment. Clays Clay miner., 19: 361-374.
Reynolds, R.C., 1980. Interstratified clay minerals. In: G.W. Brindley and G. Brown (Editors), Crystla
Structures of Clay minerals and Their X-ray Identification. Mineralogical Soc. London, 249-303.
Reynolds, W.R., 1966. Formation of cristobalite, zeolite and clay minerals in the Paleocene and Lower
Eocene of Alabama. Ph.D. Thesis, Florida State U., Tallahassee, 194 p.
770

Reynolds, R.C. Jr. and Hower, J., 1970. The nature of interlayering in mixed-layer illite-montmorillonite.
Clays Clay Miner., 18: 25-36.
Rhoads. D.C.. 1963. Rates of sediment reworking by Yoldia limatula in Buzzards Bay, Massachusetts.
and Long Island Sound. Jour. Sed. Petrology. 33: 723-727.
Rice, D.D. and Gautier, D.L.. 1983. Patterns of sedimentation, diagenesis, and hydrocarbon accumula-
tion. In: Cretaceous Rocks of the Rockey Mountains, Short Course No. 11, Soc. Econ. Paleo. Miner.,
Chap. 2.
Rich. C.I., 1956. Muscovite weathering in a soil developed in the Virginia Piedmont. Clays Clay Miner..
5 : 203-212.
k c h , C.I., 1964. Effect of cation size and pH on potassium exchange in Nason soil. Soil. Sci.. 98:
100-105.
Rich, C.I., 1968. Hydroxy interlayers in exapansible layer silicates. Clays Clay Miner.. 16: 15-30.
Rich, C.I. and Cook. M.G.. 1963. Formation of dioctahedral vermiculite in Virginia soils. Clays Clay
Miner. 10: 96-106.
Riddick, T.M., 1968. Control of colloid stability through Zeta potential. T.M. Riddick, N.Y.. 372 p.
Rieke. H.H. and Chilingarian, G.V., 1974. Compaction of Argillaceous Sediments. Developments in
Sedimentology 16, Elsevier. 424 p.
Riley, J.P. and Chester, R.. 1971. Introduction to Marine Chemistry, Academic Press. 465 p.
Rimmer, S.M. and Eberl, D.D., 1982. Origin of an underclay as revealed by vertical variations in
mineralogy and chemistry. Clays Clay Miner.. 30: 422-430.
Risebrough, R.. Huggett, R.. Griffin, J.. and Goldberg, E.D.. 1968. Pesticides: transatlantic movement in
the Northeast trades. Science, 159: 1233-1236.
Roaldset. E., 1985. Tertiary claystones North Sea - Barents Sea, composition and technical properties.
Internat. Clay Conf.. Denver, Abs., 197.
Roberson. H.E., 1964. Petrology of Tertiary bentonites of Texas. Jour. Sed. Petrol.. 34: 401-411.
Roberson. H.E., 1957. Petrology of the Carrizo and Recklow Formation, Leon County. Texas. M.A.
Thesis, U. Texas, Austin, 144 p.
Robert, C., 1980. Climats et courants CCnozoiques dans I’Atlantique Sud d’apr6s I’Ctude des minCraux
argileux (Legs 3, 39 et 40DSDP). Oceanol. Acta, 3: 369-376.
Robert. C.. 1981. Santonian to Eocene palaeographic evolution of the Rio Grand Rise (South Atlantic)
deduced from clay mineralogical data (DSDP Legs 3 and 39). Palaeogeo.. Palaeoclim.. Palaeoecol..
33: 311-325.
Robert. M., 1973. The experimental transformation of mica toward smectite; relative importance of total
charge and tetrahedral substitution. Clays Clay Miner., 21: 167-174.
Roberson, H.E.. 1974. Early diagenesis: expansible soil clay-sea water reactions. Jour. Sed. Petrol., 44:
441 -449.
Roberson. H.E. and Lahamm, R., 1981. Smectite to illite conversion rates: effects of solution chemistry.
Clays Clay Miner., 29: 129-135.
Roberts. D.G. and Montadert, L.. 1979. Margin paleoenvironments of the Northeast Atlantic. In: L.
Montadert and D.G. Roberts et al. (Editors). Init. Repts. DSDP. 48, Wash. (US.Gov. Printing
Office), 1099-1 11 8.
Robertson, R.H.S., 1963. Allophanic soil from Trail Bridge, Oregon, with notes o n mosaic growth in clay
minerals. Clay Miner. Bull., 5 : 237-247.
Robinson, P.T.. Flower, M.F.J., Schminke, H-U., and Ohnmacht, W.. 1977. Low temperature alteration
of oceanic basalt, DSDP Leg 37. In: F. Aumento, W.G. Nelson et al.. Initial Report of the Deep Sea
Drilling Project. 37: Wash. (U.S. Gov. Printing Office), 775-793.
Rochewicz, A. and Bakun, N.N.. 1980. Secondary minerals in the Rotliegend sandstones of western
Poland. Archiwun Miner., 36: 47-54.
Roen. J.B. and Hosterman, J.W.. 1982. Misuse of the term “bentonite” for ash beds of Devonian age in
the Appalachian basin. Geol. Soc. Amer. Bull., 93: 921-925.
Rogers, L.E., Martin, A.E., and Norrish. K.. 1954. Palygorskite from Queensland: Miner. Mag.. 30:
534-540.
Rogers, L.E., Quirk, J.P.. and Norrish. K., 1956. Aluminium sepiolite. Jour. Soil. Sci.. 7: 177-183.
Rohrlich, V., Price, N.B. and Calvert, S.E., 1969. Chamosite in recent sediments of Loch Etive, Scotland.
Jour. Sed. Petrol., 39: 624-631.
771

Rolfe, B.N. and Hadlety, R.F., 1964. Weathering and transport of sediment in the Cheyenne River Basin,
eastern Wyoming. Clays Clay Miner., 12 Natl. Conf. Atlanta. GA., 649-670.
Ronov, A.B., 1972. Evolution of rock composition and geochemical processes in the sedimentary shell of
the earth. Sedimentology, 19: 157-172.
Ronov. A.B.. Khain, V.E., Balukhovsky, A.N. and Seslavinsky, K.B., 1980. Quantitative analysis of
Phanerozoic sedimentation. Sed. Geol., 25: 311-325.
Ronov, A.B. and Migdisov, A.A., 1971. Geochemical history of the crystalline basement and the
sedimentary cover of the Russian and North American Platforms. Sedimentology, 16: 137-185.
Roquemore, S.K., Reynolds, W.R. and Williford, C.W., Jr., 1984. Models of clinoptiolite occurrence in
Middle Eocene strata of Alabama and Mississippi. Clay Miner. SOC.Meeting, Baton Rouge, Abs., 98.
Ross, C.A., 1986. Paleozoic evolution of southern margin of Permian basin. Geol. Soc. Amer. Bull., 97:
536-554.
Ross, C.P., 1976. The Precambrian of the United States of America: Northwestern United States - The
Belt Series. In: K. Rankama (Editor), The Precambrian, 4: 145-252.
Ross, C.S., 1946. Sauconite - a clay mineral of the montmorillonite group. Amer. Miner., 31: 411.
Ross, C.S. and Hendricks, S.B., 1945. Minerals of the montmorillonite group. U.S. Geol. Surv. Prof. Pap.
205-B, 79 p.
Ross, C.S. and Kerr, P.F., 1931. The kaolin minerals. U.S. Geol. Surv. Prof. Paper 165E, 151-175.
Ross, C.S. and Kerr, P.F., 1934. Halloysite and allophane. U.S. Geol. Surv. Prof. Paper 185-G, 135-148.
Ross, C.S., Miser, H.D. and Stephenson, L.W., 1928. Water-laid volcanic rocks of early Upper
Cretaceous age in southwestern Arkansas, southeastern Oklahoma, and northeastern Texas. U.S.
Geol. Sur. Prof. Paper 154, 175-202.
Ross, G.J., 1969. Acid dissolution of chlorites: Release of magnesium, iron and aluminum and mode of
acid attack. Clays Clay Miner., 17: 347-354.
Ross, G.J. and Kodama, H., 1974. Experimental transformation of a chlorite into a vermiculite. Clays
Clay Miner, 22: 205-211.
Ross, G.J. and Kodama, H., 1976. Experimental alteration of a chlorite into a regularly interstratified
chlorite-vermiculite by chemical oxidation. Clays Clay Miner., 24: 183-190.
Ross, G.J., Wang, C. and Protz, R., 1983. Mineralogy and mineral weathering of selected post-Wiscon-
sian and pre-Wisconsian soils and saprolites in eastern Canada. Clay Miner. Soc. Meeting. Buffalo,
Abs., 12.
Rossel, N.C., 1982. Clay mineral diagenesis in Rotliegend aeolian sandstones of the Southern North Sea.
Clay Miner, 17: 69-77.
Roth, C.B., Jackson, M.L., and Syers, J.K., 1969. Deferration effect o n structural ferrous-ferric iron ratio
and CEC of vermiculites and soils. Clays Clay MIner., 17: 253-264.
Rowsell, D.M. and De Swardt, A.M.]., 1976. Diagenesis in Cape and Karroo sediments, South Africa,
and its bearing on their hydrocarbon potential. Trans. Geol. Soc. S. Afr., 79: 81-144.
Roy, A.B., 1978. Evolution of slaty cleavage in relation to diagenesis and metamorphism: A study from
the Hunsruckschiefer. Geol. Soc. Amer. Bull, 89: 1775-1785.
Roy, B.B. and Barde, N.K., 1962. Some characteristics of the black soils of India. Soil Sci., 93: 142-147.
Rule, A.C. and Bailey, S.W., 1985. Refinement of the crystal structure of phengite - 2M. Clays Clay
Miner., 33: 403-409.
Rumeau, J.L. and Kulbicki, G.L., 1966. Evolution des minCaraux argileux dans les dolomies et les
calcaires poreux du Cretace SupCrieur de la plateforme d’Aquitaine. Proc. Internatl. Clay Conf.,
Jerusalem 2: 103-117.
Rusinov, V.L., Laputina, I.P., Muravitskaja, G.N., Zvjagin, B.B. and Gradusov, B.P., 1979. Clay minerals
in basalt from Deep Sea Drilling Project Sites 417 and 418. In: T. Donnelly, J. Franchateau, W.
Bryan, P. Robinson, M. Flower, and M. Salisburg et al. (Editors), Initial Reports of the deep Sea
Drilling Project 51, 52, 53 Part 2: Wash. (U.S. Gov. Printing Office), 1265-1271.
Russell, K.L., 1970. Geochemistry and halmyrolysis of clay minerals, Rio Ameca, Mexico. Geochim.
Gosmochem. Acta, 34: 893-907.
Rutstein, M.S., Gronwald, K., and Conrad, D., 1983. lllite crystallinity studies of Mid-Hudson River
Valley Ordovician shales. In: R. Waines (Editor), Field Trip Guidebook, Geol. SOC.Amer., N E Sec.
Ann. Meet., Kiamesha Lake, N.Y., 47-64.
772

Ruxton. B.P.. 1968. Measure of the degree of chemical weathering of rocks. Jour. Geol. 76: 518-527.
Ryer, T.A., Phillips, R.E., Bohor, B.F. and Pollastro. R.M., 1980. Use of altered volcanic ash falls in
stratigraphic studies of coal-bearing sequences: An example from the Upper Cretaceous Ferron
Sandstone Member of the Mancos Shale in central Utah. Geol. Soc. Amer. Bull.. 91: 579-586.
Saburo, A. and Kaoru. 0.. 1973. The clay minerals in the sediments of the Continental Shelf off Sanin,
the Japan Sea. Jour. Assoc. Geol. Collaboration in Japan, 27: 35-59.
Sand, L.B., 1956. On the genesis of residual kaolins. Amer. Miner., 41: 28-40.
Sarkisyan, S.G., 1972. Origin of authigenic clay minerals and their significance in petroleum geology. Sed.
Geol.. 7: 1-22.
Sarkisyan, S.G. and Kotelnikov, D.D., 1972. Genesis and thermodynamic stability of dioctahedral and
trioctahedral mixed-layer minerals in sedimentary rocks. Proc. Int. Clay Conf., Madrid, 281 -290.
Sarnthein, M. and Kapman, B., 1980. Late Quaternary deep-sea record on northwest Africa dust supply
and wind circulation. In: Sahara and Surrounding Seas: Sediments and Climatic Changes:
Palaeoecology of Africa and The Surrounding Islands, 12, Amsterdam. Balkema, 239-253.
Sawhney, B.L., 1969. Regularity of interstratification as affected by charge density in layer silicates. Soil
Sci. Soc. Amer. Proc., 33: 42-46.
Sawhney. B.L. and Jackson. M.L.. 1958. Soil montmorillonite formulas. Soil Sci. SOC.Amer. Proc. 22, 2:
115-118.
Sayles, F.L. and Bischoff, J.L.. 1973. Ferromanganoan sediments in the equatorial east Pacific. Earth
Planet. Sci. Letters, 19: 330-336.
Sayles, F.L. and Mangelsdorf, P.C., Jr., 1977. The equilibration of clay minerals with seawater: exchange
reactions. Geochim. Cosmochem. Acta, 41 : 951-960.
Sayles, F.L. and Mangelsdorf. P.C. Jr., 1979. Cation-exchange characteristics of Amazon River sus-
pended sediment and its reaction with seawater. Geochim. Cosmochim. Acta. 43: 767-779.
Schatz. A., 1963. Soil microorganisms and soil chelation. The pedogenic action of lichens and lichen
acids. Agri. Food Chem., 11: 112-118.
Schlenker. B., 1971. Petrographische Untersuchungen an Gipskeuper und Lettenkeuper von Stuttgart.
Oberrhein. Geol. Abh. 20: 69-102.
Scheidegger, K.F. and Stakes, D.S., 1979. X-ray diffraction and chemical study of secondary minerals
from Deep Sea Drilling Project Leg 51 Holes 417A and 417D. In: T. Donnelly. J. Franchateau. W.
Bryan, P. Robinson, M. Flower, and M. Salisbury et al., Initial Reports of the Deep Sea Drilling
Project 51, 52, 53 Part 2: Wash. ( U S . Gov. Printing Office), 1253-1263.
Schlocker. J. and VanHorn. R., 1958. Alteration of volcanic ash near Denver, Colorado. Jour. Sed.
Petrol., 28: 31-35.
Schneider. H., 1927. A study of Glauconite. Jour. Geology, 35: 289-310.
Schoen, R., 1964. Clay minerals of the Silurian Clinton ironstones, New York State. Jour. Sed. Petrol.,
34: 855-863.
Schofield. R.K. and Samson, H.R.. 1953. The deflocculation of kaolinite suspension and the accompany-
ing change over from positive to negative chlorite adsorption. Clay Miner., 2: 45-50.
Schrader, E.L., Rosendahl. B.R.. Furbish, W.J. and Mattey, D.P., 1980. Mineralogy and geochemistry o f
hydrothermal and pelagic sediments from the mounds hydrothermal field, Galapagos Spreading
Center DSDP leg 54. Jour. Sed. Petrol.. 50: 917-927.
Schrock, R.L., 1975. Mixed-layered clays of the Triassic Chinle Formation. northern Arizonea and
southern Utah. M.A. Thesis, Dartmouth College, 79 p.
Schroeder. R.J. and Hays, J.B., 1968. Dickite and kaolinite in Pennsylvanian limestones of southeastern
Kansas. Clays Clay Miner., 16: 41-49.
Schubel, J.R., 1969. Size distribution of the suspended particles of the Chesapeake Bay turbidity
maximum. Neth. Jour. Sea. Res.. 3: 283-309.
Schultz, L.G., 1958. Petrology of underclays. Geol. SOC.Amer. Bull., 69: 363-402.
Schultz, L.G., 1963a. Nonrnontmorillonitic composition of some bentonite beds. Clays Clay Miner., 11 :
169-177.
Schultz. L.G., 1963b. Clay minerals in Triassic Rocks of the Colorado Plateau. U S . Geol. Survey Bull.
1147-C. 71 p.
Schultz, L.G., 1969. Lithium and potassium absorption, dehydroxylation temperature and structural
water content of aluminous smectites. Clays Clay Min., 17: 115-149.
773

Schultz, L.G.. 1984. Comparison of illite-smectite and maturation of organic matter in Cretaceous shale,
Montana disturbed belt. Clay Miner. Soc. Meeting, Baton Rouge, Abs., 103.
Schultz, L.G., Shepard. A.O., Blackmon, P.D. and Starkey, H.C., 1971. Mixed-layer kaolinite-montmoril-
lonite from the Yucatan Peninsula. Clays Clay Miner., 19: 137-150.
Schultz, L.G., Tourtelot, H.A., Gill, J.R., and Boerngen, J.G., 1980. Composition and properties of the
Pierre Shale and equivalent rocks, Northern Great Plains Region. U.S. Geol. Surv. Prof. Paper
1064-B, 113 p.
Schultz, L.G. and Wright, J.C., 1963. Bentonite beds of unusual composition in The Carmel Formation,
southwest Utah. U.S. Geol. Surv. Prof. Paper, 450-E: 67-72.
Schutter, S.R. and Heckel, P.H., 1983. Missourian (early late Pennsylvanian) climate in midcontinent
North America. Geol. Soc. Amer. Bull, Prog. with Abs.. 681.
Scotford, D.M., 1965. Petrology of the Cincinnatian Series shales and environmental implications. Geol.
SOC.Amer. Bull., 76: 193-222.
Scott, T.M., 1973. The clay mineralogy of the Silurian Crab Orchard Formation of east-central Kentucky.
Eastern Kentucky U., M.S., Thesis, 57 p.
Scruton, P.C., 1956. Oceanography of Mississippi Delta sedimentary environments. Amer. Assoc. Petrol.
Geol. Bull., 40: 2864-2952.
Scruton, P.C., and Moore, D.G., 1953. Distribution of surface turbidity of the Mississippi Delta. Amer.
Assoc. Petrol. Geol. Bull., 137: 1067-1074.
Sears, D.A., 1976. The fissility of some Carboniferous shales. Sedimentology, 23: 721 -725.
Sedivy, R.A., Wampler. J.M. and Weaver, C.E., 1984. Potassium-Argon. In: C.E. Weaver and Assoc.,
Shale-Slate Metamorphism in Southern Appalachians. Elsevier, 153-183.
Sellwood, B.W. and Sladen, C.P., 1981. Mesozoic and Tertiary argillaceous units: distribution and
composition. Q.J. Eng. Geol. London, 14: 263-275.
Senkayi, A.L., Dixon, J.B., Hossner, L.R., Abder-Ruhman, M., and Fanning, D.S., 1984. Mineralogy and
genetic relationships of tonstsein, bentonite. and lignitic strata in the Eocene Yegua Formation of
east-central Texas. Clays Clay Miner., 32: 259-271.
Serstevens, A.T., Rouxhet, P.G. and Herbillon, A.J.., 1978. Alteration of mica surfaces by water
solutions. Clay Miner., 13: 4 0 - 4 1 0 .
Seyfied, W.G., Shanks, W.C. and Bischoff, J.L., 1976. Alteration and vein formation in Site 321 basalt.
In: Yeats, R.S., Hart, S.R. et al., Initial Report of the Deep Sea Drilling Project, 34, Wash. ( U S . Gov.
Printing Office), 385-392.
Shadfan, H. and Dixon, J.B., 1984. Occurrence of palygorskite in the soils and rocks of the Jordan
Valley. In: A. Singer and E. Galen (Editors), Palygorskite-Sepiolite, Developments in Sedimentology
37. Elsevier, Amsterdam, 187-198.
Shannon, P.M., 1978. The petrology of some Lower Paleozoic graywackes from southeast Ireland: a clue
to the origin of the matrix. Jour. Sed. Petrol., 48, 1185-1192.
Shaw, D.B. and Weaver, C.E., 1965. The mineralogical composition of shales. Jour. Sed. Petrol., 35:
213-222.
Shaw, D.M., 1956. Geochemistry of pelitic rocks. Part 111: Major elements and general geochemistry.
Geol. Soc. Amer. Bull., 67: 919-934.
Shaw, H.F., 1981. Mineralogy and petrology of the argillaceous sedimentary rocks of the U.K. Q.J. Eng.
Geol. London, 14: 277-290.
Shayan, A., 1984. Hisingerite material from a basalt quarry near Geelong, Victoria, Australia. Clays Clay
Miner., 32: 272-278.
Shelton, J.W., 1964. Authigenic kaolinite in sandstone. Jour. Sed. Petrol., 34: 102-11 1.
Shepard, F., P., 1950. Marginal sediments of Mississippi Delta. Amer. Assoc. Petroleum Geol., 40:
2537-2623.
Shepard, F.P., 1954. Nomenclature based on sand-silt-clay ratios. Jour. Sed. Petrol., 24: 151-158.
Shepard, F.P., 1959. The Earth Beneath the Sea. Johns Hopkins Press, Baltimore, 348 p.
Shepard, F.P. and Moore, D.G.. 1955. Central Texas coast sedimentation: characteristics of sedimentary
enviornment. recent history and diagenesis. Bull. Amer. Assoc. Petrol. Geol.. 89: 1463-1593.
Sheppard, R.A. and Gude, A.J., 111, 1973. Zeolites and associated authigenic silicate minerals in
tuffaceous rocks of the Big Sandy Formation, Mohave County, Arizona. U.S. Geol. Sur. Prof. Paper
830, 36 p.
774

Sherman, G.D., 1952. The genesis and morphology of the alumina-rich laterite clays. In: Problems of
Clay and Laterite Genesis. Amer. Inst. Min. Metall. Engr., New York, 154-161.
Sherman, I., 1953. Flocculent structure of sediment suspended in Lake Mead. Trans. Amer. Geoph.
Union, 34: 394-406.
Shimoda, S., 1969. New data for tosudite. Clays Clay Miner., 17: 179-184.
Shirozu, H., 1974. Clay minerals in altered wall rocks of the Kuroko-type deposits. Mining Geol. Special
Issue, No. 6, 303-310.
Shirozu, H. and Higashi, S., 1976. Structural investigations of sudoite and regularly interstratified
sericite/sudoite. Miner. Jour. Sapporo. 8: 158-170.
Sholkovitz, E.R., 1976. Flocculation of dissolved organic and inorganic matter during the mixing of river
water and seawater. Geochm. Cosmochim. Acta, 40: 831-845.
Shover, E.F., 1935. Petrology of Upper Paleozic clays and shales of north central Texas. Ph.D.
Dissertation, U. Illinois, 186 p.
Shover, E.F., 1964. Clay-mineral environmental relationships is Cisco (U. Penn.) clays and shales, north
central Texas. Clays Clay Miner., 12: 431-443.
Shrock, R.R., 1948. A classification of sedimentary rocks. Jour. Geol., 56: 118-129.
Shutov, V.D., Aleksandrova, A.V. and Losievskaya, S.A.. 1970. Genetic interpretation of the polymor-
phism of the kaolinite group in sedimentary rocks. Sedimentology, 15: 69-82.
Shutov, V.D., Drits, V.A. and Sakharov, B.A., 1969. On the mechanism of a postsedimentary transforma-
tion of montmorillonite into hydromica. Proc. Intern. Clay Conf., Tokyo, 1969, 523-531.
Shutov, V.D., Katz, M. Ya., Drits, V.A., Sokolova, A.L. and Kazakov, G.A., 1972. Crystallochemical
heterogeneity of glauconite as depending on the conditions of its formation and postsedimentary
changes. Proc. Inter. Clay Conf., Madrid, 269-279.
Sieffermann, G . and Millot, G.. 1968. Halloysite in the young soils above the recent basalts of Central
Cameroon. Bull. Groupe Fr. Argiles, 20: 25-38.
Siever. R. and Kastner. M., 1967. Mineralogy and petrology of some Mid-Atlantic b d g e sediments. Jour.
Mar. Res., 25: 263-278.
Siffert, B., 1962. Quelques rCactions de la s i l k en solution: la formation des argiles. Mtm. Serv. Carte
GCol. Alsace-Lorraine 21, 100 p.
Siffert, B., 1967. Some reactions of silica in solution: Formation of clay. Israel Programs for Scientific
Translations, Jerusalem, 100 p.
Siffert, B., 1978. Genesis and synthesis of clays and clay minerals: recent developments and future
prospects. Proc. Intern. Clay Conf., Oxford, 337-347.
Sigl, W., Chamley, H., Fabricius, F., d’Argoud, G.G. and Muller, J., 1978. Sedimentology and environ-
mental condtions of sapropels. In: K.J. Hsu, L. Montadert et al. (Editors), Initial Reports of the
DSDP, V. 42, Part 1: Wash. (U.S. Gov. Printing Office), 445-465.
Silverman, M.P. and Ehrlich, H.L., 1964. Microbial formation and degradation of minerals. In: W.W.
Umbreit (Editor), Adv. Applied Microbio. Academic Press, N.Y., 153-206.
Silverman, M.P. and Munoz, E.F.. 1970. Fungal attack on rock: solubilization and altered infared
spectra. Science, 169: 985-987.
Simons, L.H. and Taggard, M.S., Jr., 1954. Clay mineral content of Gulf Coast outcrop samples Natl.
Acad. Sci. - Natl. Res. Council Pub. 327: 104-110.
Sindeli?, J. and Kraus. I., 1976. New data o n the clays from the Cheb (Bohemia) and Poltir (Slovakia)
areas. Seventh Conf. Clay Miner. Petrol., Kalov Vary. 327-337.
Sinex, S.A., 1975. A chemical and mineralogical study of the saprolites of the Blue Ridge of Ashe
County, North Carolina. M.S. Thesis, Miami U., Oxford, OH., 55 p.
Singer, A,, 1979. Palygorskite in sediments: detrital, diagenetic or neoformed - a critical review. Geol.
Rdsch., 68: 996-1008.
Singer. A., 1980. The paleoclimatic interpretation of clay minerals in soils and weathering profiles.
Earth-Sci. Rev., 15: 303-326.
Singer. A., 1984. Pedogenic palygorskite in the arid environment. In: A. Singer. E. Galan (Editors),
Palygorskite-Sepiolite, Developments in Sedimentology 37. Elsevier, 169-175.
Singer, A., 1987. The clay mineralogy of aridic soils in the Near East. Clay Miner. SOC. Prog. with Ahs.,
118.
775

Singer, A,, Gal, M. and Banin, A., 1972. Clay minerals in recent sediments of Lake Kinneret (Tiberias),
Israel. Sed. Geol., 8: 289-308.
Singh, G. and Krishna Murti, G.S. R., 1975. Charge distribution in smectites developed from basalts in
Madhya Pradesh. Jour. Indian SOC.Soil Sci., 23: 177-183.
Singer, A. and Norrish, K., 1974. Pedogenic palygorskite occurrences in Australia. Amer. Miner., 59:
508-517.
Singer, A. and Miiller, G., 1983. Diagenesis in argillaceous sediments. In: G. Larson and G.V. Chilingar
(Editors), Diagenesis in Sediments and Sedimentary Rocks, 2. Elsevier, 115-212.
Singer, A,, Stoffers, P., Heller-Kallai, L. and Szafranek, D., 1984. Nontronite in a deep-sea core from the
South Pacific. Clays Clay Miner., 32: 375-383.
Sittler, C., 1965. Le PalCoghe des fossCs R h h a n et Rhodanien. Estudes ~Cdimentologiqueset paltoc-
limatiques. Mtm. Sew. Carte Geol. Als. Lorr., 24, Strasbourg, 350 p.
Sittler, C., Trauth, N. and Menillet, F., 1978. Bassin de Paris; mineraux argileux in Project 124. In: R.
Vinken (Editor), The Northwest European Tertiary Basin, 3: 9.
Skornyakova, N.S., Kurnosov, V.B. and Shevchenko, A. Ya., 1979. Saponite and vermiculite in sedimen-
tary rocks of the East Indian Ridge. Reports Acad. Sci. USSR, 24: 185-188.
Slaughter, M. and Earley, J.W., 1965. Mineralogy and geological significance of the Mowry Bentonites,
Wyoming. Geol. SOC.Amer. Special Pap. 83, 116 p.
Slaughter, M., and Milne, I.H., 1958. The formation of chlorite-like structures from montmorillonite.
Clays Clay Miner., 7: 114-24.
Sly, P.G., 1978. Sedimentary processes in lakes. In: A. Lerman (Editor), Lakes Chemistry Geology
Physics. Springer-Verlag, Heidelberg, 65-89.
Smith, D.G.W., 1967. The petrology and mineralogy of some Lower Devonian bentonites from Gaspe,
P.Q. Can. Mineral, 9: 141-165.
Smith, J.V. and Yoder, H.S., 1956. Studies of mica polymorphs. Miner. Mag., 31: 209-235.
Smith, N.D. and Syvitski, J.M., 1982. Sedimentation in a glacier-fed lake: the role of pelletization on
deposition of fine-grained suspensates. Jour. Sed. Petrol., 52: 503-513.
Smith, W.H. and OBrien, N.R., 1965. Middle and Late Pennsylvanian flint clays. Jour. Sed. Petrol, 35:
610-618.
Smith, W.W., 1959. Some interstratified clay minerals from basic igneous rocks. Miner. Mag., 32:
182-190.
Smoluchowski, M., 1917. Versuch einer mathematischen theorie der kaogulations - kinetik kolloider
losungen. Z. Phys. Chem., 92: 129-168.
Smulikowski, K.. 1954. The problem of glauconite. Polska. Akad. Nauk, Kom. Geol. Arch. Mineral.
Warsz., 18: 21-120. [In English.]
Snall, S., 1977. Silurian and Ordovician bentonites of Gotland (Sweden). Stockholm Contrib. Geol, 31 :
1-80.
Soil Survey Staff, 1960. Soil classification, a comprehensive (7th approximation). U.S. Dept. Agri., Soil
Cons. Service, 265 p.
Sorokin, V.I., Vlasov, V.V., Varfolomeeva, E.K. and Urasin, M.A., 1980. Effect of the medium on
development of glauconite composition. Lithol. Miner. Res., 14: 690-693.
Southard, R.J. and Boettinger, J.L., 1987. Occurrence and formation of smectite in three arid region soils.
Clay Miner. SOC.Prog. with Abs., 121.
Spangenberg, K., 1938. Pimelite. Naturwissenschaften, 26: 578-583.
Spears, D.A. and Kanaris-Sotiriou, R., 1979. A geochemical and mineralogical investigation of some
British and European tonsteins. Sedimentology, 26, 407-425.
Speights, D.B. and Brunton, G., 1961. Clay-mineral distribution in Permo-Pennsylvanian shales of Val
Verde Basin and Yates-Todd Arch, Texas. Amer. Assoc. Petrol. Geol. Bull., 45: 1957-1970.
Spooner, E.T.C. and Fyfe, W.S., 1973. Sub-sea-floor metamorphism, heat and mass transfer. Contr.
Mineral. Petrol, 42: 287-304.
Spronck, R., 1941. Mesures hydrographiques effectubes dans la region divagente du Bief Maritime du
Fleuve Congo. Brussels, Inst. Royale Colonial Belge Memoire, 156 p.
Spry, A,, 1968. Metamorphic Textures. Pergamon, New York, 350 p.
Spry, A,, 1969. Metamorphic Tectures. Pergamon, 350 p.
776

Spyridakis, D.E., Chesters, G . , and Wilde. S.A., 1967. Kaolinization of biotite as a result of coniferous
and deciduous seedling growth. Proc. Soil Sci. Soc. Amer., 31: 203-209.
Srodoh, J., 1976. Mixed-layer smectite/illite in the bentonites and tonsteins of the Upper Silesian Coal
Basin. Prace Miner, 49.
Srodoh. J., 1978. Correlation between coal and clay diagenesis in the Carboniferous of the Upper Silesian
Coal Basin. Proc. Internatl. Clay Conf., Oxford, Eng., 251-260.
Srodon, J., 1980. Precise identification of illite/smectite interstratifications by X-ray powder diffraction.
Clays Clay Miner., 28: 401-411.
Srodoh. J., 1984a. X-ray powder diffraction identification of illitic materials. Clays Clay Miner., 32:
337-349.
Srodoh, J., 1984b. Mixed-layer illite-smectite in low-temparature diagenesis: data from the Miocene of
the Carpathian Foredeep. Clay Minerals, 19: 205-215.
Srodoh, J., and Eberl, D.D., 1984. Illite. In: S.W. Bailey (Editor), Micas. Miner. SOC.Amer. Short Course
Notes, 495-544.
Srodoh, J., Morgan, D.J., Eslinger. E.V., Eberl, D.D., and Karlinger, M.R., 1986. Chemistry of
illite/smectite and end-member illite. Clays Clay Miner.,
Stach, E., Mackowsky, M-Th., Teichmiiller, M., Taylor, G.H.. Chandra, D.. and Teichmiiller, R., 1975.
Stach’s Textbook of Coal Petrology. Gebruder Borntraeger, Berlin, 428p.
Stanley, D.J., Addy, S.T. and Behrens, E.W., 1983. The mudline: variability of its position relative to
shelfbreak. In: D.J. Stanley and G.T. Mare (Editors). Critical Interface on Continental Margins. Soc.
Econ. Paleo Mineral, Spec. Publ. 33: 279-298.
Stanley, D.J. and Wear. C.M., 1978. The “mud-line” an erosion-deposition boundary on the upper
continental slope. Mar. Geol. 28: M19-M29.
Stanely. K.O. and Benson. L.V., 1979. Early diagenesis of High Plains Tertiary vitric and arkosic
sandstone, Wyoming and Nebraska. In: P.A. Scholle and P.R. Schluger (Editors), Aspects of
Diagenesis. SEPM Special Pub. No. 26, Tulsa, OK, 401-423.
S t a p h , F.L., 1969. Sedimentary organic matter, organic metamorphism and oil and gas occurrence. Bull.
Canad. Petrol. Geol., 17: 47-66.
Staub, J.R. and Cohen, A.D., 1978. Kaolinite-enrichment beneath coals; a modern analog. Snuggedy
Swamp, South Carolina. Jour. Sed. Petrol., 48: 203-210.
Staudigel, H. and Hart, S.R., 1983. Alteration of basaltic glass: Mechanisms and significance for the
oceanic crust-seawater budget. Geochim. Cosmochim. Acta, 47: 337-350.
Steiner. A., 1967. Clay minerals in hydrothermally altered rocks at Wairakei, New Zealand. Clays Clay
Miner. 16th Conf., Abs.. 31-33.
Stephen, I., 1954. Palygorskite from Shetland. Miner. Mag., 30: 471-480.
Sterne. E.J., Reynolds, R.C.. Jr. and Zantop. H., 1982. Natural amonium illites from black shales hosting
a stratiform base metal deposit. Delong Mountains, Northern Alaska. Clays Clay Miner., 30:
161-166.
StCvaux, J., 1962. Etude stdimentologique des formations Ordoviciennes et Mksozoiques inftrieures des
forages au S du champ d’Hassi-Messaoud (Sedimentological Study of Ordovician and Lower
Mesozoic Formations in the Boreholes Drilled to the South of the Field of Hassi-Messaoud Oil
Field). Thhse Sci., 3rd cycle, Strasbourg. 1961. Referenced by Lucas. 1962.
Stewart, F.H., 1965. The mineralogy of the British Permian evaporites. Min. Mag., 34: 460-470.
Stoch, L. and Sikora, W.. 1976. Transformation of micas in the process of kaolinitization of granites and
gneiss. Clays Clay Miner., 24: 156-162.
Stoffers, P. and Singer, A.. 1979. Clay minerals in Lake Mobutu Sese Seko (Lake Albert) ~ their
diagenetic changes as an indicator of the paleoclimate. Geol. Rundschau, 68: 1009-1024.
Stokke, P.R., Carson. B., and Baker. E.T., 1977. Comparison of the bottom nepheloid layer and late
Holocene deposition on Nitinat Fan: implications for lutite dispersal and deposition. Geol. Soc.
Amer. Bull., 88: 1586-1592.
Stone, W.J., 1973. Stratigraphy and sedimentary history of middle Cenezoic (Oligocene and Miocene)
deposits in North Dakota. Ph.D. Thesis, U. North Dakota, Grand Forks, 217 p.
St6rr. M., 1975a. Kaolin deposits of the G D R in the northern region of the Bohemian Massif.
Ernst-Moritz-Arndt-UniversitatGreifswald, 243 p.
777

Storr, M., 1975b. Distribution, age and genesis of the formation of the weathering crust in the GDR. In:
M. Storr, (Editor), Kaolin deposits of the GDR in the northern region of the Bohemian Massif.
Ernst-Moritz-Arndt-Universitat Greifswald, 29-92.
Stoessell, R.K. and Hay, R.L. 1978. The geochemical origin of sepiolite and kerolite at Amboseli, Kenya.
Contrib. Mineral. Petrol. 65: 255-267.
Strakhov. N.M., 1966. Principles of Lithogenesis: Vol. 2. Consultants Bureau and Oliver and Boyd,
Edinburgh, 609 p.
Strakhov, N.M., 1967. Principles of Lithogenesis, V . l . Oliver and Boyt, Ltd., 245 p.
Stuart, C.J. and Caughey, C.A., 1976. Form and composition of the Mississippi Fan. Trans. Gulf Coast
Assoc. Geol. Soc., 25: 333-343.
Studer, M., 1980. MCtamorphisme d’enfouissement dans le Hart Atlas central (Maroe). Essai sur
I’evolution de I’epassieur des couvertures skdimentaires. C.R. Acad. Sc. Paris, 291 : 457-460.
Suchecki, R.K., 1984. Clay-mineral diagenesis of the Tuscaloosa Sandstone: implications of hydrogen
isotopes. Clay Miner. SOC.Prog. with Abs., Baton Rouge, 113.
Suchecki, R.K., Perry, E.A., Jr., Hubert, J.F.. 1977. Clay petrology of Cambro-Ordovician continental
margin, Cow Head Klippe, Western Newfoundland. Clays Clay Mineral., 25: 163-170.
Sudo, T., 1943. On some low temperature hydrous silicates found in Japan. Bull. Chem. SOC.Japan, 18:
281-329.
Sudo, T., 1954. Iron-rich saponite found from Tertiary iron sand beds of Japan: re-examination of
“lembergite”. Jour. Geol. SOC.Japan, 60: 18-27.
Sudo, T. and Hayashi, H., 1956. A randomly interstratified kaolin-montmorillonite in acid clay deposits
in Japan. Nature, 178: 1115-1116.
Sudo, T., Hayashi, H. and Yokokura, H., 1957. Mineral associations in ore deposits. Clay Miner. Bull., 3:
258-263.
Sudo, T., and Sato, M., 1966. Dioctahedral chlorite. Proc. of the Int. Clay Conf., Israel, 33-40.
Sudo, T. and Shimoda, S., 1978. Clays and Clay Minerals of Japan. Elsevier, 326 p.
Sudo, T. and Yotsumoto, H., 1977. The formation of halloysite tubes from spherulitic halloysite. Clays
Clay Miner., 25: 155-159.
Suitch, P.R. and Young, R.A., 1983. Atom positions in highly ordered kaolinte. Clays Clay Miner., 31:
357-366.
Summer, K.V., 1976. The clay component of the Columbia River palagonites. Amer. Miner., 61:
492-494.
Summer, M.E., 1963. Effect of iron oxides on positive and negative charges in clays and soils. Clay
Miner., 5: 218-224.
Surdam, R.C. and Crossey, L.J., 1985. Mechanisms of organic/inorganic interactions in sandstone/shale
sequences. In: Relatinshiip of Organic Matter and Mineral Diagenesis, Schort Course No. 17. SOC.
&on. Paleo. Miner., Tulsa, 177-231.
Surdam, R.C. and Parker, R.D., 1972. Authigenic aluminosilicate minerals in the tuffaceous rocks of the
Green River Formation, Wyoming. Geol. SOC.Amer. Bull, 83: 689-700.
Sverdrup, H.U., 1947. Wind-driven currents in a baroclinic ocean; with application to the equatorial
currents of the eastern Pacific. Proc. Nat. Acad. Sci., 33-47.
Swain, F.M., 1966. Bottom sediments of Lake Nicaragua and Lake Managua, western Nicaragua. Jour.
Sed. Petrol., 36: 522-544.
Swift, D.J.P., 1973. Continental shelf sedimentation. In: C.A. Burk and C.L. Drake (Editors), Geology of
Continental Margins. Springer-Verlag, N.Y., 117-135.
Swindale, L.D. and Fan, P.F., 1967. Transformation of gibbsite to chlorite in ocean bottom sediments.
Science, 157: 799-800.
Swineford. A., 1955. Petrology of Upper Permian rocks in south-central Kansas. Kansas Geol. Surv. Bull.
111, 179 p.
Syvitski, J.P.M., Asprey, K.W., Clattenburg, D.A. and Hodge, G.D., 1985. The prodelta environment of a
fjord: suspended particle dynamics. Sedimentology, 32: 83-107.
Syvitski, J.P.M. and Lewis, A.G. 1980. Sediment ingestion by Tigriopus Californicus and other zooplank-
ton: mineral transformation and sedimentological considerations. Jour. Sed. Petrol., 50, 869-880.
Syvitski, J.P.M. and Murray, J.W. 1981. Particle interaction in fjord suspended sediment. Mar. Geol., 39:
215-242.
778

Tabikh, A.A., Barshad, I. and Overstreet, R., 1960. Cation exchange hysteresis in clay minerals. Soil Sci.,
90: 219-226.
Takahashi, J., 1939. Synopsis of glauconitization. In: P.D. Trask (Editor), Recent Marine Sediments.
Amer. Assoc. Petrol. Geol., Tulsa, Okla., 503-512.
Talibudeen, 0. and Goulding, K.W.T., 1983. Apparent charge heter-ogeneity in kaolins in relation to
their 2 : 1 phyllosilicate content. Clays Clay Miner., 31: 137-142.
Talwani, M., Udinstsev, G., et al., 1976. Site Report 336. Initial Reports of the Deep Sea Drilling Project,
38, Wash. ( U S . Gov. Printing Office), 23-116.
Tamura, T., 1958. Identification of clay minerals from acid soils: Jour. Soil Sci., 9: 141-147.
Tamura, T., Jackson, M.L. and Sherman, G.D., 1953. Mineral content of low humic, humic and hydro1
humic latosols of Hawaii. Soil Sci. Soc. Amer. Proc., 17: 343-346.
Tamura, T., Jackson, M.L. and Sherman, G.D., 1955. Mineral content of a latosolic brown forest soil and
a humic ferruginous latosol of Hawaii. Soil Sci. Soc. Amer. Proc., 20: 435-439.
Tan, K.H. and McCreery, R.A., 1975. Humic acid complex formation and intermicellar adsorption by
bentonite. Proc. Inter. Clay Conf., Mexico City, 629-641.
Tank, R.W., 1956. Clay mineralogy of the Morrison Formation, Black Hills area, Wyoming and South
Dakota. Amer. Assoc. Petroleum Geol. Bull.. 55: 1097-1114.
Tank, R.W., 1963. Clay mineralogy of some Lower Tertiary (Paleogene) sediments from Denmark. Geol.
Survey Denmark IV Series, 4, No. 9, 45p.
Tank, R.W., 1972. Clay minerals of the Green River Formation (Eocene) of Wyoming. Clay Minerals, 9:
297-308.
Tank, R.W. and McNeely, L., 1970. Clay mineraIs associated with the Precambrian Gowganda Forma-
tion of Ontario. Clay Min., 8: 471-477.
Tardy. Y., Bocquier, G., Paquet, H. and Millot, G., 1973. Formation of clay from granite and its
distribution in relation to climate and topography. Geoderma, 10: 271-284.
Tardy, Y., Cheverry, C., and Fritz, B., 1974. Ntoformation d'une argile magntsienne dans les dtpressions
interdunaires du lac Tchad: Application aux domaines de stabilitt des phyllosilicates alumineux
magnksiens et ferrifkres. C.R. Acad. Sci., Paris, SCr. D, 258: 1999-2002.
Taskey, R.D., Harward, M.E. and Youngberg, C.T., 1978. Relationship of clay mineralogy to landscape
stability. In: C.T. Youngberg (Editor), Forest Soils and Land Use, Proc. 5th North American Forest
Soils Conf., Fort Collins, Colorado, Colorado State U., Fort Collins, Colorado, 140-162.
Tazaki, K. and Fyfe, W.S., 1986. Primitive clay precursors formed on feldspar. Can. Jour. Earth Sci., In
press.
Tchukhrov, F.V. and Anosov, F.Y.. 1950. Medmontite. Mem. Soc. r u s e Minne., 79: 23-33.
Tedorovich, G.I., Chernov, A.A. and Kotel'nikov, D.D., 1967. Post-sedimentation alteration of clayey
material in sediments of the lower division of the Azerbaydzhan productive series. Dokl. Acad. Sci.
U.S.S.R., Earth Sci. Sect., 182: 152-155.
Teichmiiller, M., Teichmiiller, R. and Weber, K., 1979. Inkohlung und illit-kristallinitat (vergleichende
untersuchungen im Mesozoikum und Palaozoikum von WestFalen). Fortschr. Geol. Rheinl. Westfal.,
27: 201-276.
Teodorovich, G.I. and Konyukov, A.I., 1970. Mixed-layer minerals in sedimentary rocks as indicators of
the depth of their catagenetic alteration. Dokladay Akad. Nauk. SSSR, 191: 174-176.
Thayer, P.A., Hostettler, J. and Smith, S., 1974. Grain-size distribution of sediments from the Eastern
Indian Ocean: In J.J. Veevers, J.R. Heirtzler, et al. Initial Reports of the Deep Sea Drilling Project,
Vol. 27, Wash. (U.S. Gov. Printing Office) 507-522.
Theng, B.K.G., 1979. Formation and Properties of Clay-Polymer Complexes. Elsevier 362p.
Theng, B.K.G., Churchman, G.J., Whitton. J.S. and Claridge, G.G.C., 1984. Comparison of intercalation
methods for differentiating halloysite from kaolinite. Clays Clay Miner., 32: 249-258.
Thomas, R.L., Kemp, A.L.W. and Lewis, C.F.M., 1972. Distribution, composition and characteristics of
the surficial sediments of Lake Ontario. Jour. Sed. Petrol., 42: 66-84.
Thomas, R.L.. Kemp, A.L.W. and Lewis, C.F.M., 1973. The surficial sediments of Lake Huron. Can.
Jour. Earth Sci., 10: 226-271.
Thompson, A.M., 1970. Geochemistry of color genesis in red-bed sequence. Juniata and Bald Eagle
Formations, Pennsylvania Jour. Sed. Petrol., 40: 599-615.
779

Thompson, G., 1983. Basalt-seawater interaction. In: P.A. Rona, K. Bostrom, L. Laubier and K.L. Smith,
Jr. (Editors), Hydrothermal Processes at Seafloor Spreading Centers, Plenum Press, N.Y., 225-278.
Thompson, G. and Jennings, S., 1985. Kinetic factors affecting the diagenetic transformation of smectite
to illite. Inter. Clay Conf. Prog. with Abs., Denver, 239.
Thompson, G.R. and Hower, J., 1973. An explanation for low radio- metric ages from glauconite.
Geochim. Cosmochim. Acta., 37: 1473-1491.
Thompson, G.R. and Hower, J., 1975. The mineralogy of glauconite. Clays Clay Miner., 23: 289-300.
Thorez, J. and Bourguignon, P., 1973. Mintraux argileux des argiles de dissolution des calcaires
Dinantiens en Condroz. Extrait Annales SOC.Gtol. Belgique, 96: 59-85.
Thorez, J., and Pirlet, H., 1978. Petrology of K-bentonite beds in the carbonate series of the Visean and
Tournaisian stages of Belgium. Proc. Int. Clay Conf., Oxford, 323-332.
Thurman, H.V., 1981. Introductory Oceanography. Charles E. Merrill Pub. Co., Columbus, 464p.
Timofeev, P.P. and Bogolyubova, L.I., 1975. Facies and authigenic clay formation in peat bogs of the
Rioni intermontane trough. Septibme Congrb Inter. de Stratig. et de Gtol. du Carboniferous, Compte
Rendu, 292-306.
Timofeev, P.P., Eremeev, V.V., and Rateev, M.A., 1977. Palygorskite, sepiolite and other clay minerals in
Leg 41 ocean sediments: mineralogy, facies and genesis. In: Y. Lancelot, E. Seibold, et al., Initial
Reports of the DSDP 41, Wash. (US.Gov. Printing Office), 1087-1101.
Tissot, B.P., Demaison, G., Masson, P., Delteil, J.R. and Combaz, A,, 1980. Paleoenvironment and
petroleum potential of Middle Cretaceous black shales in Atlantic Basins. Amer. Assoc. Petrol. Geol.
Bull., 64: 2051-2063.
Tissot, B.P. and Welte, D.H., 1984. Petroleum Formation and Occurence. Springer-Verlag, Berlin, 699p.
Tomasson, J. and Kristmannsdottir, M., 1972. High temperature alteration minerals and thermal brines,
Rejkjanes, Iceland. Contr. Miner. Petrol., 36: 123-134.
Tomita, K. and Takahashi, H., 1985. Curves for the quantification of mica/smectite and chlorite/smec-
tite interstratifications by x-ray powder diffractrion. Clays Clay Miner., 33: 379-390.
Tomita, K. and Takahashi, H., 1986. Quantification curves for the x-ray powder diffraction analysis of
mixed-layer kaolinite/ smectite. Clays Clay Miner., 34: 323-329.
Tomlinson, C.W., 1916. The origin of red beds. Amer. Jour. Sci., 24: 153-179.
Tompkins, R.E., 1981. Scanning electron microscope of a regular chlorite/smectite (corrensite) from a
hydrocarbon reservoir sandstone. Clays Clay Mineral., 29: 233-235.
Tooker, E. W., 1960. Clay minerals in rocks of the lower part of The Oquirrh Formation, Utah. Clays
Clay Miner., 9: 355-364.
Tourtelot, H.A., 1960. Origin and use of the word “shale”. Amer. Jour. Sci., 258-A: 335-343.
Tourtelot, H.A., 1962. Preliminary investigation of the geologic setting and chemical composition of the
Pierre Shale Great Plains Region. U.S. Geol. Survey Prof. Paper 390, 74p.
Towe, K.M. and Grim, R.E., 1963. Variations in clay mineralogy, across facies boundaries in The Middle
Devonian (Ludlowville) New York. Amer. Jour. Sci., 261: 839-861.
Trauth, N., 1977. Argiles Cvaporitiques das la skdimentation carbonatte continentale et tpicontinental
Tertiaire. Sciences Gtol., U. Louis Pasteur de Strasbourg, Mkm. 49, 187p.
Trauth, N., Cavelier, C., Sommer, F., Tourencq, J., Pomerol, C. and Thiry, M., 1977. A p er p sur la
stdimentation PalCogine du synclinal de Bouxwiller comprise entre les marine i Rhynchonelles
(Bathonien) et le conglomkrat du Bastberg (Oligodne). Sci. Geol., Bull., 30: 91-100.
Trauth, N., Sommer, F. and Lucas, J., 1969. Evolution gkochimique d‘une strie stdimentaire Paltogtne
dans le Bassin de Paris. Bull. Serv. Carte Geol. Als. Lorr., 22: 279-310.
Triplehorn, D.M., 1966. Morphology, internal structure and origin of glauconite pellets. Sedimentology,
6: 247-266.
Triplehorn, D.M., 1967. Occurrence of pure, well-crystallized IM illite in Cambro-Ordovician sandstone
from Rhourde El Baquel Field, Algeria. Jour. Sed. Petrol., 37: 879-884.
Turekian, K.K., 1967. Estimates of the average Pacific deep-sea clay accumulation rate from material
balance calculations Progress in Oceanography, 4: 227-244.
Turekian. K.K., 1976. Oceans. Prentice-Hall, Englewood Cliffs, N.J., 149p.
Turekian, K.K. and Wedepohl, K.H., 1961. Distribution of the elements in some minor units of the
earth’s crust. Geol. Soc.Amer. Bull., 72: 175-192.
780

Turner, F.J., 1981. Metamorphic Petrology. McGraw-Hill, N.Y., 524p.


Turner, P., 1974. Origin of red beds in the Ringerike Group Group (Silurian) of Norway. Sed. Geol., 12:
215-235.
Turner, P. 1980. Continental Red Beds. Elsevier, 562p.
Twenhofel, W.H., 1936a. Terminology of the fine-grained mechanical sediments. Exhibit F - Report of
Committee on Sedimentation 1936-1937: Natl. Res. Council, Div. Geol. Geog.. 81-104.
Twenhofel, W.H., 1936b. The greensands of Wisconsin. Econ. Geol., 31,472-487.
Twenhofel, W.H., 1950. Principles of Sedimentation. McGraw-Hill Co., N.Y.. 673p.
Twiss, P.C. and Stindl, H., 1975. Clay mineralogy of Cottonwood Limestone. Internatl. Clay Conf. Abs.,
Mexico City, 301.
Vail, P.R., Mitchum, R.M. and Thompson S. 111, 1977. Seismic stratigraphy and global changes of sea
level. In: C.E. Payton (Editor), Seismic Stratigraphy-Application to Hydrocarbon Exploration, Amer.
Assoc. Petrol. Geol. Mem.26, Tulsa, 83-97.
Valeton, I., 1960. Vulkanishe Tuffitein lagerung in der nordwestdeutschen Oberkreide. Mitt. Geol. St
Inst. Hamb., 29: 26-41, from Bib. and Index of Geol. V 25, 1960, 560.
Van Andel, Tj.H., 1967. The Orinoco Delta. Jour. Sed. Petrol., 37: 297-310.
Van de Kamp, P.C., 1977. Inorganic and organic metamorphism in silicaclastic rocks. Amer. Assoc.
~

Petrol. Geol. Bull., 61, Abs., 729.


Van den Heuvel, R.C., 1966. The occurrance of sepiolite and attapulgite in the calcareous zone of a soil
near Las Cruces, New Mexico. Clays Clay Miner.: 13, 193-207.
Van der Marel, H.W., 1954. Potassium fixation in Dutch soils: Mineralogical Analyses: Soil Sci.. 78:
163-179.
Van der Marel, H.W., 1958. Quantitative analysis of kaolinite. Silicate Industriele, 25: 23-31, 76-86.
Van der Merwe, C.R. and Weber, H.W., 1963. The clay minerals of South African soils developed from
granite under different climatic conditions. S. Afr. Jour. Agric. Sci., 6: 411-454.
Van Houten, F.B., 1965. Composition of Triassic Lockatong and associated formations of Newark Group
central New Jersey and adjacent Pennsylvania. Amer. Jour. Sci., 263: 825-863.
Van Houten, F.B., 1972. Iron and clay in tropical savanna alluvium Northern Columbia: A contribution
to the origin of red beds. Geol. Soc. Arner. Bull. 83: 2761-2772.
Van Houten, F.B., 1973. Origin of red beds: a review - 1961-1972. Ann. Rev. Earth Planet. Sci., 1 :
39-42.
Van Houten, F.B. and Purucker, M.E., 1984. Glauconitic peloids and chamositic ooids-favorable factors,
constraints, and problems. Earth-Science Rev., 20: 21 1-243.
Van Olphen, H., 1963. An Introduction to Clay Colloid Chemistry. Interscience, 301p.
Van Stratten, L.M.J.U. and Kuenen, Ph.H., 1958. Tidal action as a cause of clay accumulation. Jour Sed.
Petrol., 28: 406-413.
Varsinyi, I., 1975. Clay minerals of the southern Great Hungarian Plain. Acta Mineral.-Petrog., Szeged,
22: 51-60.
Veith, J.A. and Jackson, M.L.. 1974. Iron oxidation and reduction effects on structural hydroxy and layer
charge in a aqueous suspensions of micaceous vermiculites. Clays Clay Miner.. 22: 345-353.
Velde, B., 1965a. Experimental determination of muscovite polymorphs stabilities. Amer. Miner., 50:
436-449.
Velde, B., 1965b. Phengite micas: synthesis, stability and natural occurrence. Amer. Jour. Sci., 263:
886-913.
Velde, B., 1977. Clays and Clay Minerals in Natural and Synthetic Systems. Elsevier, 218p.
Velde, B.. 1984. Polytypism and compositional variation as factors in the P-T stability of illite/smectite
mixed layered clays. Clay Miner. Soc. Prog. with Abs., Baton Rouge, 117.
Velde, B. and Brusewitz, A.M., 1982. Metasomatic and non-metasomatic low grade metamorphism of
Ordovician meta-bentonites in Sweden. Geochem. Cosmochim. Acta, 46: 447-452.
Velde, B. and Hower, J., 1963. Petrological significance of illite polymorphism in Paleozoic sedimentary
rocks. Amer. Min., 48: 1239-1254.
Velde, B. and Odin, G.S., 1975. Further information related to the origin of glauconite. Clays Clay
Miner., 23: 376-381.
781

Veniale, F., 1966. Sur la genese de la sepiolite dans certains sediments argileux Bigarres de L’Apennin
Septentrional (Italie du Nord). Proc. Inter. Clay Conf., Jerusalem, 2: 91.
Veniale, F. and Van der Marel, H.W., 1969. Identification of some 1 :1 regular interstratified trioc-
thahedral clay minerals. Proc. Inter. Clay Conf.. Tokyo, 233-244.
Venkatarathnam, K. and Ryan, W.B.F., 1971. Dispersal patterns of clay minerals in the sediments of the
eastern Mediterranean. Marine Geol., 11 : 261-282.
Vergo, N. and April, R.H., 1982. Interstratified clay minerals in contact aureoles West Rock, Connecti-
cut. Clays Clay Miner., 30: 237-240.
Venvey, J., 1952. On the ecology of distribution of cockle and mussel in the Dutch Waddensea, their role
in sedimentation and the source of their food supply, with a short review of the feeding behavior of
bivalve mollusks. Archives Nterlandaise de Zoologie, 10: 172-239.
Vicziin, I., 1975. A review of the clay mineralogy of Hungarian sedimentary rocks with special regard to
the distribution of diagenetic zones). Acta Geol. Acad. Sci. Hungaricae, 19: 243-256.
Vikulova, M.F., 1964. Effect of the origin of the Lower Carboni-ferous clays in the western part of the
Moscow Basin on the alterations of their clay minerals. In: I.M.J.U. VanStraaten (Editor), Deltaic
and Shallow Marine Deposition, Elsevier, 417-428.
Vinogradov, A.P. and Ronov, A.B., 1956. Evolution of the chemical composition of clays of the Russian
Platform. Geochem. 2: 123-139.
Vlodarskaya, V.R., 1962. Distribution of clay minerals in Paleozoic and Mesozoic rocks, in connection
with their origin. Inter-nat. Geol. Rev., 6: 1305-1309.
Von Bennekom, A.J. and Van der Gaast, S.J., 1976. Possible clay structures in frustules of living diatoms.
Geochim. Cos-mochim. Acta, 40: 1149-1152.
Von Englehardt, W., 1967. Interstitial solutions and diagenesis in sediments. In: G. Larson, and G.V.
Chilingar, (Editors). Diagenesis in Sediments: Developments in Sedimentology 8, Elsevier, 503-522.
WaagC, K.M., 1961. Stratigraphy and refractory clay rocks of the Dakota Group along the Northern
Front Range, Colorado. U.S. Geol. Sur. Bull. 1102, 154p.
Wada, K., 1977. Allophane and imogolite. In: J.B. Dixon, S.B. and Weed (Editors), Minerals in Soil
Environments, Soil Sci. SOC.Amer., Madison, Wisc., 603-638.
Wada, K., 1979. Structural formulas of allophanes. Proc. Inter. Clay Conf., Oxford, 537-545.
Wada, K., 1981. Amorphous clay minerals - chemical composition, crystalline state, synthesis and
surface properties. Proc. Inter. Clay Conf., Bologna, 385-398.
Wada, K. and Yoshinaga, N., 1969. The structure of imogolite. Amer. Miner., 54: 50-71.
Wada, K., Yoshinaga, N., Yotsumota, H., Ibe, K. and Aida, S., 1970. High resolution electron
micrographs of imogolite. Clay Miner., 8: 487-489.
Wahlstrom, E.E., 1966. Geochemistry and petrology of the Morrison Formation, Dillon, Colorado. Geol.
SOC.Amer. Bull., 77: 727-740.
Walker, G.F., 1949. The decomposition of biotite in the soil. Miner. Mag., 28: 693-703.
Walker, G.F., 1956. The mechanism of dehydration of Mg-vermiculite. Clays Clay Miner. Nat. Acad. Sci.
Nat. Res. Council Pub. 456, 101-115.
Walker, G.F., 1958. Reactions of expanding-lattice clay minerals with glycerol and ethylene glycol. Clay
Miner. Bull., 3: 302-312.
Walker, G.F., 1975. Vermiculites. In: J.E. Gieseking (Editor), Soil Component, V.2, Inorganic Compo-
nents, Springer-Verlag, 155-190.
Walker, T.R., 1976. Diagenetic origin of continental red beds. In: H. Falke, D. Reidel (Editors), The
Continental Permian in Central, West and South Europe, Dordrecht-Holland, 240-282.
Walker, T.R. and Honea, R.M., 1969. Iron content of modern deposits in the Sonoran Desert: A
contribution to the origin of red beds. Geol. SOC.Amer. Bull., 80: 535-544.
Walker, T.R., Ribbe, P.H. and Honea, R.M.. 1967. Geochemistry of hornblende alteration in Pliocene
red beds Baja California, Mexico. Geol. SOC.Amer. Bull., 78: 1055-1060.
Wall, G.J., Wilding, L.P. and Miller, R.H., 1974. Biological trans-formation of clay-sized sediments in
simulated aquatic environments. Inter. Conf. Great Lakes Res.. Proc. No. 17, 207-211.
Wallace, J.M. and Hobbs, P.V., 1977. Atmospheric Science. Academic Press, 467p.
Wampler, J.M. and Weaver, C.E., 1986. Sample analysis for K-Ar studies of Palo Duro Basin bedded
salts. Bendix Corp. Field Eng. Corp. Tech. Tpt.. 64p.
782

Wampler, J.M. and Weaver, C.E., 1987. Sample analysis for K-Ar studies of Palo Duro Basin bedded
salts. UNC Geotech. Tech. Rpt., 77 p.
Wanless, H.R., Baroffio, J.R., Gamble, J.R., Horne, J.C., Orlopp, D.R., Rocha-Campos, A,, Sauer. J.E.,
Trescott, P.C., Vail, R.S. and Wright, C.A., 1970. Late Paleozoic deltas in the central and eastern
United States: In: J.P. Morgan (Editor), Deltaic Sedimentation, Modern and Ancient, SOC. Econ.
Paleontologist and Mineralogist, Spec. Pub. 15. 215-245.
Waples. D.W., 1980. Time and temperature in petroleum formation application of Lopatins method to
petroleum exploration. Amer. Assoc. Petrol. Geol. Bull., 67: 916-926.
Waples. D.W., 1984. Thermal models for oil generation. In: J. Brooks and D. Welte (Editors), Advances
in Petroleum Geochemistry, Academic Press, 7-68.
Warde, J.M., 1950. Refractory clays in the Union of South Africa. Jour. Amer. Ceram. Soc., 29: 257-260.
Warshaw, C.M., 1957. The mineralogy of glauconite Ph.D. Dissertation, Penn. State U.. University Park,
Pa.. 147p.
Watanabe, T., 1981. Identification of illite/montmorillonite interstratifications by x-ray powder diffrac-
tion. Jour. Miner. Soc. Japan, Spec. Issue 15, 32-41 (in Japanese).
Watts, N.L., 1976. Paleopedogenic palygorskite from the basal Permo-Triassic of northwest Scotland.
Amer. Miner., 61: 299-302.
Weaver, C.E.. 1953a. A lath shaped nonexpanded dioctahedral 2:l clay mineral. Amer. Miner., 38:
270-289.
Weaver, C.E., 1953b. Mineralogy and petrology of some Ordovician K-bentonites and related limestones.
Geol. Soc. Amer. Bull., 64: 921-944.
Weaver, C.E., 1955. Mineralogy and petrology of the rocks near the Quadrant-Phosphoria boundary in
southwest Montana. Jour. Sed. Petrol., 25: 163-192.
Weaver, C.E., 1956. The distribution and identification of mixed-layer clays in sedimentary rocks. Amer.
Miner., 41: 202-221.
Weaver, C.E., 1957. Clay mineral facies in the Texas offshore Miocene and Post-Miocene, High Island
(Neptune) area. Shell Res. Div. Co. Rpt., 12 p.
Weaver, C.E.. 1958a. The effects and geologic significance of potassium “fixation” by expandable clay
minerals derived from muscovite, biotite, chlorite and volcanic material. Amer. Min., 43: 839-861.
Weaver, C.E.. 1958b. Geologic interpretation of argillaceous sediments. Bull. Amer. Assoc. Petrol. Geol.,
42: 254-271.
Weaver, C.E., 1958c. A discussion of the origin of clay minerals in sedimentary rocks. Clays Clay Miner.,
5: 159-173.
Weaver, C.E., 1959. The clay petrology of sediments. Clays Clay Min., 6: 154-187.
Weaver, C.E., 1960. Possible use of clay minerals in search for oil. Amer. Assoc. Petrol. Geol. Bull., 44:
1505-1518.
Weaver, C.E.. 1961a. Clay minerals of the Ouachita structural belt and adjacent foreland. The Ouachita
System, Flawn, P.T., Goldstein Jr., A,, King, P.B. and Weaver, C.E.. U. Texas, Pub. No. 6120,
147- 162.
Weaver, C.E., 1961b. Clay mineralogy of the late Cretaceous rocks of the Washakie Basin. Wyoming
Geol. Soc. Field Trip, 148-154.
Weaver, C.E., 1965. The potassium content of illite. Science, 147: 603-605.
Weaver, C.E.. 1967a. Variability of a river clay suite. Jour. Sed. Petrol., 37: 971-974.
Weaver, C.E.. 1967b. Potassium, illite and the ocean. Geochim. Cosmochim. Acta, 31 : 2181-2196.
Weaver, C.E., 1 9 6 7 ~ .The significance of clay minerals in sediments. In: B. Nagy and U. Colombo
(Editors), Fundamental Aspects of Petroleum Geochemistry, Elsevier, 37-76.
Weaver, C.E., 1968. Mineral facies in the Teritiary of the Continental Shelf and Blake Plateau.
Southeastern Geol.. 9: 57-63.
Weaver. C.E., 1976. The nature of T i 0 2 in kaolin. Clays Clay Miner., 24: 215-218.
Weaver, C.E., 1978. Mn-Fe coatings on saprolite fracture surfaces. Jour. Sed. Petrol., 48: 595-610.
Weaver, C.E., 1979. Geothermal alteration of clay minerals and shales: diagenesis. Office Nuclear Waste
Isolation, Tech. Rpt. 70, Battelle, Columbus, Oh., 176p.
Weaver, C.E., 1980. Fine-grained rocks: shales of physilites. Sed. Geol.. 27: 301-313.
783

Weaver. C.E., 1984a. Origin and geologic implications of the palygorskite of the S.E. United States. In:
A. Singer and E. Galen (Editors), Palygorskite-Sepiolite, Elsevier, 39-58.
Weaver, C.E. and Associates, 1984. Shale-slate Metamorphism In Southern Appalachians. Elsevier, 239p.
Weaver.C.E. and Beck, K.C., 1971. Clay water diagenesis during burial: how mud becomes gneiss. Geol.
SOC.Amer. Spec. Paper 134, 96p.
Weaver, C.E. and Beck, K.C., 1977. Miocene of the S.E. United States: A Model for Chemical
Sedimentation in a Peri-Marine Environment. Elsevier, 234p.
Weaver, C.E. and Brockstra, B.R., 1984. Illite-mica. In Weaver, C.E. and Associates. Shale-Slate
Metamorphism in Southern Appalachians, Elsevier, 67-98.
Weaver, C.E., Eslinger, E.V. and Yeh, H.W., 1984. Oxygen Isotopes. In Shale-Slate Metamorphism in
Southern Appalachian, Weaver, C.E. and Associates, Elsevier, 142-152.
Weaver, C.E., Highsmith, P.B and Wampler, J.M., 1984. Chlorite. In Shale-Slate Metamorphism in
Southern Appalachians, Weaver, C.E. and Assoiciates, Elsevier, 99-140.
Weaver, C.E. and Padan, A., 1983. Detailed mineralogy, petrology and water content of repository salts.
Battelle Proj. Manag. Rpt. No. 8, 57p.
Weaver, C.E. and Pollard, L.D., 1973. The Chemistry of Clay Minerals. Elsevier, 213 p.
Weaver, C.E. and Stevenson Jr., R.G., 1971. Clay minerals in the Cretaceous rocks of Florida. Geol. Soc.
Amer. Bull., 82: 3457-3460.
Weaver, C.E. and Wampler, J.M., 1970. K, Ar, illite burial. Geol. SOC.Amer. Bull., 81: 3423-3430.
Weaver, C.E. and Windom, H.L., 1988. Origin of marsh clays in the Southeastern United States. In press.
Weber, K., 1976. Gefugeuntersuchungen an transversalgeschieferten Gesteinen aus dem ostlichen
Rheinischen Schiefergebirge (Ein Bietrag zur Genese der transversalen Schieferung): Geologisches
Jahrbuch, ser. D. 15: 3-98.
Weed, S.B., Davey, C.B. and Cook, M.G., 1969. Weathering of mica by fungi. Soil Sci. SOC.Amer. Proc.,
33: 702-706.
Weiss, A.A., Koch, G., and Hoffman, U., 1955. Saponite. Ber. Dtsch. Keram. Ges., 32: 12-17.
Weir, A.H. and Catt, J.A., 1965. The mineralogy of some Upper Chalk samples from the Arundel area,
Sussex. Clay Minerals, 6: 97-109.
Weir, A.H. and Greene-Kelly, R., 1962. Beidellite. Amer. Miner., 47: 137-146.
Wells, J.T., 1987. Entrapment of shelf-destined mud, Cape Lookout Bight, N.C.. Southeastern Sec. Geol.
SOC.Amer. Abs. with Prog.. 135.
Wentworth, C.K., 1922. A scale of grade and class terms for clastic sediments. Jour. Geol., 30: 377-392.
White, E.M. and Riecken, F.F., 1955. Brunizem-gray brown podsolic soil biosequences. Soil Sci. SOC.
Amer. Proc., 19: 504-509.
White, S.H. and Knipe, R.J.. 1978. Microstructure and cleavage development in selected slates. Contrib.
to Miner. Petrol., 66: 165-174.
Whitehouse. U.G., Jeffrey, L.M. and Debbrecht, J.D., 1960. Differential settling tendencies of clay
minerals in saline waters. Clays Clay Miner. 7: 1-79.
Whittig, L.D. and Jackson, M.L., 1955. Mineral content and distri-bution as indexes of weathering in the
Omega and Ahmeek soils of northern Wisconsin. Clays Clay Miner., 4: 362-371.
Whittington, H.B. and Hughes, C.P., 1972. Ordovician geography and faunal provinces deduced from
trilobite distribution. Philo. Trans. Royal COC.London, 263: 235-278.
Wiersma, J., 1970. Provenance, genesis and paleo-geographical implication of macrominerals occurring in
sedimentary rocks of Jordan Valley Area. Publ. Fys. - Geogr. Bodenk. Lab Univ. Amsterdam, 15,
240p.
Wiewibra, A,, 1972. Mixed-layer kaolinite-smectite from Lower Silesia, Poland: Final report. Proc. Inter.
Clay Conf. Madrid, 75-86.
Wiewibra, A,, Dubinska, E. and Iwasihska, I., 1981. Mixed-layering in Ni-containing talc-like minerals
from Szklary, Lower Silesia, Poland. Proc. Int. Clay Conf., Bologna, 111-124.
Wilding. L.P., Smeck, N.E. and Hall, G.F., 1983. Pedogenesis and Soil Taxonomy, Developments in Soil
Science 11A. 303p and 11B, 410p.
Willliams, D.L., Von Herzen, R.P., Sclater, J.G. and Anderson, R.N., 1974. The Galapagos Spreading
Center: lithospheric cooling and hydrothermal circulation. Geophys. Jour. Roy.Astron. SOC..38:
587-608.
784

Williams, E.G., Bergenback, R. E.. Falla. W.S. and Udagawa. S.. 1968. Origin of the Pennsylvania
underclays in western Pennsylvania. Jour. Sed. Petrol., 38: 1179-1 193.
Williams. P.F., 1972. Development of metamorphic layering in the low grade metamorphic rocks at
Bermagui, Australia. Amer. Jour. Sci.: 272, 1-47.
Williams. P.F., 1977. Folation: A review and discussion. In: G.S. Lister, P.F. Williams, H.J. Zwart and
R.J. Lisle (Editors), Fabrics, Microtextures and Microtectonics. Tectonophysics, 39: 305-308.
Williamson, I.A.. 1961. Tonsteins: a possible additional aid to coalfield correlation. Mining Mag.
(London), 104: 9-14.
Wilson. J.T.. 1966. Did the Atlantic close and then reopen? Nature, 21: 676-681.
Wilson, M.D. and Pittman, E.D.. 1977. Authigenic clays in sandstones: Recognition and influence o n
reservoir properties and paleoenvironmental analysis. Jour. Sed. Petrol., 47: 3-31.
Wilson. M.D. and Tillman. R.W., 1974. Diagenetic destruction of feldspar and genesis of clay: their
influence on sandstone classification and grain size analysis. Geol. Soc. Amer. Abs. with Programs. 5:
130- 131.
Wilson, M.J.. 1965. The origin and geologic significance of the South Wales underclays. Jour. Sed.
Petrol.. 35: 91-99.
Wilson, M.J., 1970. A study of weathering in a soil derived from a biotite-hornblende rock. I Weathering
of biotite. Clay Min., 8: 291-303.
Wilson, M.J., 1971. Clay mineralogy of the Old Red Sandstone (Devonian) of Scotland. Jour. Sed.
Petrol., 41: 995-1007.
Wilson, M.J. and Bain, D.C.. 1970. The clay mineralogy of the Scottish Dalradian meta-limestones,
Contrib. Miner. Petrol., 26: 285-295.
Wilson. M.J., Bain, D.C. and McHardy, W.J., 1971. Clay mineral formation in a deeply weathered
boulder conglomerate in northeast Scotland. Clays Clay Miner., 19: 345-352.
Wilson, M.J., Bain, D.C.. McHardy.W.J. and Berrow, M.L.. 1972. Clay-mineral studies o n some
Carboniferous sediments in Scotland. Sed. Geol., 8: 137-150.
Wilson, M.J. and Farmer, V.C., 1970. A study of weathering in a soil derived from a biotite-hornhlende
rock 11. The weathering of hornblende. Clay Minerals, 8: 435-444.
Wilson. M.J. and Nadeau. P.H.. 1984. Interstratified clay minerals and weathering processes. I n : J.I.
Drever (Editor), The Chemistry of Weathering, Reidel, D., Pub. Co.. Dordretch, 97-118.
Windiey. B. F., 1977. The Evolving Continents. John Wiley and Son, N.Y., 385p.
Windom, H.L., 1969. Atmospheric dust records in permanent snow-fields: implications t o marine
sedimentation. Geol. Soc. Amer. Bull.. 80: 761 -782.
Windom. H.L.. 1975. Eolian contribution to marine sediments. Jour. Sed. Petrol., 45: 520-529.
Windom, H.L., 1976. Lithogenous material in marine sediments. In: J.P. Riley and R. Chester (Editors).
Chemical Oceanography. Academic Press, 103-1 36.
Windom. H.L. and Chamberlain, C.F., 1978. Dust-storm transport of sediments to the North Atlantic
Ocean. Jour. Sed. Petrol.. 48: 385-388.
Windom. H., Griffin, J. and Goldberg. E.D., 1967. Talc in atmospheric dust. Envir. Sci. Tech.. 1:
923-926.
Windom, H.L., Neal, W.J. and Beck, K.C.. 1971. Mineralogy of sediments in three Georgia estuaries.
Jour. Sed. Petrol., 41: 497-504.
Winkler. H.G.F.. 1979. Petrogenesis of metamorphic rocks. Springer-Verlag. N.Y.. 348p.
Winston, D.. 1973. The Precambrian Missoula Group of Montana as a braided stream and sea margin
sequence., In: Belt Symposium. Vol.1. Dept. Geol., U. Idaho and Idaho Bureau Mines and Geol..
Moscow, Idaho, 208-220.
Wise, W.S. and Eugster. H.P., 1964. Celadonite: synthesis, thermal stability and occurrences. Amer.
Miner., 49: 1031-1083.
Wolf. M., 1975. Uber die Beziehungen zwischen illite-kristallinitat und inkohlung. N.Jb. Geol. Paliiont
Mh., 437-447.
Wollast. R., 1967. Kinetics of the alteration of K-feldspar in buffered solutions at low temperature.
Geochim. Cosmochim. Acta, 31: 635-648.
Wollast, R., Mackenzie, F.T. and Bricker, O.P., 1968. Experimental precipittion and genesis o f sepiolite
at earth-surface condi lions. Amer. Miner,, 53: 1645-1662.
785

Wright, L.D., 1978. River deltas. In: R.A. Davis, Jr. (Editor), Coastal Sedimentary Environments
Springer-Verlag, N.Y., 5-68.
Wyrwicki, R. and Wiewibra, A., 1981. Clay minerals of the Upper Neogene sediments in Poland. Bull.
Acad. Polonaise Sci., 29: 67-71.
Yaalon, D.H., 1962a. Mineral composition of the average shale. Clay Min. Bull., 5: 31-36p.
Yaalon, D.H., 1962b. Weathering and soil development through geologic time. Bull. Res. Coun. Israel,
Sec. G., V. 11G, Proc. Israel Geol. SOC.,4th Congr. Israel Assoc. Adv. Sci., 1961, 147-148.
Yaalon, D.H., Nathan, Y., Koyumdjisky, H., and Dan, J., 1966. Weathering and catenary differentiation
of clay minerals on various parent materials in Israel. Proc. Inter. Clay Conf., Jerusalem, 187-198.
Yariv, S. and Shoval, S., 1975. The nature of the interaction between water molecules and kaolin-like
layers in hydrated halloysite. Clays Clay Miner., 23: 473-474.
Yau, L.Y.C. and Peacor, D.R., 1984. TEM/AEM study of chlorite diagenesis in well I.I.D. No. 2 in
Salton Sea Geothermal Field, California. Clay Miner. SOC.Program with Abs., Baton Rouge, 120.
Yoder, H.S. and Eugster, H.P., 1955. Synthetic and natural muscovite. Geochim. Cocmochim. Acta, 8:
225-280.
Zabawa, C.F., 1978. Microstructure of agglomerated suspended sediments in Northern Chesapeake Bay
estuary. Science, 202: 49-51.
Zaytseva, E.D., 1966. Capacity of exchange and exchange cations of sediments of the Pacific Ocean. In:
S.V. Brujewicz (Editor), Khimiya Tikhogs Okeana (Chemistry of the Pacific Ocean), Izd. “Nauka”,
110-112.
Zemmels, I., Cook, H.E. and Hathaway, J.C., 1972. X-ray mineralogy studies-Leg XI. In: C.D. Hollister,
J.E. Ewing, et al., 1972. Initial Reports of the Deep Sea Drilling Project, XI. Wash. (U.S. Gov.
Printing Office) 729-790.
Zemmels, I., Harrold, P.J. and Cook, H.E., 1976. X-ray mineralogy data from the Argentine Basin-Leg 38
Deep Sea Drilling Project. In Barker, P.F., Dalziel, I.W.D. et al., 1976. Initial Reports of the Deep
Sea Drilling Project, 36, Wash. (U.S. Gov. Printing Office), 1017-1032.
Zen, E-An, 1959. Mineralogy and petrology of marine bottom sediment samples off the coast of Peru and
Chile. Jour. Sed. Petrol., 29: 513-539.
Ziegler, P.A., 1982. Geological Atlas of Western and Central Europe. Elsevier, 130p.
Zitzmann, A,, 1977. The iron ore deposits of Europe and adjacent areas, Vol. 1. Fed. Inst. Geosci. Nat.
Resour., Hannover, 418 p.
Zitzmann, A., 1978. The iron ore deposits of Europe and adjacent areas, Vol. 2. Fed. Inst. Geosci. Nat.
Resour., Hannover, 386 p.
Zharkov, M.A., 1981. History of Paleozoic Salt Accumulations. Springer-Verlag, 308p.
Zkhus, I.D., 1961. Clay mineral associations of the Lower Carboniferous in the southwestern part of the
Moscow Basin. U S . Office Tech Services, Tech. Translations, 11402, 1-3, 430-431.
Zvyagin, B.B., 1957. Determination of the structure of celadonite by electron diffraction. Kristallografiya,
2: 388-394.
Zvyagin, B.B., 1960. Electron-diffraction determination of the structure of kaolinite. Kristallografiya, 5:
40-50.
This Page Intentionally Left Blank
787

AUTHOR INDEX1

Aagaard, P., 129, 723 Andrews, J.E., 375, 724


Aba-Husayn, M.M., 738 Andrews, P.B., 8 , 724
Abbot, P.L., 147, 723 Angel, B.R., 8 , 24, 724
Abder-Ruhman, M., 773 Angino, E.A., 729
Adams, J.K., 678, 751 Anosov, F.Y., 57, 778
Adams, W.A., 582, 739 Antia, D.D.J., 519, 724
Addy, S.T., 776 Antiweiler, R.C., 134, 724
Ager, D.V., 588, 635, 680, 681, 723 Aoki, S., 328, 337, 368, 724
Aguayo, E., 738 Aomine, S., 25, 143, 724
Aguayo, L., 687, 764 April, R.H., 79, 80, 634, 635, 648, 724, 781
Ahn, J.H., 419-421, 427, 432, 436-440, Armitage, T.M., 724, 743
443, 447, 533, 723 Arnaud Sr., R.J., 179, 724
Ahn, J.O., 756 Arnone, R., 215-218, 724
Aida, S., 781 Aronson, J.L., 158, 427, 428, 496, 499,
Al-Bakir, K.F., 342, 723 500, 724, 756
Al-Ghadban, A., 342, 723 Arrese, F., 687, 724
Alberti, A., 51, 723 Arrhenius, G., 302, 386, 724, 728
Albrecht, P., 522, 723 Arthur, M.A., 694, 724
Aleksandrova, A.V., 774 Artru, Ph., 419, 725
Alexander, E., 746 Asprey, K.W., 777
Alexander, L.T., 30, 723 Ataman, G., 277, 725
Alexandrova, V.A., 754 Atsuyuki, I., 752
Alexiades, C.A., 164, 723 Atterberg, A., 5 , 725
Alietti, A., 54, 94, 97, 723 Attiwel, P.M., 751
Allen, B.L., 183, 723, 727, 758 Aubry, M.P., 400, 727
Aubry, M.-P., 738
Allen, R.C., 350, 386, 758
Axelrod, D.I., 576, 625, 654, 679, 725
Alling, H.L., 6, 723
Axelrod, J.M., 740
Almon, W.R., 79, 541, 542, 546, 553, 660,
723, 735
Baadsgaard, E.G., 743
Alt, J.C., 363, 364, 724
Baadsgaard, H.B., 731
Altaman, G., 408, 639, 642, 757
Baadsgaard, H.S.S., 398, 725
Altaner, S.P., 419, 426, 432, 724, 727, 748
Badaut, D., 275, 725
Altmann, R.S., 756 Bader, R.G., 193, 747
Altschuler, Z.S., 147, 724 Bailey, S.W., 1, 2, 13, 15, 16, 19, 26, 31,
Amador, E.S., 692, 693, 724 45, 67-71, 76, 82, 87, 88, 93, 96-98, 362,
Amaud, R.J. St., 760 486, 569, 725, 728, 730, 738, 743, 751,
Ames, L.L., 5 5 , 724 771
Amiri-Garroussi, K . , 650, 724 Bain, D.C., 59, 79, 160, 167, 725, 762, 784
Amouric, M., 87, 91, 724 Baird, S.L., 762
Anand, R.R., 131, 133, 724 Baker, E.T., 264, 290, 519, 725, 776
Anderson, R.F., 736 Bakun, N.N., 534, 545, 770
Anderson, R.N., 783 Bale, A.J., 761
Ball, D.F., 584, 725
'Page references to text are in Roman Ballanca, A., 725
type, to bibliographical entries in italics. Baltrusaitis, E.J., 589, 725
788

Balukhovsky, A.N., 771 Berggren, W.A., 400, 664, 694, 727


Bambach, R.K., 572, 725 Berner, R.A., 123, 395, 721, 727
Banes, P.M., 736 Berrow, M.L., 784
Banin, A., 775 Berry, R.W., 317, 320, 321, 727
Bannister, F.A., 32, 44, 725 Bertrand, R., 747
Barbieri, M., 688, 725 Beskin, E.A., 543, 662, 727
Barde, N.K., 155, 771 Bethke, C.M., 419, 727
Barker, P.F., 696, 699, 725 Bevan, J.C., 730
Barnhisel, R.I., 77, 170-172, 726 Bhattacharya, N., 409, 737, 768
Baroffio, J.R., 782 Bhattacharyya, D.P., 31, 32, 388, 727
Barrett, P.J., 8, 726 Biggs, R.B., 224, 225, 727
Barretto, H.T., 760 Bigham, G.N., 325, 326, 727
Barshad, I., 84, 85, 113-117, 151, 161, Bigham, J.M., 182, 727
162, 169, 179, 726, 778 Billings, G.K., 747
Bartholomk, P., 652, 726 Bin, W.C., 131, 145, 739
Barzanji, A.F., 183, 726 Birch, G.F., 322, 391, 395, 727
Barzanji, J.F., 183, 739
Birkeland, P.W., 104, 107, 112, 114, 115,
Basan, P.B., 244, 741 124, 151, 727
Bassett, R.L., 409, 410, 726
Biscaye, P.E., 202, 291, 292, 310, 312, 314,
Bassett, W.A., 37, 162, 726
317, 318, 323, 727, 737, 754
Bates, T.F., 28, 77, 172, 245, 246, 726, 749
Bischoff, J.L., 338, 365, 368, 369, 727, 772,
Baum, G.E., 400, 745
773
Bausch, W.M., 651, 726
Bish, D.L., 172, 727, 730
Baxter, J.W., 741
Bismuth, H., 669, 727
Bayliss, P., 594, 657, 661, 679, 726, 732,
Bjorlykke, K., 752
757
Bjorlykke, K., 537, 540, 577-579, 587, 727
Baysal, O., 277, 725
Beardsley, G.F., 731 Black, P.M., 487, 489, 728
Beattie, J.A., 673, 734 Blackmon, P.D., 773
Beaufort, D., 768 Blake, D.F., 433, 728
Beaver, P.J., 120, 726 Blanche, J.B., 554, 728
Beck, C.W., 446, 745 Blancher Jr., D.W., 599, 728
Beck, K.C., 102, 151, 152, 174, 182-184, Bland, D.J., 751
252, 276, 342, 380, 383-386, 400, 401, Blatt, H., 8, 539, 540, 715, 717, 718, 728,
405, 408, 419, 421, 437, 439, 440, 444, 751
510-512, 537-539, 565, 677, 691, 699, Blatter, C.L., 79, 446, 656, 728
701, 707, 783, 784 Bledsoe, A.O., 752, 753
Behrens, E.W., 776 Blokh, A.M., 52, 728
Bell, J.M., 742 Blount, A.M., 15, 17, 728
Bell, T.E., 419, 435, 436, 726 Bocquier, G., 146, 762, 778
Belt Symposium, 570, 726 Bode, G.W., 8, 728
Bennett, R.H., 710, 726, 752 Bodenheimer, W., 726
Benson, L.V., 550, 551, 675, 726, 776 Bodine Jr., M.W., 75, 94, 408-410, 587,
Bentor, Y.K., 89, 399, 577, 652, 668, 690, 617, 627, 628, 728
726 Boerngen, J.G., 773
BCrczi, I., 641, 727 Boettinger, J.L., 157, 775
Berg-Madsen, V., 391, 727 Bogolyubova, L.I., 137, 779
Bergenback, R.E., 783 Bohor, B.F., 674, 728, 772
Berger, W.H., 384, 385, 727 Boland, J.N., 131, 132, 738
789

Boles, J.R., 419, 420, 423, 425, 428, 429, Brown, B.E., 68, 159, 725, 730
437, 440, 447, 464, 539, 545, 551, 552, Brown, E.H., 487, 730
676, 728, 768 Brown, G., 3, 69, 84, 93, 169, 730, 757,
Boltenhagen, C., 727 769
Bonatti, E., 102, 338, 339, 372, 375, 380, Brown, J.L., 756
386, 728, 754, 767 Brown, K.M., 730
Bonis, S.B., 735 Brown Jr., L.F., 610, 730
Bonnot-Courtois, C., 733 Brown, P.E., 751
Booth, J.S., 581, 728 Brown, R.E., 69, 486, 730
Borbunosa, Z.N., 768 Bruce, C.H., 422-424, 429, 730
Borchardt, G.A., 49, 154, 166, 728 Brun, J., 584, 730
Borchert, H., 88, 728 Brunton, G., 607, 627, 775
Bormann, F.H., 751 Brusewitz, A.M., 426, 585, 779
Bornhold, B.D., 391, 393, 394, 728 Bryant, W.R., 726, 729
Borovec, Z., 754 Brydon, J.E., 157, 159, 167, 169, 171, 730,
Borst, R.L., 589, 728 740, 754, 758
Bossi, G., 630, 764 Buckley, D.E., 768
Bossi, G.E., 174, 630, 645, 729 Buckley, H.A., 87, 88, 91, 94, 362, 730
Bostick, N.H., 517, 518, 729, 757 Buie, B.F., 678, 765
Bottino, M.L., 399, 589, 742, 745 Bundy, W.B., 446, 730
Boudda, A., 732 Burger, K., 624, 730
Boujo, A., 691, 729 Burley, S.D., 636, 730
Bourguignon, P., 621, 779 Burrell, D.C., 244, 764
Bouroz, A., 624, 729 Burst, J.F., 87, 92, 390, 391, 418, 677, 731
Bowles, F.A., 102, 352, 375, 380, 710, 729 Burtner, R.L., 524, 731
Bowser, C.J., 270, 751 Buseck, P.R., 130, 178, 738
Boyce, R.E., 8, 729 Bustillo, A.M., 650, 731
Boyer, P.S., 388, 729 Butler, J.C., 746
Boyle, J.R., 38, 729 Butuzova, G.Yu., 366, 731
Bradley, W.F., 80, 81, 95, 96, 101, 405, Butze, H., 308, 731
409, 602, 729, 743, 762 Buxton, B.P., 125, 731
Bradley, W.H., 674, 729 Bystrom, A.M., 63, 426, 445, 584, 731
Bradshaw, M.J., 649, 729 Bystrom-Asklund, A.M., 584, 731
Braid, S.P., 692, 729
Braitsch, O., 94, 408-410, 629, 729 Caballero, M.A., 642, 731
Braun, H., 88, 728 Cadigan, R.A., 646, 731
Brauner, K., 95, 96, 729 Cahet, G., 393, 731
Bray, R.H., 743 Cahoon, H.P., 56, 731
Breitschmid, A., 741 Caillkre, S., 75, 77, 100, 731
Brell, J.M., 758 Callen, R.A., 102, 182, 342, 380, 382, 383,
Brereton, N.R., 740 400, 690, 731
Bricker, O.P., 130, 729, 756, 784 Calvert, C., 21, 731
Brigatti, M.F., 51, 54, 79, 84, 723, 729 Calvert, S.E., 769
Brindley, G.W., 3, 13-18, 26, 28, 31, 34, Cameron, E.M., 512, 731
69, 71, 77, 93, 97, 99, 145, 171-173, 725, Camez, T., 84, 760
730, 737, 767 Campbell, F.A., 512, 661, 731
Brockstra, B.R., 39, 783 Cano-Ruiz, J., 98, 758
Brookins, D.G., 398, 399, 730, 732 Carden, K.L., 293, 731
Brooks, R.A., 269, 730 Carlo, E.A., 741
790

Carlson, T.N., 767 CiEel, B., 48, 733


Carmichael, I.S.E., 729 Cimbinik6vb, A., 88, 733
Carmouze, J.P., 765 Claridge, G.G.C., 778
Caron, J.-M., 756 Clark, F.W., 8, 487, 512, 733
Carothers, W.D., 753 Clark, P.E., 742
Carozzi, A.V., 721, 731 Clarke Jr., O.M., 149, 663, 733
Carrigy, M.A., 657, 661, 731 Clattenburg, D.A., 777
Carroll, D., 334, 346, 732 Clauer, N., 398, 695, 733, 756
Carson, B., 776 Clayton, R.N., 328, 733
Carstea, D.D., 77, 171, 172, 732, 745 Clemency, C.V., 145, 733
Carver, R.E., 678, 732 Cluff, R.M., 713, 733, 745
Case, E.C., 625, 732 Coakley, R.D., 267, 733
Cashion, J.D., 742 Cobban, W.A., 743
Cashman, S.M., 758 Cobler, R., 737
Castaiio, J.R., 517, 732 Cocker, J.D., 533, 733
Castillo, A., 687, 742 Cohen, A.D., 269, 776
Cauffman Jr., L.B., 594, 732 Cole, M.J., 31, 733
Caughey, C.A., 287, 777 Cole, R.D., 714, 716, 734
Cavaroc, V.V., 597, 732 Cole, T.G., 338, 365, 366, 734
Cavelier, C., 779 Coleman, D.D., 742
Cecconi, S., 141, 732 Coleman, J.M., 8, 248, 249, 734, 747
Cecil, C.B., 618, 732, 769 Collins, K., 710, 734
Cervantes, A., 752 Colomb, E., 682, 732
Chaabani, F., 757 Colquhoun, D.J., 766
Chagnon, A., 584, 730, 747 Commeau, J.A., 754
Chamberlain, C.F., 303, 784 Condie, K.C., 572, 734
Chamley, H., 375, 380, 571, 665, 668, 682, Conley, F.R., 736
688, 696-699, 701-703, 732, 737, 774 Conomos, T.J., 754
Chandra, D., 776 Conrad, D., 771
Chang, H.K., 550, 732 Cook, H.E., 758, 785
Charollais, J., 755 Cook, M.G., 162, 163, 168, 734, 770, 783
Chaudhuri, S., 399, 627, 732, 756 Cooke, G.E., 9, 734
Chen, B.Y., 446, 732 Coombs, D.S., 418, 734
Chen, H.S., 742 Cooper, W.A., 749
Chen, P.Y., 328, 330, 732 Copeland, R.A., 353, 734
Chennaux, G., 597, 696, 700, 701, 733 Copin, E., 750
Chernov, A.A., 778 Corliss, J.B., 369, 734, 737
Chesnut, D.R., 615, 733 Cormier, R.F., 747, 749
Chester, R., 313, 341-343, 733, 770 Cornet, C., 667, 685, 734
Chester, R.H., 305, 307, 317, 319, 733 Correns, C.W., 5 , 734
Chesters, G., 776 Cott, J.A., 666, 783
Cheverry, C., 778 Court, J.E., 267, 734
Chichester, F.W., 157, 733 Courtois, C., 748, 751
Chilingarian, G.V., 705, 707, 708, 733, 769 Cousins, P., 760
Choukroune, P., 748 Couto Anjos, S.M., 426, 669, 734
Chowdhury, A.N., 650, 733 Couture, R.A., 102, 380, 734
Christ, C.L., 99, 733 Cowperthwaite, I.A., 666, 734
Chudaev, B., 755 Craig, J.R., 758
Churchman, G.J., 167, 169, 733, 778 Crane, K., 734
791

Crocket, D.S., 710-712, 739 Deynoux, M., 732


Cronan, D.S., 746 Dickinson, W.R., 743
Crossey, L.J., 537, 538, 777 Dillon, W., 760
Cubitt, J.M., 605, 627, 734 Dimichele, W.A., 766
Cummings, J.C., 673, 734 Divis, A.F., 177, 736
Curray, J., 738 Dixon, J.B., 27, 143, 690, 736, 738, 773
Curry, D., 764 Dmitriev, Y., 746
Curry, J.R., 283, 734 Dmitrik, A.L., 731, 754
Curtin, D., 160, 734 Dobryansky, A.F., 518, 736
Curtis, C.D., 146, 446, 537, 538, 543, 734, Dodd, C.G., 43, 545, 628, 736, 758
750, 757 Dodge, C.F., 746
Curtis, G.H., 739 Dodson, M.H., 390, 397, 398, 725, 764
Doebl, F., 424, 447, 736
Dahl, W.M., 541, 543, 662, 734 Dolcater, D.L., 23, 736
Dallmeyer, R.D., 492, 734 Donaldson, A.C., 618, 736
Dalrymple, G.B., 492, 493, 735 Donnelly, T.W., 361, 366, 375, 654, 679,
Dalziel, I.W.D., 725 736
Damuth, J.E., 288, 735 Donze, P., 727
Dan, J., 785 Douglas, L.A., 50, 161, 166, 174, 736, 758,
Danberger, H.H., 518, 729 764
Dandois, par Ph., 596, 621, 735 Doyle, L.J., 254, 255, 736
Dangit, A., 146, 735 Drever, J.I., 129, 134, 242, 337, 349, 382,
D’Anglejan, B.F., 204, 215, 322, 735 724, 736
d’Argoud, G., 732 Drits, V.A., 44, 408, 731, 754, 774
d’Argoud, G.G., 375, 732, 774 Droste, J.B., 84, 158, 159, 271, 276, 407-
Davey, C.B., 783 409, 587, 589, 590, 594, 600, 601, 673,
Davies, D.K., 541, 542, 551, 723, 735, 745 737
Davis, A.G., 638, 735 Droubi, A., 742
Davis, J.C., 656, 735 Drymond, J., 734
Davis Jr., R.A., 244, 248-250, 656, 735 Dubinska, E., 783
de Oliveira, J.J., 740 Dubreuilh, J., 682, 737
De Swardt, A.M.I., 597, 630, 771 Duff, P.M.D., 624, 767
Dean, R.S., 43, 440, 533, 582, 735, 754 Dulong, F., 769
Dean, W.E., 751 Dulong, F.T., 732
Debbrecht, J.D., 783 Dumbleton, M.J., 120, 638, 726, 737
Deer, W.A., 73, 93, 735 DuMontelle, P.B., 745
Deffeyes, K.S., 352, 735 Duncan, J., 570, 762
Degens, E.T., 610, 735 Duncan, J.R., 264, 737
Deike, R.G., 273, 735 Dunoyer de Segonzac, G., 419, 424, 425,
DeKimpe, C., 146, 735 444, 447, 450, 452, 458, 523, 541, 571,
DeLange, G.J., 350, 735 577, 588, 596, 597, 621, 623, 629, 643,
Delany, A.C., 304, 305, 735 651, 668, 669, 686, 732, 733, 737
Demaison, G.J., 445, 447, 516, 520, 735 Duplay, J., 765
Dennison, J.W., 589, 735 Dutta, P.K., 558, 737
Denny, M.V., 749 Dutton, S.P., 543, 755
Depetris, P.J., 323, 736 Dwornik, E.J., 724
Derouane, E.G., 747 Dyer, K.R., 213, 220, 737
Desprairies, A., 101, 102, 736 Dymond, J., 179, 336, 338, 368, 369, 737,
Deuser, W.G., 293, 736 746
792

Dypvik, H., 651, 737 Eswbran, H., 726


Eswaran, J., 183, 739
Eargle, D.H., 677, 737 Etheridge, M.A., 460, 739
Earley, J.W., 252, 628, 657, 671, 737, 761, Eugster, H.P., 39, 41, 270, 739, 784, 785
775 Evans, L.J., 582, 739
Eaton, G.P., 309, 737 Evernden, J.F., 496, 739
Eberl, D., 430, 431, 737, 738 Ewing, J.I., 748
Eberl, D.D., 33, 39, 42, 43, 49, 59, 61, Ewing, M., 260, 261, 322, 324, 750, 761
62, 94, 97, 180, 276, 420, 422, 425, 431, Eymery, J.-P., 768
534-536, 613, 614, 738, 753, 769, 776
Echle, W., 638, 738 Faas, R.W., 710-712, 739
Eckel, E.C., 512, 738 Fabricius, F., 774
Eckhardt, F.J., 447, 738 Fahey, J.J., 31, 98, 675, 740
Edger, N.T., 766 Fairbairn, H.W., 748, 749
Edzwald, J.K., 209-212, 214, 233, 235, Falla, W.S., 783
738 Fallis, J.M., 79, 627, 754
Egashira, K., 159, 738 Fan, P.F., 334, 386, 777
Eggleton, R.A., 76, 130-132, 178, 738, 743 Fanning, D.S., 175-177, 740, 768, 773
Ehlers, E.G., 585, 738 Farmer, V.C., 160, 730, 762, 784
Ehlmann, A.J., 195, 391, 738 Faust, G.T., 31, 55, 723, 740
Ehrlich, H.L., 137, 774 Favretto, L., 757
Einsele, G., 371, 384, 738 Faye, B., 729
Einstein, H.A., 246, 738 Ferrell Jr., R.E., 269, 730, 747
Eisma, D., 322, 738 Ferrero, A., 687, 742
Eittreim, S.L., 291, 292, 727 Feuillet, J.P., 233, 234, 740
Eklund, W., 368, 369, 737 Fisher, M.J., 629, 636-638, 642, 740
Eldenfield, H., 733 Fisher, W.R., 179, 762
Elderfield, H., 372, 381, 405, 738 Fiskell, J.G.A., 147, 740
Elders, W.A., 75, 446, 487, 488, 491, 758 Fitch, F.I., 666, 734
Elizalde, Par G., 597, 623, 630, 738 Fitch, F.J., 398, 740
Ellis, B.D., 313, 738 Flach, K.W., 762
Elprince, M., 183, 738 Flack, K.W., 762
Emelfanov, E.M., 768 Flawn, P.T., 7, 603, 740
Emery, K.O., 279, 280, 739 Fleischer, P., 233, 234, 265, 380, 740
Enos, P., 385, 699, 739, 750 Flemal, R.C., 671, 740
Enu, E.I., 669, 739 Flexor, J.M., 156, 740
Epshteyn, O.G., 147, 739 Flower, M.f.J., 769
Epstein, A.G., 517, 591, 739 Fodor, R.V., 746
Epstein, J.B., 739 Folger, D., 760
Eremeev, V.V., 779 Folinsbee, R.E., 731
Ernst, W.G., 45, 739 Folk, R.L., 7, 740
Eroshchev-Shak, V.A., 166, 739 Follett, E.A.C., 23, 740
Eslinger, E.V., 427, 445, 458, 489, 570, Fontes, J.C., 681, 740
739, 748, 783 Force, L.M., 664, 761
Esquavin, J., 450, 739 Forester, R.W., 537, 541, 741
Esquevin, J., 733 Forman, S.A., 157, 159, 740
Esteoule, J., 686, 739, 757 Formoso, M.L.L., 597, 623, 630, 644, 768
Esteoule-Choux, J., 682, 739, 757 Fornari, D., 738
Eswaran, H., 131, 145, 146, 155, 739 Forstner, U., 270, 762
793

Foscolos, A.E., 63,419,424,440,458,524, Galloway, W.E., 547, 742


542, 657, 740, 741, 767 Galbn, E., 758
Foss, J.E., 768 Gamble, J.R., 782
Foster, M.D., 73, 84, 486-488, 741 Garcia, J.G., 758
Foth, H.D., 105, 107, 108, 122, 741 Garcia Palacios, M.C., 687, 742
Fournier, R.O., 52, 741 Gardner, J.V., 751
Frakes, L.A., 146, 147,566,581, 654, 661, Gardner, L.R., 178, 742
664, 671,679, 741 Garrels, R.M., 122, 123, 130, 147, 300,
Francheteau, J., 748 309, 350, 512, 513, 729, 731, 742, 758
Francis, E.H., 624, 741 G a t , R.G., 437, 742
Frank-Kamonetzky, V.A., 448, 741 Gates, G.R., 673, 742
Frankart, R., 747 Gaudette, H.E., 43, 742
Frankel, L., 138, 741 Gaultier, J.P., 180, 742
Franks, P.C., 656, 741 Gauthier, J., 419, 725, 740
Franks, S.G., 419, 420,423,425,437,440, Gautier, D.L., 654, 770
447, 464, 537, 539, 541, 545, 551, 552, Ghabru, S.K., 760
676, 728, 741 Ghent, E.D., 540, 742
Fraser, G.S., 600, 741 Gibbs, R.J., 115, 189, 192, 193, 195-197,
Freas, D.H., 675, 741 212, 213, 220, 259, 314, 320, 322, 323,
Freed, R.I., 424, 677, 741 743
Freeman, T.,385, 699, 739 Giese Jr., R.F., 163, 172, 727, 743
Frey, F.A., 734 Gieskes, J., 738
Frey, M., 398, 417, 452, 490, 507, 509, Gieskes, J.M., 377, 379, 743
520-522, 563, 641, 651, 741 Gile, L.H., 106, 743
Frey, R.W., 244, 741 Giletti, B.J., 503, 743
Frezon, S.E., 603, 741 Gilkes, R.J., 152,680, 724, 743
Friend, P.F., 590, 741 Gill, J.R., 661, 743, 773
Fritz, B., 742, 778 Gill, W.D., 521,620, 743
Fritz, P., 740 Gillery, F.H., 31, 71, 730, 743
Fritz, S.J., 179, 741 Gillette, D.A., 305, 765
Frost, J.K., 745 Gilmore, J.S., 767
Fry, J.C., 166, 742 Giot, D., 729
Frye, J.D., 675, 742 Gipson Jr., M., 6, 743
Fiichtbauer, H., 409, 444, 526, 540, 629, Giresse, P.,391,393, 394, 728, 731
742 Gjems, O., 84, 166, 743
Fukama, A., 754 Glaccum, R.A., 304, 743
Fukushima, I<., 754 Glass, H.D., 610, 662, 664, 742, 743, 767,
Fullager, P.D., 589, 742 768
Fullerton, L.B., 723 Glass, H.W., 520, 768
Furbish, W.J., 772 Glover, E.D., 738
Fyfe, W.S., 130, 353, 764, 775, 778 Gluskater, H.J., 615, 768
Fysch, S.A., 23, 742 Gluskoter, H.J., 615, 743
Goh, T.B., 172, 743
Gac, J.Y., 273, 742 Goldberg, E.D., 310-313, 327-330, 341,
Gal, M., 775 342, 735, 743, 769, 784
Galan, E., 686,687, 742 Goldich, S.S., 55, 496, 724, 743
Galdn, E.,665, 760 Goldman, C.R., 734
Galle, O.K., 729 Goldschmidt, 444, 742
Galliher, E.W., 390, 742 Goldschmidt, H., 409, 629, 742
794

Goldstein, A.P., 740 Hagemann, F., 584, 587, 745


Goll, R.M., 746 Hajek, B.J., 172, 753
Goodwin, S., 766 Hall, A., 628, 745
Gorbunova, Z.N., 337, 743 Hall, A.M., 408, 745
Goulding, K.W.T., 21, 778 IIall, F.L., 24, 724
Govett, G.J.S., 568, 569, 743 Hall, G.F., 783
Gradstein, F.M., 750 Hall, P.L., 8, 724
Gradusov, B.P., 154, 160, 174, 175, 584, Hallam, A., 649, 666, 745
588, 744, 768, 769, 771 Hamilton, R., 115, 745
Grady, W.C., 615, 743 Hancock, N . J . , 553, 557, 745
Graff-Peterson, 650, 743 Hann, H.H., 208, 745
Graham, S.A., 603, 743 Hanson, R.F., 752
Greenberg, S.A., 125, 743 Harder, H., 56, 126, 177, 178, 395, 745
Greene-Kelly, R., 50, 53, 744, 783 Hardie, L.A., 270, 739
Greenland, D.J., 136, 213, 743 Harris, L.D., 490, 739, 745
Griffin, G.M., 157, 158, 200, 233, 241, Harris, W., 399, 400, 745
252-254, 313, 323, 329, 330, 676, 743, Harrison, J.L., 600, 737
758 Harrison, S.C., 246, 745
Grifin, J., 769, 784 IIarrold, P.J., 785
Griffin, J.J., 153, 196, 308, 310-313, 316- Hart, S.R., 355, 373, 745, 773, 776
318, 320, 322, 323, 327, 334, 338, 341- Hartwell, J.W., 748
343, 378, 733, 735, 736, 743, 766 Harvey, R.D., 446, 715, 741, 745
Griggs, G.B., 737 Harward, M.E., 83, 733, 745, 778
Grim, R.E., 5, 7, 13, 14, 19, 32, 33, 40, Harwood, M.E., 732
46, 51, 52, 242, 252, 563, 589, 601, 677, Hassouba, H., 183, 406, 745
743, 751, 779 Hatcher Jr., R.D., 500, 746
Grimrn, W.E., 580, 743 Hathaway, J.C., 84, 98, 99, 168, 230-233,
Gronwald, K., 771 327, 380, 733, 740, 746, 755, 767, 785
Groot, J.J., 664, 743 Hauff, P.L., 617, 746
Gross, D.L., 267, 743 Haven, D.S., 225, 716
Grossman, R.B., 743 Hawkins, D.B., 432, 746
Grout, F.F., 8, 743 Hawkins, G.P., 662, 746
Cruner, J.W., 14, 32, 82, 743, 744 Hawley, J.W., 743
Gruner, U., 741 Hay, R.L., 271, 277, 674, 746, 750, 777
Gude 111, A.J., 674, 773 Hayashi, H., 21, 777
Guerrero, J., 738 Hayes, J.B., 68, 166, 330, 331, 543, 584,
Guggenheirn, S., 31, 743 746, 762
Guinness, E.A., 729 Heacock, R.L., 748
Guldrnan, S.G., 271, 746 Heath, G.R., 331-336, 338, 746
Gunther, P.R., 767 Heckel, P.H., 618, 773
Gunthier, P.R., 741 Heezen, B.C., 342, 746, 749
Gushchina, E.B., 754 Heirn, D., 666, 746
Guthrie, J.M., 522, 523, 743 Hein, F.J., 230, 746
Gutjahr, C.C.M., 748 IIein, J.R., 338, 339, 368, 375, 376, 7.16
Gutschick, R.C., 599, 745 IIekinian, R., 319, 369, 372, 746, 748
Giiven, N . , 37, 46, 51, 276, 535, 677, 743, Helgeson, H.C., 129, 723
745 Heling, D., 419, 424, 686, 736, 747, 767
Heller, L., 726
IIadley, R.F., 195, 771 Heller-Kallai, L., 644, 747, 775
795

Helmold, K.P., 546-549, 747 Honnorez, J., 371, 748


Hendricks, S.B., 3, 32, 50, 52, 54, 88, 723, Hood, A., 517-521, 748
747, 771 Hooker, P.J., 740
Heng, Y.Y., 146, 739 Hooks, W.C., 634, 748
HCnin, S., 77, 731 Hoover, K.V., 585, 738
Henley, J.J., 448, 747 Horne, J.C., 782
Henmi, T., 25, 747 Hossner, L.R., 738, 773
Henry, V.J.Jr.., 193, 747 Hosterman, J.W., 589, 591, 614, 748, 765,
Henson, M.R., 636, 747 769
Hepburn, J.C., 616, 747 Hostermann, J.W., 729
Herbillon, A.J., 24, 143, 156, 165-167, Hostetler, P.B., 733
747, 758, 760, 773 Hostet tler , J., 778
Heron Jr., S.D., 664, 678, 747 Houseknecht, D.W., 743
Hdroux, Y., 513, 747, 755 Howard, J.J., 552, 748
Herzog, L., 398, 399, 747 Howard, P., 123, 742
Hey, M.H., 73, 747 Howell, B.A., 225, 727
Hickel, D., 450, 737 Hower, J., 9, 10, 41, 42, 57, 87-92, 158,
Higashi, S., 76, 774 390, 399, 419-421, 423, 425, 427-432,
High Jr., L.R., 266, 633, 747, 766 435, 437, 440, 443, 458, 463, 464, 496,
Highsmith, P.B., 783 499, 500, 512, 569, 573, 576, 656, 677,
Hillyer, J.W., 724 724, 738, 748, 749, 758, 766, 770, 779
Hiltabrand, F., 448, 747 Hower, M.E., 748
Hinch, H.H., 421, 424, 706, 747 Hower, W.F., 745
Hinckley, D.N., 19, 21, 29, 747 Howie, R.A., 735
Ho, C., 8, 747 Howland, R.J.M., 761
Hobbs, B.E., 478, 510, 739, 748 Hsu, P.H., 77, 171, 172, 748, 749
Hobbs, P.V., 300, 781 Huang, P.M., 172, 174, 743, 749
Hodge, G.D., 777 Huang, W.H., 125, 134, 749, 752
Hodge, T., 184, 406, 748 Hubert, J.F., 777
Hoeve, J., 661, 748 Huck, G., 517, 749
Hoffert, M., 319, 338, 367-370, 372, 746, Huertas, F., 135, 756
748 Huff, W.D., 63, 64, 426, 584, 749
Hoffman, J., 440, 656, 748 Huggett, J.M., 540, 541, 749
Hoffman, U., 783 Huggett, R., 769
Hofmann, B., 446, 766 Huggins, C.W., 101, 749
Hofmann, U., 83, 748 Hughes, C.P., 579, 783
Hogberg, R.K., 656, 765 Hughes, C.R., 757
Holdoway, K., 627, 748 Hughes, R.E., 614, 615, 749
Hole, F.D., 728 Hull, R.W., 753
Holeman, J.N., 189, 198, 285, 748 Hullings, N.C., 738
Holland, H.D., 352, 762 Humphris, S.E., 357, 746, 749
Hollister, C.D., 648, 664, 694, 696, 727, Huneke, J.C., 493, 749
748, 755 Hunt, C.B., 106-108, 110, 749
Holman, M.G., 756 Hunter, B.E., 735
Holser, W.T., 408, 748 Hunter, K.A., 213, 214, 749
Holtzapfel, T., 733 Hunziker, J.C., 398, 502, 503, 507, 741,
Homann, W., 736 749, 764
Honea, R.M., 157, 781 Hurley, P.M., 312, 320, 398, 399, 496, 748,
Honjo, S., 280, 739 74 9
796

Hurst, A,, 534, 539, 540, 553, 650, 749 Joenson, O., 102, 380, 728
Hurst, C.W., 712, 761 Johns, W.D., 19, 242, 252, 317, 320, 321,
Hurst, V.J., 570, 749 690, 727, 743, 751
Hutchinson, G.E., 265, 749 Johnson Jr., H.S., 747
Hyme, N.J., 734 Johnson, L.J., 79, 149, 166, 751
Hyne, N.J., 268, 749 Johnson, L.R., 730, 733
Johnson, N.M., 125, 751
Ibe, K., 781 Johnson, R.G., 214, 297, 751
Ichikawa, A., 78, 749 Jones, B.F., 270, 272, 273, 735, 738, 751,
Iglesia, A.La., 650, 731 753
Iglesia Fernandez, A., 128, 749 Jones, L.H., 169, 751
Iijima, A., 32, 75, 445, 547, 674, 750 Jones, R.L., 717, 718, 751
Ildefonse, P., 161, 167, 750 Jordan, R.R., 678, 751
Imam, M.B., 541, 750 Jorgensen, P., 752
Ingersoll, R.V., 743 Jourdan, A., 733
Ingram, R.L., 6, 7, 233, 634, 743, 748, 750 Juster, T.C., 44, 751
Iniguez, A.M., 623, 750 Juteau, T., 356, 357, 751
Insby, H., 723
Ireland, B.J., 393, 734, 750 Kaiser, W.R., 537, 543-545, 751, 752
Irwin, H., 534, 539, 553, 749 Kames, W.H., 732
Ismail, F.T., 84, 156, 162, 750 Kanaris-Sotiriou, R., 624, 775
Isphording, W.C., 94, 750 Kane, J.E., 293, 752
Ivarson, K.C., 138, 615, 750 Kantor, W., 144, 155, 752
Iwasihska, I., 783 Kantorwicz, J., 557, 650, 752
Kao, C.-C., 77, 171-173, 730
Jaakkimainen, M., 754 Kaoru, O., 328, 330, 772
Jack, R.N., 729 Karlsson, W., 680, 752
Jackson, M.L., 25, 53, 109, 121, 154, 158, Karpoff, A.M., 748
159, 161-164, 166, 168, 171-174, 421, Karpova, G.V., 452, 458, 622, 752
723, 724, 728, 730, 733, 736, 750, 756, Karweil, J., 517, 736, 749, 752
768, 771, 772, 778, 779, 783 Kastner, M., 89, 102, 353, 371, 376, 380,
Jackson, T.A., 140, 750 381, 385-387, 399, 695, 726, 738, 752,
Jacobs, M.B., 260, 261, 293, 322, 324, 338, 753, 774
701, 750 Kator, N.V., 741
Jagodzinksi, H., 59, 750 Katz, M.Ya., 774
Jakobsson, S.P., 372, 750 Kautz, K., 277, 752
James, H.L., 569, 750 Kazakov, G.A., 774
Janke, N.C., 167, 767 Keene, J.B., 387, 752
Jansa, L.F., 654, 665, 666, 679, 694, 750, Keith, M.L., 735
751 Keler, W.D., 674, 753
Jarman, C.B., 655, 751 Keller, G.H., 281, 726, 752
Jaynes, W.F., 727 Keller, W.D., 22, 23, 30, 43, 103, 123-126,
Jeans, C.V., 629, 636-638, 642, 666, 667, 134, 140, 147-149, 439, 441, 610, 614,
740, 751 615, 646, 673, 749, 750, 752, 753, 767
Jednatok, J., 758 Kelley, J.T., 247, 753
Jeffrey, L.M., 783 Kelly, J.C., 754
Jennings, S., 432, 779 Kelts, K., 371, 738, 753
Jenny, H., 112, 116, 751 Kemp, A.L.W., 778
Jepson, W.B., 23, 751 Kennedy, V.C., 191, 192, 345, 753
797

Kent, O.B., 381, 753 Konta, J., 176, 205, 206, 754
Keppens, E., 400, 753 Konwar, L., 743
Keramidas, V.Z., 175-177, 740 Konyukov, A.K., 419, 424, 778
Kerr, P.F., 13, 23, 25, 30, 52, 54, 276, 745, Koopman, B., 315, 772
753, 771 Kopp, O.C., 79, 627, 754
Khain, V.E., 771 Kossovskaya, A.G., 44, 408, 445,447, 452,
Khalaf, F.I., 743 541, 629, 695, 696, 754
Kharaka, Y.K., 536, 537, 539, 753 Kostecki, J.A., 754
Kheirov, M.B., 769 Kotel’nikes, D.D., 622, 769
Kheirov, m.B., 768 Kotel’nikov, D.D., 79, 772, 778
Kheisov, M.B., 769 Kotov, N.V., 755
Kheoruenromne, I., 742 Kotovich, V.A., 755
Kholodkevich, I.V., 755 Koyumdjisky, H., 785
Khoury, H., 738 Kraeuter, J.N., 245, 754
Khoury, H.N., 94, 276, 674, 738, 753 Kramer, H., 724
Kiersch, G.A., 674, 753 Kramers, J.W., 657, 731
Kimbara, K., 80, 753 Kranck, K., 214-216, 219, 754
Kimberley, M.M., 396, 753 Kraus, I., 689, 754, 774
Kimura, H.S., 761 Krauskopf, K.B., 125, 754
Kind, P.B., 740 Krishna Murti, G.S.R., 155, 769, 775
Kirkland, D.L., 172, 753 Kristmannsdotter, H., 352, 779
Kisch, H.J., 417, 418, 445, 447-449, 458, Kristmannsdottir, H., 79, 353, 546, 754
490, 520, 753 Kristmannsdottir, M., 446, 779
Kishk, F.M., 84, 85, 161, 169, 726 Kronberg, B., 764
Kistler, R., 739 Krone, R.B., 246, 738
Kitagawa, Y., 25, 753 Kroner, A., 398, 733
Kitano, Y., 755 Krumm, H., 641, 755
Klameth, L.C., 762 Krylov, A.Ya., 312, 398, 755
Klanderman, E.S., 764 Kubler, B., 10, 61, 409, 418, 444, 448-450,
Klein, G.DeV., 636, 753 452, 507, 520, 582, 755
Klimentidis, R.E., 439, 753 Kuenen, Ph.H., 246, 248, 779
Klingebiel, A., 685, 754 Kulbicki, G., 51, 52, 685, 740, 743, 755
Klotchkova, G.N., 741 Kulbicki, G.L., 667, 771
Knebel, H., 760 Kulm, L.D., 737
Knebel, H.J., 204, 242-244, 754 Kumar, N., 288, 735
Knight, L., 705, 733 Kunze, G.W., 158, 244, 755
Knipe, R.J., 463, 478, 783 Kiinzi, B., 741
Knowles, L.L., 755 Kurnosov, V.B., 351, 352, 755, 775
Knox, E.G., 732 Kurzweil, H., 690, 751
Knox, R.W.O’B., 649, 680, 754
Kobayashi, K., 330, 764 La Iglesia, A., 135, 755
Koch, G., 783 Lagache, M., 448, 755
Kodarna, H., 43, 63, 134, 135, 137, 165, Lagaly, G., 46, 48, 755
169, 419, 424, 440, 458, 524, 657, 725, Laglesia, A., 742
740, 754, 771 Lahamm, R., 429, 431, 769
Kohyama, N., 29, 754 Lahann, R.W., 431, 432, 755, 761
Kohyana, N., 724 Lahav, N., 171, 755
Kokorina, L.P., 755 Laidi, L.W., 749
Kolla, V., 328, 342, 754 Lambert, D.N., 752
798

Lambert-Aikhionbare, D.O., 692, 755 Lisitzyn, A.P., 768


Lamphere, M.A., 492, 493, 735 Liss, P.S., 213, 749
Lancelot, Y., 186, 187, 755 Logvinenko, N.V., 398, 755
Land, L.S., 543, 755 Lombard, A.L., 726
Laplante, R.E., 518, 755 Lomova, O.S., 400, 754, 757
Laputina, I.P., 771 Long, G., 643, 757
Lashkov, E.M., 582, 587, 755 Long, L.E., 398-400, 761
Lashkova, L.N., 755 Longchambon, H., 98, 757
Lassin, R.J., 764 Longstaffe, F.J., 230, 540, 541, 543, 558,
Latouche, C., 187, 685, 688, 699, 754, 755 746, 757
Laverne, C., 724 Lonsdale, P., 366, 757
Lawrence, J.R., 356, 377, 379, 743, 756 Lopatin, N.V., 517, 518, 757
Le Pichon, X., 748 Lopez-Aguayo, F., 665, 757
Lebauer, L.R., 576, 756 L6pez Aguayo, F., 764
Leckie, G.G., 576, 593, 601, 617, 756 Lopez Azcona, M.C., 758
Leckie, J.O., 756 L6pez De Azcona, M.C., 668, 761
Ledbetter, M.T., 375, 756 Losievskaya, S.A., 774
Lee, J.H., 419, 421, 433, 481-483, 557, Louail, J., 667, 757
629, 756 Loughman, F.C., 30, 757
Lee, M., 421, 481-483, 557, 629, 756 Loughnan, F.C., 107, 757
Lee, M.J., 627, 756 Love, S.K., 192, 757
Lee, S.Y., 21, 174, 749, 756 Lovering, T.S., 142, 757
LeFkvre, J., 727 L U C ~ SJ.,
, 408, 571, 638-640, 642, 643,
Leininger, R.K., 158, 762 668, 692, 729, 757, 779
Lelong, F., 143, 756 Ludmila, P., 755
Lenzi, S.R., 630, 729 Lyle, M., 734, 738
Leonard, A.B., 742 Lynch, L., 420, 757
Leonard, R.A., 162, 756 Lynch-Blosse, M.A., 729
Leopold, L.B., 190, 191, 756 Lynn, W.C., 174, 757, 762
Lerbekmo, J.F., 661, 731 Lyons, S.C., 19, 20, 762
Lerman, A., 350, 756
Lever, A., 696, 756 MacBustin, R., 661, 679, 757
Levinson, A.A., 657, 661, 726 Macchi, L., 553, 757
Levison, A., 757 MacEwan, D.M.C., 22, 34, 57, 70, 424,
Lewis, A.G., 224, 295, 296, 756, 777 730, 757, 758
Lewis, C.F.M., 267, 756, 778 Machajdik, D., 48, 733
Lewis, D.D., 756 Mackenzie, F.T., 122, 147, 300, 309, 350,
Lewis, S.R., 634, 756 513, 732, 742, 756, 758, 784
Liebling, R.S., 590, 756 Mackenzie, R.C., 5, 56, 183, 758
Liewig, N., 452, 756 Mackin, J.E., 350, 386, 758
Likens, G.E., 751 Mackinnon, I.D.R., 439, 753
Linares, J., 135, 756 Mackowsky, M.-Th., 776
Lindsey, D.A., 569, 756 MacNeil, S., 638, 758
Lineback, J.A., 733, 743 Mahjoory, R.A., 179, 758
Lion, L.W., 349, 756 Mahoney, J.J., 380, 762
Lippmann, F., 79, 409, 638, 757 Maisano, M.D., 675, 758
lisitsyna, N.A., 731 Maksimovib, Z., 94, 758
Lisitzin, A.P., 280, 282-285, 287, 301, 302, Makumbi, L., 165, 758
308, 309, 317, 757 Makumbi, M.N., 156, 166, 167, 747
799

Malachoff, A., 368, 369, 758 McHardy, W.J., 59, 533, 535, 553, 554,
Malcolm, R.L., 159, 174, 758 740, 758, 762, 784
Malden, P.J., 23, 24, 758, 760 McKeague, J.A., 159, 167, 758
Malla, P.B., 50, 758 McKee, E.D., 6, 617, 746, 758
Mamy, J., 180, 742 McKee, T.R., 27, 143, 736
Mangelsdorf Jr., P.C., 346-348, 772 McKenzie, J., 177, 736
Manghnani, M.H., 89, 90, 758 McKyes, E., 131, 758
Manivit, H., 729 McLean, S.A., 98, 276, 758
Manker, J.P., 241, 758 McMurdie, H.F., 723
Mankin, C.J., 43, 758 McMurthy, G.M., 52, 54, 366-369, 371,
Manley, F.H., 327, 581, 758 758, 760
Marchadour, P., 737 McNeely, L., 569, 778
Markovris, G., 762 McPherson, H.J., 192, 760
Marshall, C.E., 437, 758 McRae, S.G., 87, 760
McVeagh, W.J., 737
Martin, B.D., 387, 769
Meade, R.H., 189, 190, 220, 221, 224, 279,
Martin, F.M., 687, 758
285, 705, 760
Martin, I., 650, 758
Meads, R.E., 23, 24, 758, 760
Martin, J.M., 207, 758 Means, W.D., 748
Martin, R.T., 50, 708, 725, 759, 762 Mehmel, M., 28, 760
Martin Vivaldi, J.L., 70, 98, 128, 135, 642, Meisner, J., 94, 723
665, 731, 749, 755, 757, 758
Mdlikres, F., 732
Mashhady, A.S., 738, 758 Mellon, G.B., 661, 731
Masood, H., 661, 758 Melson, W., 375, 736
Masse, J.-P., 665, 732 Melson, W.G., 352, 353, 360, 760
Massoud, M.S., 743 MCndez, P.A., 665, 760
Masui, J., 160, 174, 378, 379, 758 Menillet, F., 775
Mather, A., 87, 88, 93, 387-391, 764 Merino, E., 558, 760
Mathieson, A.McL., 82, 758 Mermut, A.R., 174, 760
Matoba, M., 738 Merrill, W.M., 600, 602, 760
Matsumoto, R., 32, 75, 750 Merriman, J.G., 751
Mattey, D.P., 746, 772 Mestdagh, M.M., 24, 747, 760
Matti, J.C., 384, 385, 758 Meunier, A., 133, 178, 760
Maxwell, D.T., 41, 458, 569, 758 Middleton, M.F., 519, 760
Maynard, J.B., 767 Midgley, H.G., 98, 760
McAllister Jr., R.F., 252, 758 Migdisov, A.A., 568, 771
McBride, M.B., 24, 758 Miles, N.M., 750
McCave, I.M., 696, 756 Milici, R.C., 490, 745
McCave, I.N., 280, 282-285, 290, 291, 293, Miller, B.E., 540, 742
758 Miller, J.A., 740
McCoy Jr., F.W., 761 Miller, J.P., 756
McCracken, R.J., 758 Miller, R., 327, 760
McCreery, R.A., 137, 778 Miller, R.H., 781
McCrone, A.W., 349, 758 Milliman, J.D., 189, 190, 279, 285, 322,
McCulloh, T.H., 706, 758 324, 739, 760
McDowell, S.D., 75, 446, 487, 488, 491, Millot, G., 7, 79, 84, 98, 120, 143, 152,
758 160, 169, 179, 182, 183, 389, 390, 409,
McDuff, R.E., 379, 758 577, 650, 665, 666, 669, 682, 685, 690,
McGown, A., 710, 734 740, 760, 765, 774, 778
800

Milne, A.A., 751 Muller, J., 774


Milne, I.H., 77, 78, 252, 349, 671, 761, 775 Muller, J.-P., 146, 762
Minch, J.A., 723 Mullis, J., 741
Mingarro Martin, P.F., 668, 761 Mulvaney, R., 734
Mintz, L.W., 572, 579, 593, 598, 626, 631, Munoz, E.F., 140, 774
646, 653, 672, 761 Murata, K.J., 55, 740
Miser, H.D., 771 Muratli, C.M., 737
Mitchell, B.D., 740 Muravitskaja, G.N., 771
Mitchell, J.G., 751 Murray, H.H., 19, 20, 158, 446, 512, 570,
Mitchell, W.A., 166, 761 610, 693, 730, 751, 762, 767
Mitchum, R.M., 779 Murray, H.M., 673, 676, 762
Miyashiro, A., 350, 352, 761 Murray, J.W., 225, 226, 230, 777
Miyauchi, N., 143, 724 Mustoe, G.E., 766
Moberly Jr., R., 330, 761
Mohr, D.W., 179, 741 Nadeau, P.H., 53, 59, 166, 439, 762, 784
Mohr, E.C.J., 120, 143, 761 Nadeau, P.N., 426, 658, 762
Moiola, R.J., 271, 746 Nadler, L., 754
Molina-Cruz, A., 738 Nagy, B., 95, 762
Monciardini, C., 729 Nahon, D., 765
Moncure, G.K., 545, 664, 761 Naidu, A S . , 330, 762
Mongiorgi, R., 56, 761 Nakozawa, K., 764
Montadernt, L., 700, 769 Nanz, R.H., 512, 762
Montague, C.L., 225, 761 Narkis, N., 214, 762
Moon, C.F., 712, 761 Nathan, Y., 747, 785
Moore, D., 738 Natland, J., 746
Moore, D.G., 8, 284, 773 Natland, J.H., 380, 694, 724, 762
Moore, H.B., 225, 297, 761 NeacSu, G., 651, 762
Moore, J.E., 267, 761 NeacSu, V., 651, 762
Moore, J.G., 372, 373, 750, 761 Neal, W.J., 784
Moore Jr., T.C., 746 Needham, H.D., 748
Morales-Alamo, R., 225, 746 Neglia, S., 757
Morandi, N., 56, 761 Neiheisel, J., 202, 215, 232, 233, 235-239,
Morey, G.B., 569, 761 241, 762, 763
Morgues, F., 98, 757 Neira, E., 724
Moriarty, K.C., 339, 761 Nelson, B.W., 74, 233, 262, 615, 762
Morris, A.W., 350, 761 Nelson, C.H., 268, 762
Morris, R.C., 600, 761 Nelson, D.D., 408, 599, 762, 766
Morton, J.P., 398-400, 427, 428, 677, 761 Nelson, R.E., 762
Morton, R.W., 223, 761 Nelson, W.A., 584, 762
Mossler, J.H., 584, 762 Nelson, W.G., 749
Mottl, M.J., 352, 353, 762 Neri, R., 725
Motto, H.L., 764 Nesbitt, H.W., 125, 569, 762
Moulin, C., 726 Nesterhoff, W.D., 746
Mowatt, T.C., 42, 330, 748, 762 Nesteroff, W.D., 385, 762
Muehlenbachs, K., 724 Nettleton, W.D., 111, 157, 179, 183, 758,
Muffler, L.J.P., 445, 762 762
Muller, A., 639, 762 Neuzil, S.G., 732
Muller, G., 526, 742 Newnham, R.E., 13, 762
Muller, G., 267, 270, 418, 762, 775 Nichols, M.M., 223, 762
80 1

Nicolai, V., 755 Padgham, R.C., 315, 733, 765


Niederbudde, E.A., 179, 762 Pak, H., 731
Nielson, O.B., 680, 762 Palacio, M.R., 693, 765
Niem, A.R., 601, 762 Palmer, D.P., 409-411, 627, 628, 726, 765
Niemitz, J., 738 Papke, K.G., 98, 276, 673, 765
Niggli, E., 452, 520, 741 Paquet, H., 120, 155, 156, 160, 169, 179,
Nishiyama, T., 66, 762 182, 185, 760, 765, 778
Noack, Y., 751 Parachoniak, W., 624, 765
Norrish, K., 46, 152, 161-163, 181, 183, Parfenova, E.I., 138, 140, 177, 765
762, 769, 775 Parham, W.E., 581, 613, 656, 765
Northrop, H.R., 738 Parker, R.D., 675, 777
Nortland, M.M., 179, 724 Parkin, D.W., 315, 765
Nosov, G.L., 768 Parkins, D.W., 735
Novak, R.J., 151, 764 Parmenter, C.M., 767
Novick, B.E., 50, 762 Parrish, J.T., 631, 765
Nuhfer, E.B., 713, 764 Parron, C., 87, 91, 724
Parrott, B.S., 252, 743
Oades, J.M., 748 Parry, W.T., 98, 276, 765
Oberlin, A., 746 Parting, H., 762
Obradovich, J., 739 Partridge, P., 693, 762
O’Brien, N.R., 6, 244, 288, 614, 707-709, Pashai, A., 183, 723
764, 775 Pasteels, P., 400, 753
Odin, G.S., 32, 87, 88, 93, 387-392, 397, Paterson, M.S., 739
398, 764, 779 Patterson, E.M., 305, 765
Odom, I.E., 392, 559, 712, 764 Patterson, S.H., 614, 673, 676, 678, 748,
Ogren, D.F., 758 762, 765
Ohmacht, W., 769 Pe-Piper, G., 654, 665, 666, 679, 751
Oinum, K., 330, 764 Peacor, D.R., 419-421,427,432,433,436-
Oinuma, K., 328, 724 440, 443, 446, 447, 533, 723, 728, 756,
Ojanuga, A.G., 146, 764 785
Okada, H., 385, 764 Pearl, H.W., 225, 294, 765
Okusami, T.A., 144, 764 Pearson, M.J., 419,425, 524, 650, 680, 765
Oleg, V., 755 Pedio, G., 273, 765
Olorunfemi, B.N., 256, 764 Pelzer, E.E., 592, 600, 765
O’Melia, C.R., 209-212,214, 233,235, 738 Peppers, R.A., 766
Ordonez, S., 687, 764 Perkins, H.F., 147, 740
Orlopp, D.R., 782 Perrin, R.M.S., 588, 594, 620, 629, 649,
Ortega-Huertas, M., 687, 764 665, 666, 680, 766
Orth, C.J., 767 Perry Jr., E.A., 337, 419, 423, 427, 439,
Osborne, R.H., 581, 728 496, 677, 766, 777
Ourisson, G., 522, 723 Perry, G.A., 748
Ovejero, R., 630, 764 Perseil, A., 748
Ovenshine, A.T., 8, 724 Person, A., 748
Overstreet, R., 778 Persoz, F., 651, 665, 766
Peters, Tj., 446, 766
Packman, G., 724 Petersen, N., 746
Padan, A., 93, 94, 97, 408, 409, 411, 412, Peterson, G.L., 723
414, 617, 765, 783 Peterson, M.N.A., 79, 378, 384, 409, 602,
Padgett, G., 732 766
802

Petrov, V.P., 571, 651, 766 Price, E.W., 125, 743


Petruk, W., 71, 766 Price, N.B., 624, 767, 769
Pettijohn, F.J., 5-9, 581, 720, 766 Price, R.C., 762
Pevear, D., 21, 731 Prospero, J.M., 300-303, 316, 339, 767
Pevear, D.R., 673, 766 Protz, R., 771
Philippi, G., 518, 766 Proust, D., 168, 768
Phillips, R.E., 772 Pryor, W.A., 246, 297-300, 389, 393, 662,
Phillips, T.L., 618, 766 767, 768
Picard, M.D., 7, 266, 633, 714, 716, 734, Purton, M.J., 620, 768
747, 766 Purucker, M.E., 395, 396, 779
Pickering, J.G., 152, 162, 163, 181, 762 Pusey, W.C., 518, 523, 768
Pickett, E.E., 752
Pierce, J.W., 288, 293, 324, 325, 766 Quaide, W.L., 671, 768
Pierce, R.S., 751 Quakernaat, J., 262, 267, 762, 768
Pillmore, C.L., 767 Quigley, R.M., 708, 768
Pinsak, A.P., 767 Quinn, A.W., 520, 768
Pinson, W.H., 748, 749 Quirk, D., 748
Pinson Jr., W.H., 747, 749 Quirk, J.P., 769
Piper, D.J., 204, 767
Pirlet, H., 624, 779 Raam, A., 550, 768
Pittion, J.L., 755 Rabenhourst, M.C., 168, 768
Pittman, E.D., 525-531, 533, 784 Radoslovich, E.W., 13, 28, 33, 35-38, 71,
PlanGon, A., 19, 767 768
Pollard Jr., C.O., 419, 767 Ramamohana, R., 486, 487, 768
Pollard, L.D., 3, 30, 41, 51, 65, 66, 76, 79, Ramon, K.V., 163, 421, 768
88, 91, 95, 98, 100, 102, 366, 368, 380, Ramos, A.N., 597, 623, 630, 644, 768
431, 450, 783 Ramseyer, K., 428, 429, 768
Pollastro, R.M., 420, 553, 674, 767, 772 Ransom, B., 558, 760
Pomeraol, C., 779 Rao, C.G., 409, 768
Poncelet, G.M., 145, 767 Rao, C.R., 615, 768
Ponder, H., 673, 767 Rapaire, J.L., 740
Poppe, L.J., 323, 767 Rashid, M.A., 136, 768
Poppi, L., 79, 84, 729 Rateev, M.A., 263, 310, 322, 327, 334-
Porada, H., 277, 752 336, 338, 366, 368, 369, 577, 580, 584,
Porrenga, D.H., 43, 255-257, 297, 767 588, 596, 622, 768, 769, 779
Porthault, B., 652, 665, 767 Rateyev, M.A., 66, 769
Post, J.L., 100, 167, 767 Raup, O.B., 616, 618, 627, 769
Potter, P.E., 610, 712, 713, 716, 720, 721, Rawsthorn, K., 748
743, 766, 767 Rayner, J.H., 93, 769
Potter, p.E., 8, 104, 158, 195, 197-201, Rebhum, M., 762
76 7 Reesman, A.L., 752
Powell, T.G., 419, 440, 542, 740, 741, 767 Reeves, C.C., 98, 276, 765
Power, P.E., 602, 616, 767 Rehmer, J.A., 616, 747
Powers, M.C., 418, 767 Reiche, P., 103, 769
Pravdic, V., 758 Reichenbach, H.G., 175, 769
PravdiC, V., 207, 767 Reimann, B.E.F., 735
Preisinger, A., 95, 96, 729 Reineck, H.E., 248, 769
Presley, M.W., 736 Reinhold, M.L., 733
PrkvBt, L., 729, 757 Rengarten, N.W., 769
803

Rengasamy, P., 23, 24, 769 Ronov, A.B., 567, 568, 583, 588, 629, 771,
Renton, J.J., 269, 615, 736, 769 781
Rettke, R.C., 655, 769 Roquemore, S.K., 678, 771
Rex, R.W., 327, 328, 385-387, 733, 737, Rosendahl, B.R., 746, 772
766, 769 Ross, C.A., 603, 606, 771
Reynolds, R.C., 57, 60, 61, 120, 121, 166, Ross, C.P., 570, 771
424, 458, 570, 752, 769 Ross, C.S., 3, 13, 23, 25, 30, 50, 52, 54,
Reynolds Jr., R.C., 57, 420, 426, 658, 757, 57, 88, 662, 747, 771
762, 770, 776 Ross, E.H., 736
Reynolds, W.R., 678, 769, 771 Ross, G.J., 159, 165, 750, 771
Rhoads, D.C., 225, 769 Ross, M., 740
Ribbe, P.H., 781 Rossel, N.C., 553, 555, 771
Rice, D.D., 654, 770 Roth, C.B., 86, 771
Rich, C.E., 171, 726 Rouxhet, P.G., 773
Rich, C.I., 77, 84, 162-164, 168, 169, 171, Rowse, J.B., 23, 751
172, 174, 175, 734, 749, 769, 770 Rowsell, D.M., 597, 630, 771
Richman, D.L., 537, 752 Roy, A.B., 478, 483, 771
Rickard, R.S., 727 Roy, B.B., 155, 771
Riddick, T.M., 207, 769 Roy, R., 74, 432, 746, 762
Riecken, F.F., 115, 783 Rueda, J., 738
Rieke, H.H., 705, 707, 708, 769 Rueger, B.F., 617, 627, 728
Riley, J.P., 313, 770 Ruellen, A., 760
Rimmer, S.M., 613, 614, 769 Ruiz, A.R., 757
Risacher, F., 275, 725 Rule, A.C., 45, 771
Risebrough, R.,323, 324, 769 Rumeau, J.L., 667, 771
Rispens, F.B., 350, 735 Ruppert, L.F., 732
Ristori, G.G., 732 Rusinov, V.L., 358-361, 771
Roaldset, E., 680, 769 Russell, J.D., 160, 725
Roberson, H.E., 349, 429, 431, 432, 676, Russell, K.L., 242, 346, 771
677, 728, 755, 769 Rust, B.R., 267, 733
Robert, C., 699, 701, 769 Rutstein, M.S., 582, 771
Robert, M., 163, 769 Ruxton, B.P., 772
Roberts, D.G., 700, 769 Ryan, W.B.F., 263, 781
Roberts, G.L., 746 Ryer, T.A., 657, 772
Robertson, K.R., 768
Robertson, R.H.S., 151, 742, 769 Sabatier, G., 746, 755
Robinson, F., 754 Saburo, A., 328, 330, 772
Robinson, G.C., 747 Sachs, P.L., 98, 99, 380, 746
Robinson, K., 28, 730 Saint-Marc, P., 727
Robinson, P.T., 354, 355, 769 Sakharov, B.A., 774
Rocha-Campos, A., 782 Salisbury, M.H., 736
Rochewicz, A., 534, 545, 770 Samson, H.R., 24, 772
Rodriguez, J., 724 Sand, L.B., 149, 178, 772
Rodriguez-Gallego, M., 764 Sandberg, C.A., 599, 745
Roen, J.B., 589, 769 Sanders, G.S., 733
Rogers, L.E., 98, 99, 769 Sarkisyan, S.G., 79, 444, 450, 772
Roggenthen, W., 746 Sarma, V.A.K., 769
Rohrlich, V., 32, 769 Sarntheim, M., 315, 772
Rolfe, B.N., 195, 771 Sato, M., 75, 777
804

Sauer, J.E., 782 Shannon, P.M., 551, 773


Saunders, A., 738 Sharples, E.J., 733
Savin, S.M., 427, 445, 458, 489, 570, 739, Shaw, D.B., 8 , 9, 252, 715, 716, 718, 773
756 Shaw, D.M., 512, 773
Sawhney, B.L., 53, 158, 419, 729, 772 Shaw, H.F., 183, 338, 365, 366, 406, 541,
Sayegh, A.H., 745 680, 692, 734, 745, 750, 755, 773
Sayles, F.L., 338, 346-348, 772 Shayan, A., 54, 773
Schatz, A., 140, 772 Shell, H.R., 749
Scheidegger, K.F., 358, 361, 772 Shelton, J.W., 540, 773
Schlee, J.S., 570, 749 Shepard, A.O., 773
Schlenker, B., 638, 772 Shepard, F.P., 7, 8, 285, 773
Schminke, H.-U., 769 Sheppard, A.O., 733
Schneider, H., 91, 772 Sheppard, R.A., 674, 773
Schnitzer, M., 137, 754 Sheridan, R.E., 750
Schoen, R., 75, 586, 772 Sherman, G.D., 118, 119, 145, 774, 778
Schofield, R.K., 24, 772 Sherp, H.S., 590, 756
School, D.W., 375, 376, 746 Shevchenko, A.Ya., 775
Schoonmaker, J., 732 Shido, F., 761
Schrader, E.L., 368, 369, 746, 772 Shimoda, S., 66, 78, 80, 151, 152, 160, 749,
Schrader, H., 738 762, 774, 777
Schrock, R.L., 633, 772 Shimp, N.F., 743, 767
Schubel, J.R., 220, 772 Shirozu, H., 76, 446, 774
Schubert, C., 767 Shirshov, P.P., 337, 743
Schultz, D.J., 8 , 728 Shoji, S., 160, 174, 378, 379, 758
Schultz, L.G., 21, 51, 174, 524, 603, 610, Sholkovitz, E.R., 214, 774
611, 613, 632, 647, 658-661, 741, 743, Shott, W.L., 349, 761
773 Shoval, S., 27, 28, 785
Schutter, S.R., 618, 773 Shover, E.F., 607, 608, 774
Schwertmann, U., 144, 155, 752 Shrock, R.R., 7, 774
Schwizer, B., 741 Shutov, V.D., 452, 541, 577, 622, 754, 774
Sclater, J.G., 783 Shutove, V.D., 396, 774
Scotese, C.R., 725, 765 Siderenko, G.A., 52, 728
Scotford, D.M., 8, 512, 581, 773 Siebert, R.M., 761
Scott, T.M., 586, 773 Siefferman, G., 740
Scruton, P.C., 284, 773 Sieffermann, G., 143, 774
Sears, D.A., 6, 773 Siever, R., 353, 743, 766, 774
Sedivy, R.A., 491, 773 Siffert, B., 126-128, 275, 277, 774
Sellors, B., 570, 739 Sigl, W., 137, 774
Sellwood, B.W., 649, 666, 680, 745, 773 Sikora, W., 132, 152, 178, 197, 776
Senkayi, A.L., 677, 773 Sib, Yu.I., 312, 755
Serebrennikova, N.D., 754 Silverman, M.P., 137, 140, 774
Serstevens, A.T., 164, 773 Simoneit, B., 738
Seslavinksy, K.B., 771 Simons, L.H., 676, 774
Sethi, A., 758 Sindelii, J., 689, 774
Seyfied, W.G., 355, 356, 773 Sinex, S.A., 149, 774
Shabtai, J., 755 Singer, A., 52, 102, 146, 179, 183, 268,
Shadfan, H., 690, 773 273, 274, 380, 400, 418, 774-776
Shani, U., 755 Singh, G., 155, 775
Shanks, W.C., 773 Singh, I.B., 248, 769
805

Sittler, C., 683-685, 775 Starkey, H., 346, 732


Skornyakova, N.S., 352, 775 Starkey, H.C., 773
Sladen, C.P., 680, 773 Staub, J.R., 269, 776
Slatt, R.M., 204, 767 Staudigel, H., 355, 373, 745, 776
Slaughter, M., 77, 78, 657, 775 Steinberg, M., 597, 623, 630, 733, 738
Sly, P.G., 266, 775 Steiner, A., 445, 776
Small, J.S., 765 Stephan, J.F., 388, 764
Smeck, N.E., 783 Stephen, I., 101, 776
Smillie, G.W., 160, 734 Stephens, D.G., 732
Smith, B.F.L., 740 Stephenson, L.W., 771
Smith, D.G.W., 589, 775 Sterne, E.J., 44, 776
Smith, E.C., 204, 215, 322, 735 StCvaux, J., 644, 776
Smith, J.V., 33, 37, 38, 775 Stevenson Jr., R.G., 663, 783
Smith, N.D., 268, 775 Stewart, F.H., 94, 409, 776
Smith, S., 778 Stewart, R.J., 758
Smith, W.H., 614, 775 Stindl, H., 627, 779
Smoluchowski, M., 209, 775 Stoch, E., 776
Smulikowski, K., 88, 775 Stoch, L., 132, 152, 178, 197, 776
Sndl, S., 587, 775 Stoessell, R.K., 277, 777
Snowdon, L.R., 767 Stoffers, P., 273, 274, 775, 776
Soil Survey Staff, 108, 775 Stokke, P.R., 288-290, 776
Sokolova, A.L., 774 Stolzman, R.A., 729
Sommer, F., 779 Stone, W.J., 675, 776
Sorokin, V.I., 397, 775 Stoner, J.H., 733
Southard, R.J., 157, 775 Storr, M., 152, 582, 621, 638, 651, 665,
Sparks, D.M., 517, 732 689, 777
Sparks, T.N., 254, 255, 736 Storr, M., 152, 582, 621, 638, 651, 665,
Spears, D.A., 146, 446, 624, 734, 775 689, 776
Speights, D.B., 607, 627, 775 Strakhov, N.M., 112, 297, 622, 777
Spencer, R.J., 272, 751 Stuart, C.J., 287, 777
Sperber, H., 762 Studer, M., 652, 777
Spjeldnaes, N., 584, 745 Stumm, W., 208, 745
Spooner, E.T.C., 353, 775 Sucheki, R.K., 79,444, 541, 543, 577, 578,
Spronck, R., 193, 775 777
Spry, A., 478, 776 Suddhiprakarn, A., 743
Spyridakis, D.E., 141, 776 Sudo, T., 21, 56, 75, 151, 152, 160, 446,
Sridhar, K., 166, 750 724, 777
Srivastava, R.K., 746 Suitch, P.R., 15, 777
Srodoh, J., 33, 39, 42, 43, 49, 58-61, 420, Summer, K.V., 157, 777
422, 423, 425, 426, 513, 515, 534-536, Summer, M.E., 24, 777
621, 624, 689, 738, 765, 776 Summerhayes, C.P., 760
Srbmek, J., 754 Sunderman, J.A., 737
Stakes, D.S., 358, 361, 772 Surdam, R.C., 537, 538, 675, 777
Standaert, R.R., 75, 408, 409, 587, 728 Suttner, L.J., 558, 737
Stanley, D.J., 280, 281, 776 Sverdrup, H.U., 328, 777
Stanley, K.O., 550, 551, 675, 776 Swain, F.M., 268, 777
Stanton, R., 769 Sweeney, J.W., 748
Stanton, R.W., 732 Swift, D.J.P., 283, 777
S t a p h , F.L., 517, 776 Swindale, L.D., 334, 386, 777
806

Swineford, A., 627, 777 Threadgold, I.M., 728


Syers, J.K., 733, 736, 771 Tillman, R.W., 526, 784
Sys, C., 155, 726, 739 Timofeev, P.P., 137, 384, 622, 769, 779
Syvitski, J.M., 268, 775 Timofeev, P.R., 769
Syvitski, J.P.M., 224-230, 251, 295, 296, Tissot, B.P., 519, 779
756, 777 Tixier, M., 652, 767
Szafranek, D., 775 Tolar, V.I., 754
Tomasson, J., 352, 446, 779
Tabikh, A.A., 437, 778 Tomita, K., 21, 58, 385, 764, 779
Taggard Jr., M.S., 676, 774 Tomlinson, C.W., 721, 779
Tait, J.M., 758, 762 Tompkins, R.E., 546, 779
Takahashi, H., 21, 58, 779 Tooker, E.W., 617, 779
Takahashi, J., 389, 778 Totten, M.W., 715, 717, 718, 728
Takuhashi, S., 764 Tourencq, J., 779
Talibuden, O., 21, 778 Tourtelot, H.A., 5, 9, 773, 779
Talwani, M., 375, 778 Towe, K.M., 589, 779
Tamura, T., 86, 143, 145, 778 Trauth, D., 748
Tan, K.H., 137, 778 Trauth, N., 618, 681, 682, 685, 692, 775,
Tank, R.W., 569, 647, 675, 680, 778 779
Tardy, Y., 133, 143, 146, 273, 735, 742, Trescott, P.C., 782
778 Triplehorn, D.M., 43, 388, 389, 779
Taskey, R.D., 151, 778 Trompette, R., 732, 757
Taylor, A.M., 553, 557, 745 Tschudy, R.H., 767
Taylor, G.H., 776 Tsipurskii, S.I., 731
Tazaki, K., 130, 778
Tucholke, B.E., 750
Tchoubar, C., 19, 767
Turchenek, L.W., 748
Tchukhrov, F.V., 57, 778
Turekian, K.K., 302, 314, 512, 779, 780
Teague, L.S., 551, 726
Turk, L.M., 108, 122, 741
Tedorovich, G.I., 419, 778
Tiirkmenoglu, A.D., 63, 64, 426, 584, 749
Teichmuller, M., 736
Turner, F.J., 417, 483, 779
Teichmiiller, M., 520, 741, 776, 778
Teichmiiller, R., 741, 776, 778 Turner, P., 587, 720, 779
Templin, E.H., 158, 755 Turner, R.C., 171, 730
Teodorovich, G.I., 419, 424, 778 Twenhofel, W.H., 5, 7, 389, 779
Terchunian, A., 743 Twiss, P.C., 627, 779
Thayer, P.A., 8, 778 Tyler, S.A., 569, 725
Theng, B.K.G., 22, 136, 778
Thiry, M., 737, 779 Udagawa, S., 783
Thomas, R.L., 267, 778 Udinstev, G., 778
Thompson, A.M., 581, 779 Urasin, M.A., 775
Thompson, C.D., 708, 768 Utada, M., 445, 547, 750
Thompson, G., 360, 371, 432, 736, 760,
779 Vacquir, V., 738
Thompson, G.R., 87-89, 91, 92, 399, 728, Vail, P.R., 396, 567, 779
779 Vail, R.S., 782
Thompson, R.N., 749 Valeton, I., 667, 779
Thompson 111, S., 779 Van Andel, Tj.H., 258, 779
Thorez, J., 621, 624, 762, 779 van Andel, Tj.H., 353, 760
Thorndike, E.M., 750 Van Baren, F.A., 120, 143, 761
807

Van de Kamp, P.C., 521, 546-549, 747, Wang, C., 771


779 Wang, C.-H., 760
Van den Heuvel, R.C., 98, 737, 779 Wanless, H.R., 609, 782
Van der Marel, H.W., 22-24, 84, 94, 322, Waples, D.W., 519, 782
738, 779, 781 Warde, J.M., 22, 782
Van der Merwe, C.R., 118, 179, 779 Warner, M.A., 524, 731
Van Houten, F.B., 143, 395, 396, 633, 720, Warren, N., 746
779 Warshaw, C.M., 87, 90, 91, 782
van Moort, J.C., 419, 761 Watanabe, T., 60, 62, 782
Van Olphen, H., 206, 779 Watkins, D., 765
Van Stratten, L.M.J.U., 246, 248, 779 Watts, N.L., 185, 782
VanWie, W., 767 Wear, C.M., 280, 776
Varfolomeeva, E.K., 775 Weaver, C.E., 1, 3, 8-10, 23, 30, 33, 39, 41,
Varsdnyi, I., 689, 779 43-45, 51, 58, 63, 65, 66, 71, 75, 76, 79-
Veith, J.A., 162, 779 81, 83, 84, 88, 91, 95, 98, 100, 102, 141,
Velde, B., 39, 133, 178, 392, 426, 439, 487, 149, 151, 152, 158, 159, 164, 165, 169,
508, 573, 576, 585, 750, 760, 765, 779 174, 179, 180, 182-184, 193, 194, 202,
Veniale, F., 94, 667, 781 233, 235-239, 241, 246, 252, 276, 302,
Venkatarathnam, K., 263, 781 312, 320, 342, 348, 366, 368, 380, 383-
Vergo, N., 634, 635, 781 386, 395, 400, 401, 405, 407-409, 414,
Verway, J., 225, 297, 781 417-421, 423, 424, 427, 431, 432, 437,
Viczidn, I., 641, 652, 666, 689, 781 439, 440, 442, 444-446, 449-459, 462-
Vidrich, V., 732 468, 470-476, 484-490, 492, 493, 495-
Vielvoye, L., 747, 760 499, 501, 504, 505, 507, 510-514, 525,
Viguier, C., 688, 755 535, 537-539, 561, 563, 565, 573, 575,
Vikulova, M.F., 622, 781 578, 580, 584, 600-602, 606, 617, 627,
Vinogradov, A.P., 583, 588, 629, 781 628, 659, 663, 677, 678, 691, 699, 701,
Vinopal, R.J., 764 706, 707, 715, 716, 718, 729, 740, 743,
Vitaliano, C.J., 589, 590, 737 745, 763, 765, 773, 782, 783
Vlasov, V.V., 775 Webb, L.C., 758
Vlodarskaya, V.R., 596, 629, 781 Weber, H.W., 118, 179, 779
Voigt, G.K., 729 Weber, K., 417, 452, 778, 783
Vollset, J., 752 Wedepohl, K.H., 512, 780
Von der Borch, C.C., 766 Weed, S.B., 140, 162, 756, 783
Von Herzen, R.P., 783 Weidmann, M., 755
von Rad, U., 384, 385, 727, 738 Weir, A.H., 53, 272, 666, 751, 783
Voznesenskays, T.A., 769 Weir, G.W., 6, 758
Weiss, A., 46, 48, 755
Waagk, K.M., 655, 781 Weiss, A.A., 83, 783
Wada, K., 25, 26, 143, 747, 781 Wells, J.T., 226, 230, 783
Wahlstrom, E.E., 647, 781 Welte, D., 736
Walker, G.F., 82-84, 758, 781 Welte, D.H., 519, 779
Walker, T.R., 157, 627, 781 West, G., 638, 737
Wall, G.J., 138, 781 Westcott, J.F., 752, 753
Wallace, J.M., 300, 781 Wheeler, W.H., 664, 747
Wallin, C., 729 Whetten, J.T., 754
Wampler, J.M., 158, 408, 427, 442, 496, Whitaker, J.H.MeD., 554, 728
765, 773, 782, 783 White, D.E., 445, 762
Wan, H.-M., 730 White, E.M., 115, 783
808

White, J.L., 758 Wollast, R., 123, 275, 380, 381, 784
White, S.H., 463, 478, 783 Wones, D.R., 734
White, W.A., 614, 615, 743, 745, 749 Woodward, K., 757
Whitechurch, H., 751 Woollen, I.D., 732
Whitehouse, U.G., 208, 211, 783 Wright, C.A., 782
Whiteman, J.A., 734 Wright, J.C., 647, 773
Whitlow, S.I., 591, 748 Wright, L.D., 248-250, 734, 784
Whitman, J.A., 750 Wyrwicki, R., 689, 785
Whitney, G., 724, 738
Whittards, W.F., 32, 725 Yaalon, D.H., 8, 9, 120, 160, 785
Whittig, L.D., 159, 166, 174, 757, 783 Yarilova, E.A., 138, 140, 177, 765
Whittington, H.B., 579, 783 Yariv, S., 27, 28, 785
Whittle, C.K., 734 Yart, J.W., 755
Whitton, J.S., 778 Yau, L.Y.C., 446, 785
Wiersma, J . , 101, 690, 783 Yeats, R.S., 773
Wiewicira, A., 21, 94, 689, 783, 785 Yeh, H.-W., 368, 371, 739, 758, 760
Wilde, S.A., 776 Yeh, H.W., 746, 783
Wilding, L.P., 107, 108, 781, 783 Yoder, H.S., 33, 37-39, 41, 775, 785
Wilkes, P., 743 Yokokura, H., 777
Willand, T.N., 764 Yong, R.N., 758
Williams, D.L., 352, 783 Yoshinaga, N., 26, 781
Williams, E.G., 614, 735, 783 Yotsumota, H., 781
Williams, G.D., 512, 731 Young, G.M., 125, 569, 762
Williams, P.E., 748 Young, R.A., 15, 777
Williams, P.F., 460, 463, 478, 784 Youngberg, C.T., 733, 778
Williams, V.E., 766 Yowell, R.F., 620, 768
Williamson, I.A., 624, 784
Williford Jr., C.W., 771 Zabawa, C.F., 220, 225, 785
Willis, J.P., 727 Zak, I., 747
Willman, H.B., 742 Zalan, P.V., 692, 693, 724
Wilson, J.T., 578, 784 Zalba, P.E., 623, 750
Wilson, M.D., 525-531, 533, 784 Zantop, H., 776
Wilson, M.J., 22, 79, 152, 160, 166, 594, Zemmels, I., 649, 696, 697, 758, 785
620, 758, 762, 784 Zen, E.-An., 334, 785
Wiltshire, J.C., 758 Zharkov, M.A., 407, 785
Winar, R.M., 600, 602, 760 Ziegler, A.M., 725, 765
Windley, B.F., 619, 625, 784 Ziegler, P.A., 650, 680, 785
Windom, H., 308, 323, 743, 784 Zinsmeister, W.J., 758
Windom, H.L., 153, 238, 246, 303, 305- Zitzmann, A., 396, 785
307, 310-313, 315, 316, 323, 328, 333, Zkhus, I.D., 622, 785
783, 784 Zuberec, J., 689, 754
Winkler, H.G.F., 417, 448, 483, 490, 784 Zussman, J., 735
Wintsch, R.P., 756 Zvjagin, B.B., 771
Wise, W.S., 784 Zvyagin, B.B., 13, 362, 785
Wolf, M., 520, 784
809

SUBJECT INDEX

Abyssal Plain, 285 - greenschist, 352, 363


Activity diagram, 381 - high-temperature, 352
- sandstones, 544 - high to low temperature, 363
Adriatic Sea, 262 - hydrothermally altered sediments, 371
Allegheny Orogeny, 602 - low-temperature, 354
Allophane - montmorillonite, 360
- electron micrograph, 25 - nontronite, 366
- X-ray patterns, 26 - palagonite, 356, 372
Amargosa Desert, 276 - palygorskite, 379, 400
Amesite, chemical analysis, 32 - saponite, 352, 363
Anchizone, 418, 452, 464 - sepiolite, 379, 400
Antarctica, 337, 339, 701 - smectite, 352, 354, 365
Appalachians, 204 - volcanic ash, 375
40Ar-39Ar, 491 Authigenic physils, 414, 525
Arabian Sea, 383 - criteria, 527
Arcadian Orogeny, 573, 575 - recognition, 527
Arctic fiords, 230 Bacteria, 395
Arctic Ocean, 703 - SEM, 228, 294
Argon-release patterns, 493 Banderas Bay, 349
Ash beds, 375 Basalt, 355, 373, 377
Astoria Canyon, 264 Basin, 410
Atlantic Ocean, 285, 303, 372, 703 Bauer Deep, 338
- chlorite, 316 Beidellite, 51
- illite, 321 - chemical analyses, 53
- kaolinite, 315 - origin, 53
- montmorillonite, 318 Bentonite, 375, 378, 667
- physils, 314, 695 Berthierine, 35, 387, 395
- pyrophyllite, 323 - chemical analysis, 32
- suspended material, 291 - distribution, 395
- suspended physils, 324 - electron micrographs, 482
- talc, 323 - occurrence, 388
Atmospheric transport, 300 Biotite, 465, 504
Australian desert, 343 Black shales, 595
Authigenic marine, 345, 351 Calcretes, 183
- ambient marine smectite, 372 Caledonian orogeny, 573, 588, 594
- beidellite, 360 Callianassa, 297
- berthierine, 387 Carbonate, 663
- celadonite, 352, 355, 361 Carboniferous paleogeography, 598
- celadonite-nontronite, 363 Caribbean Sea, 261, 322
- chlorite, 352 Cascadia Basin, 289
- chloritic, 363 CEC, 345
- Ch/S, 352 - river, 346
- discharge deposits, 365 - sea, 346
- Fe-rich montmorillonite, 366 Celadonite, 86, 361
- glauconite, 355 - chemical analyses, 360
- glaucony, 387 - chemical composition, 95
810

Celadonite (cont’d) - prophyroblasts, 483


- chemistry, 87 - sandstone, 533, 543
- origin, 93 - SEM, 463, 471, 532
- X-ray data, 90 - structure, 67
- X-ray patterns, 88, 356 - X-ray, 68
Celadonite/smectite, 358 - X-ray data, 70
Chamosite, 395 - X-ray patterns, 72
- distribution, 395
Chlorite classification, 487
Chemical analyses Chlorite/smectite, 78
- amesite, 32
Chlorite/vermiculite, 78
- beidellite, 53
Chloritic, 363
- celadonite, 95, 360
Chloritic physils, 77
- chlorite, 75, 80
- chemical composition, 80
- cronstedtite, 32
- magnesium hydroxide, 78
- Fe-smectite, 365
- origin, 80
- glaucrinite, 91
Ch/S, 407, 443, 578, 619
- greenalite, 32
- sandstone, 543
- halloysites, 30
- SEM, 413, 548
- hectorite, 55
- X-ray patterns, 412, 549, 635
- illite, 43
ChIV, 619
- I/S, 63
Classification, 1, 11
- kaolinite, 22
- kerolite, 97
Clay fabric, 710
- SEM, 711
- montmorillonite, 52, 173
- nontronite, 54 Cleavage lamellae, 472
- SEM, 473
- nontronite, marine, 369
- palygorskite, 101
Climate, 112, 576, 701, 702
- phengite, 45
chlorite, 186
- rectorite, 66
Ch/V, 120
- saponites, 56 clay content, 113
- sepiolite, 100 cretaceous, 188
- smectite, 360 desert, 284
- stevensite, 55 gibbsite, 119
- talc, 97 illite, 119, 188
- vermiculite, 85 I/S, 188
Chemical weathering, 104 I/V, 120
- experimental, 122 kaolinite, 119, 186
- inorganic, 122 montmorillonite, 118, 120
Chesapeake Bay, 230 palygorskite, 120
China Sea, 328 physil types, 116
Chlorite, 316, 330, 408 pleistocene, 186
- chemical composition, 75 smectite, 186
- chemistry, 72 temperate, 146, 156
- classification, 73 tropical, 142, 155, 284
- composition, 486 vermiculite, 120, 121
- electron micrographs, 436, 480, 482 Climatic zones, 284
- origin, 76 Coalification, 513
- pods SEM, 465, 468 Coals, 513
- polytype, 486 Color, 720
811

Compaction - montmorillonite composition, 429


- illite, 705 - New Zealand, 427
- kaolinite, 705 - R b S r , 427
- montmorillonite, 705 - Reykjanes, 445
Composition, 7 - Rhine Graben, 424, 447
Conasauga shale, 452 - Salton Sea, 446
Continental drift, 571 - San Joaquin Basin, 428
Continental shelf, 232, 258 - SEM,441
- transport, 279, 285 - smectite, 418
Corrensite, 78, 407, 443, 619 - smectite/illite, 418
- X-ray pattern, 81 - Sverdrup Basin, 440
Cretaceous paleogeography, 653 - TEM, 432, 447
Cronstedtite, chemical analysis, 32 - temperature, 420, 423, 432, 447
Crystallinity index, 21, 318 - vermiculitic, 439
Definitions, 6 - Wairakei, 445
Degraded illite, X-ray patterns, 640 Dickite
Delta, 248 - sandstone, 531
- Amazon, 258 - structure, 13
- Charleston, 235 Differential flocculation, 208, 213, 259
- Congo, 315 Dioctahedral vermiculite, 168
- glauconite, 257 Double layer, 206
- Mississippi, 251, 279, 285 Dust, 303, 323, 324, 341
- Niger, 255, 315 - glaciers, 305
- Nile, 263 - physils, 304, 307
- Orinoco, 279 - size, 305
- PO, 262 - snowfield, 305
- RhBne, 279 - talc, 308
- Santee, 235 Dusts, 319
Desert, 343 Electron micrograph
Devonian paleogeography, 593 - allophane, 25
Diagenesis, 417, 452, 461 - berthierine, 482
- Beaufort Sea, 440 - chlorite, 436, 480, 482
- berthierine, 35 - halloysite, 27
- chemistry, 440 - illite, 261, 436, 438, 480, 482
- chlorite, 35, 446, 484 - imogolite, 25
- chlorite porphyroblasts, 462 - I/S, 434
- Ch/S, 443 - kalonite, 150
- corrensite, 443 - montmorillonite, 47
- division, 418 - slate, 482
- Gulf Coast, 422, 428, 439 - smectite, 150, 433
- hydrothermal experiments, 430 Electron microscopy, 178
- illite, 445 Eolian, 327
- I/S/Ch, 421 - aerosols, 300
- K-Ar, 427, 442 - deposition, 301
- K-bentonites, 426, 442 - palygorskite, 183
- K-feldspar, 422, 424, 442 - Quartz, 303
- kaolinite, 446 - tranport, 300
- mancos shale, 426 Epizone, 418, 452, 469
- mechanism, 430 Erosion, 103
812

Estuaries Florida Keys, 327


~ Banderas Bay, 242 Galapagos, 338, 365
- biodeposits, 224 - temperature, 367
- Brunswick Harbor, 236 Galapagos Islands, 335
- Charleston Harbor, 238 Glaciers, 103, 244
- Chesapeake Bay, 225 Glauconite, 86, 363
- Delaware Bay, 225, 232 - C.E.C., 90
- deposition, 220 - chemical composition, 91
- fecal pellets, 224, 226 - composition, 391
- filter feeders, 224 - distribution, 395
~ fulvic, 213 - K-Ar, 397
- Glacier Bay, 244 - morphology, 389
- Howe Sound, 225 - origin, 93, 389
- humic, 213 - Rb-Sr age, 399
- James River, 222 - X-ray data, 90
- material, 215 - X-ray patterns, 87-89, 356
- organic, 215 Glaucony, 92, 387
- physils, 230 - K-Ar age, 397
- Port Royal Sound, 235 - morphology, 388
- San Francisco Bay, 242 - occurrence, 388
- Satilla River, 237 - origin, 389
- settling, 205 - SEM, 392, 394
- tidal currents, 221 - X-ray patterns, 390
- York River, 225 Great Smoky fault, 453
Evaporite physils Greenalite, chemical analysis, 32
- Mg-rich physils, 407 Greenschist, 352
- Palo Duro, 410 Gulf Coast, 510, 552, 707
- Paradox Basin, 411 Gulf of Mexico, 252, 261, 326
- Salado Formation, 410 H alloysi te
- Zechstein, 409 - chemical analyses, 30
Fan - electron micrograph, 27
- Amazon, 288 - SEM, 148
- Indus, 341 - structure, 26
- Mississippian, 286 - X-ray powder data, 28
- Nitinat, 289 Hectorite, 54
Fe, 475 - chemical analyses, 55
Fe-montmorillonite, X-ray patterns, 370 Hemipelagic clays, 707
Fecal pellets, 246, 293, 295, 393 Hercynian Orogeny, 602, 625
- SEM, 227 Hisingerite, 54
- X-ray patterns, 299 Hydrothermal, 380, 448
Feldspar, 540, 554 Hydroxy interlayers, 173
Fiord, 226 Iapetus Ocean, 578
Fissility, 6 Ice transport, 309
Flocculation, 206, 210, 212, 226 Illite, 32, 174
Flocs, 211, 213, 218, 222, 226, 705 - chemical analyses, 43
- organic, 216 - chemistry, 41
- SEM, 217, 218, 227, 708 - electron micrograph, 261
- size, 214 - electron micrographs, 436, 438, 480, 482
Florida Bay, 241 - fibrous, 535
813

Illite (con t’d) - diagenesis, 446


- lMd, 177 - electron micrograph, 150
- polymorphs, 37 - Fe2O3, 23
- sandstone, 533, 551 - sandstone, 531, 540
- SEM, 464, 470, 534, 535, 554, 559 - SEM, 148, 531, 559
- structure, 33 - structure, 13
- X-ray patterns, 42 - TiO2, 23
Illite/smectite (I/S), 418, 614 - X-ray patterns, 18, 20
- chemical composition, 63 - X-ray powder data, 16
- chemistry, 62 Kaolinite-illite
- electron micrographs, 434 - SEM, 555-557
- I/S ratio, 61 Kaolinite/smectite, 21
- K, 64 Kerolite
- ordering, 61 - chemical composition, 97
- origin of, 67 - chemistry, 93
- sandstone, 552 - origin, 94
- SEM, 441 - X-ray, 93
- X-ray, 57 - X-ray patterns, 94
- X-ray pattern, 321 K.I., 452, 454, 455, 505, 520, 522
- X-ray patterns, 60, 274, 318 Labrador, 204
Imogolite Lake
- electron micrograph, 25 - Albert, 272, 273
- X-ray pattern, 26 - Amboseli, 277
Indian Ocean - Bolivian, 275
- Fe-montmorillonite, 340 - Bow Lake, 268
- illite, 339 - Chad, 273
- kaolinite, 343 - clastic, 265
- montmorillonite, 341 - Constance, 267
- palygorskite, 342 - Crater, 268
- smectite, 342 - Erie, 267
I/S, see Illite/smectite, 614 - fecal pellets, 268
Jurassic paleogeography, 646 - hectorite, 277
K-Ar, 312, 491, 496, 584 - illite, 273
- Alpine, 502 - kerolite, 277
- carbonate residues, 498 - kerolite-stevensite, 276
- Conasauga, 499 - Kinneret, 268
- tertiary physilites, 499 - Manyara, 273
I<-Ar age, 427, 442 - Maracaibo, 268
- glauconite, 397 - Maurepas, 269
K-bentonite, 426, 439, 442, 575, 584 - Michigan, 267
K-smectites, 180 - montronite, 273
Kaolin, 13 - Mound, 276
- polymorphism, 15 - Nicaragua, 268
Kaolinite, 315 - Ontario, 267
- cation exchange capacity, 24 - palygorskite, 276
- chemical composition of, 22 - Pleistocene, 276
- chemistry, 22 - Pontchartrain, 269
- crystallinity index, 21 - Saline, 270
- degree of crystallinity, 17 - saponite, 276
814

Lake (con t 'd) - porphyroblasts, 483


- Searles, 271 - quartz, 510
- sepiolite, 275, 277 - sharpness ratio, 449
- Silver, 272 - SiOz/A1203, 511
- Snuggedy Swamp, 269 - slates, 478
- Stanwell-Fletcher, 267 - slaty cleavage, 460
- stevensite, 272, 275 - stability range, 507
- Superior, 267 - tectonics, 503
- Tahoe, 267, 294 - temperature, 489, 503
Lattice fringe image, 433, 481 - temperatures, 458
Layer charge, 64 - terminology, 418
Layer silicate films, 478 - tourmaline, 510
Lithification, 705 - vacuoles, 479
Loess, 329, 332 - Weaver Index, 449
Mancos shale, 426 - W.I., 469, 505
Marine deposition, 279 Mg-physil, 565
Marine transport, 279 Mica, 174
Marshes, deposition, 244 - SEM, 467, 577
Mediterranean, 263 Microorganisms, 294
- organic, 513 Mississippi Delta mud, 511
Metamorphism, 417, 452 Mobile Bay, 255
- (002)/(001), 458 Montmorillonite, 46, 51, 318, 429
- anchizone, 464 - chemical analyses, 52, 429
- 40Ar-39Ar, 492 - electron micrograph, 47
- biotite, 465, 471, 503 - layer charge, 46
- carbonate rocks, 457 - SEM, 550
- chlorite, 469, 483, 484 - X-ray pattern, 337, 359
- chlorite composition, 486 Muds, 715
- chlorite porphyroblasts, 462 Mudstone, 481
- cleavage lamellae, 469, 472 Muscovite, X-ray d a t a for, 40
- Conasauga shale, 452 Nacrite, structure, 13
- crystallinity index, 449 Namib Desert, 316
- epizone, 469 Neogene paleogeography, 672
- greenschist, 475 Nepheloid layer, 288
- illite-mica, 454 Nomenclature, 5
- K-Ar, 475, 491, 496 Nontronite, 51, 393
- - data, 471 - chemical analyses, 54, 369
- K.I., 469, 505 - origin, 53
- kinking, 474 Oceans, 309
- KzO/A1203, 511 - chlorite, 311
- Kubler Index, 449 - illite, 312
- layer silicate films, 478 - kaolinite, 310
- 2M, 460 - montmorillonite, 311
- mica, 475 - physils, 310
- 018/016, 458 - quartz, 313
- oxygen isotope, 489 - suspended material, 283
- petrology, 460 Octahedral charge, 64
- phengite, 465, 469 Odinite, 397
- planar foliation, 474 Ordovician paleogeography, 579
815

Organic Material, 517 - Conasauga, 563


Orientation, 461, 707, 710 - Cretaceous, 653
Oxidation, 134 - Devonian, 589
Oxygen isotopes, 489 - early paleozoic, 571
Pacific Ocean, 285, 703 - England, 636
- chlorite, 330 - Europe, 571, 577, 582, 587, 594, 619,
- eolian, 327, 333 629, 634, 649, 665, 679
- Fe-montmorilline, 338 - France, 638
- Fe-montmorillonite, 336 - Germany, 638
- illite, 327, 332 - Gulf Coast, 676
- I/S, 329 - Israel, 644
- montmorillonite, 332 - Jurassic, 645
- quartz, 331 - Mesozoic, 696
- smectite, 333 - Mississippian, 598
Palagonite, 356 - North America, 568, 576, 578, 585, 589,
Pdeogene, 672 598, 632, 645, 655, 670
Paleosoils, 185 - Ordovician, 578
Paleothermometers - Paleoatlantic physils, 693
- organic, 513, 520 - Pennysylvanian, 602
- physil, 513, 520 - Permian, 625
Palygorskite (=Attapulgite), 94 - Pleistocene, 693
Palygorskites, 342, 379, 400 - Precambrian, 568
- chemical composition, 101 - Silurian, 585
- chemistry, 98 - Sinai, 644
- occurrences, 382 - South America, 594, 619, 623, 629, 644,
- origin, 102 669, 692
- recent, 406 - Spain, 641
- SEM, 184, 404 - Switzerland, 641
- stability relations, 405 - Triassic, 630
- structure, 94 - X-ray patterns, 606
- X-ray data, 99 Physils, 1
Paragonite, 454 Pierre Shale, 660
Paris Basin, 683 Planar foliation, SEM, 474, 475
Pellets, 528 Pleistocene, 287
Permian paleogeography, 626 Porphyroblasts, 506
Petrology, 705 Precipitation, mean global, 566
Phengite, 37, 44, 465, 469, 476, 504 Proto-Atlantic, 578
- chemical composition, 45 Proto-Atlantic Ocean, 573
Physical weathering, 103 Protosoils, 188
Physil dissolution, 349 Pysilites
Physilites, 7 Quartz, 558, 715, 717
- Africa, 577, 582, 587, 594, 619, 623, 629, Rb-Sr age, 427
649, 668, 690 - glauconite, 399
- Atlantic Ocean, 696, 699 Rectorite
- Cambrian, 576 - composition, 66
- Carboniferous, 598 Red Sea, 365
- Cenozoic, 670, 699 Reichwite numbers, 59, 420
- Coastal Plains, 661 Rhine, 684
- composition, 714 Rhine Graben, 424
816

Rh6ne Graben, 684 Saharan air layer, 303


Rise, 285 Salton Sea, 446
River San Antonio Bay, 242
- Altamaha, 202, 236 Sandstone
- Amazon, 189, 192, 195, 203, 2159, 286, - activity diagrams, 544
314, 322, 325 - authigenic physils, 525
- Apalachicola, 253 - chlorite, 533, 543, 560
- Arkansas, 193 - Ch/S, 543, 560
- Ashley, 238 - C 0 2 , 536, 540
- Brahmaputra, 341 - diagenetic reactions, 539
- Chari, 273 - dickite, 531
- clay suite, 193 - illite, 533, 542, 551, 560
- Columbia, 204, 264 - illite/smectite, 536
- Congo, 286, 315 - I/S, 542, 552, 560
- Cooper, 238 - kaolinite, 531, 540, 560
- Delaware, 232 - kaolinite-illite, 554
- discharge, 190 - meteoric water, 536
- dissolved load, 191 - montmorillonite, 560
- Ganges, 189, 286, 341 - mud water, 539
- Indus, 286, 341 - organic acids, 536
- Jordan, 268 - oxygen isotope, 543
- Mackay, 236 - pore water, 536
- major rivers, 205 - smectite, 536, 551
- Mississippi, 196, 253, 286 - true pH, 537
- Mobile, 349 Santa Barbara Basin, 265
- Neuse, 233 Saponite, 54, 409
- Niger, 315, 319 - chemical analyses, 56, 160
- Ogeechee, 247 - origin, 56
- Pamlico, 233 Saprolite, 149
- Pascagoula, 349 Sargasso Sea, 293
- Pee Dee, 233, 246 Sea level, 567
- PO, 262 Sedimentary structures, 712
- Potomac, 189 Sedimentation rates, 279
- Rappahannoch, 233 SEM, 178
- Rio Ameca, 242 - bacteria, 228, 294
- Sacremento-San Joaquin, 242 - bent mica, 467
- San Francisco, 196 - chlorite, 471, 532
- Santa Clara, 265 - chlorite pod, 463, 465, 468
- Santee, 202, 233, 238 - Ch/S, 413, 548
- Sao Francisco, 314 - clay fabric, 711
- Satilla, 238 - cleavage lamellae, 472, 473
- Savannah, 202, 247 - fecal pellet, 227
- Snake, 265 - fibrils, 294
- St. Lawrence, 204 - floes, 216-218, 227, 228, 708
- suspended sediment, 191 - glauconite-smectite, 394
- Wando, 238 - glaucony, 392
- Yellow, 189, 279, 285 - halloysite, 148
Rotation, 15 - hemipelagic clays, 709
Sahara Desert, 153, 315 - illite, 470, 534, 535, 554, 559
817

SEM (cont’d) - chemical analyses, 155, 158, 360


- illite folds, 464 - chemistry, 50
- I/S, 441 - cheto, 51
- kaolinite, 148, 531, 559 - electron micrographs, 433
- kaolinit e-illit e, 555-557 - sandstone, 551
- matting, 229 - SEM, 367
- mica, 577 - Wyoming, 51
- montmorillonite, 550 - X-ray, 49
- mucous, 229 - X-ray patterns, 49, 358, 376, 378
- palygorskite, 184, 404 Soil, 104
- planar foliation, 474, 475 - basalt, 143, 195
- serpentine, 413 - bauxites, 142
- shale, 461 - classification, 107
- slate, 466 - dominant clay mineral, 109
- slate, incipient, 462 - gibbsite, 195
- slaty cleavage, 468 - halloysite, 195
- smectite granule, 367 - hydroxy interlayers, 170
- talc, 413 - illite, 174, 186
- turbiditic clays, 709 - kaolin, 142
Sepiolite, 379, 400 - kaolinite, 186
- chemical composition, 100 - laterites, 142
- chemistry, 98 - montmorillonite, 153, 195
- occurrences, 382 - palygorskite, 182
- origin, 99 - pyrophyllite, 185
- structure, 94 - saprolitic, 201
- X-ray data, 99 - sepiolite, 182
Serpentine, 407 - smectite, 153, 186
- chemistry, 31 - talc, 185
- SEM, 413 - vermiculite, 161
- structure, 30 Soil formation, 103
Settling velocities, 208 Soil profile, 105
Shale colors, 720 S o hbility
Shales, 424, 517 - Al, 125
- composition, 561 - Si, 125
- Conasauga, 563 S.R., 521
- diagenetic changes, 563 Stability diagrams, 128
- K-Ar, 496 Stability range, 507
- mineral distribution, 561 Stevensite, 54
- SEM, 461 - chemical analyses, 55
- temporal changes, 563 Structural formula, 3
Sharpness ratio, 450 Submarine canyons, 286
Slate, 478 Suspended material, 291
- electron micrographs, 482 Suspended physils, 299, 324
- SEM, 466 Suspended sediment, 280, 293
Slaty cleavage, 469 Synt hesis
- SEM, 468 - chlorite, 126
Slope, 285, 327 - illite, 126
Smectite, 358 - smectite, 126
- CEC, 46 Taconic Orogeny, 573, 575, 581, 601
818

Talc, 308, 323, 324, 407 - gibbsite, 143, 149


- chemical composition, 97 - halloysite, 131, 148
- chemistry, 93 - hematite, 143
- origin, 94 - hydroxy interlayers, 170
- SEM, 413 - illite, 133, 159, 174, 179, 186
- X-ray, 93 - kaolin, 142
- X-ray patterns, 94 - kaolinite, 131, 143, 146, 149, 186
Temperature, 489, 516 - kaolinite-montmorillonite, 144
- coal, 516 - kaolinite-smectite, 151
- LOM, 518 - laterites, 142
- mean global, 566 - micas, 131
Tethys, 406, 653, 703 - microbial, 137
Tetrahedra, 1 - montmorillonite, 153, 157
- rotation of, 36 - muscovite, 133, 159, 162
Tetrahedral, 15 - nontronite, 156
Tetrahedral charge, 64 - organic, 134
Thin sections, 721 - orthopyroxene, 131
Ti, 475 - palygorskite, 182
Tidal flats, 244 - saponite, 160
TiO2, 374 - sepiolite, 182
Tonsteins, 624 - smectite, 131, 144, 153, 186
Tosudite, 78 - talc, 131, 185
Triassic paleogeograpy, 631 . - TiO2, 143
Trioctahedral vermiculite, 166 - vermiculite, 159, 161, 169, 179
Turbiditic, 707 - volcanic ash, 143, 145
Underclays, 610 Weathering index, 125
Vermiculite White mica, 465
- chemical analyses, 168 W.I., 452, 456, 505, 522
- chemistry, 83 X-ray pattern
- origin, 86 - allophane, 26
- soil, 85 - black shales, 595
- X-ray, 82 - celadonite, 88, 356
- X-ray patterns, 171 - chlorite, 72
Vertisols, 155 - Ch/S, 412, 549, 635
Vitrinite reflectance, 517, 519, 524 - corrensite, 81
Volcanic ash, 375 - degraded illite, 640
Volcanogenic sedimentation, 308 - Fe-montmorillonite, 370
Washakie basin, 659 - fecal pellets, 296, 299
Weathering, 112, 165 - glauconite, 89, 356
- allophane, 143 - glaucony, 390
- amorphous material, 131, 143 - glaudonite, 88
- basalt, 143, 145 - Gulf of Alaska, 331
- bauxites, 142 - Gulf of Mexico, 260
- biotite, 133, 138, 156, 162, 166, 179 - illite, 42
- chlorite, 156, 159, 165, 166, 186 - imogolite, 26
- Ch/S, 165, 169 - I/S, 42, 60, 274, 318, 321
- Ch/V, 165, 169 - kaolinite, 18, 20
- Fe-smectite, 185 - kaolinite, montmorillonite, 144, 150
- feldspars, 130, 148 - kerolite, 94
819

X-ray pattern (con t’d) - soil, 177


- K.I., 454 - soil smectites, 378
- marsh, 321 - soil vermiculite, 171

- Mg hydroxide montmorillonite, 78 - suspended physils, 325

- montmorillonite, 337 - talc, 94


- muscovite, 169 X-ray radiographs, 715
- North Pacific, 329 Zechstein, 553, 629
- physilites, 606 Zeta Potential, 207
- smectite, 358, 359, 376
This Page Intentionally Left Blank

You might also like