You are on page 1of 227

Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.

org/
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Seismic Petrophysics in Quantitative Interpretation

Investigations in Geophysics Series No. 18

Lev Vernik

Rebecca Latimer, managing editor


Tad Smith, volume editor

SM

00-Vernik_FM.indd 1 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ISBN 978-0-931830-46-4 (Series)


ISBN 978-1-56080-324-9 (Volume)

Library of Congress Control Number: 2016945722

Copyright 2016
Society of Exploration Geophysicists
8801 S. Yale, Ste. 500
Tulsa, OK U.S.A. 74137-3575

All rights reserved. No part of this book may be reproduced, stored in a retrieval system, or transcribed in any form or by any
means, electronic or mechanical, including photocopying and recording, without prior written permission of the publisher.

Published 2016
Printed in the United States of America

00-Vernik_FM.indd 2 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Contents

About the Author . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .  ix

Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . xi

Chapter 1: Petrophysics of Siliciclastic Rocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1


Petrophysical classification of siliciclastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Petrographic data and petrophysical classification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1
Application to the core/log database . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
Petrophysical model building . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Lithologic parameters Vcl and Vsh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
Total porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11
Permeability prediction in siliciclastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
Seismic petrophysics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
Log editing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Sonic log . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Density log . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
Anisotropic correction of sonic logs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
Chapter 2: Pore Pressure and Stress State . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Shale-compaction model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Porosity reduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29
Density model for stress computation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
Vertical effective stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
Pore-pressure prediction from shale velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Sonic velocity versus effective stress in shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
Pore-pressure prediction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36
The effective-stress tensor and the Shmin gradient . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
Chapter 3: Seismic Rock Properties and Rock Physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Theoretical models in rock physics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Fluid-saturation effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43
Drained rock-frame moduli . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45
Cracks in dry and fluid-saturated rocks: The effect of aspect ratios . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46
Drained rock-frame moduli: A mixture of cracks and pores . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
Effects of pore/crack interactions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
Contact models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

iii

00-Vernik_FM.indd 3 12-08-2016 20:04:40


iv Seismic Petrophysics in Quantitative Interpretation

Drained frame moduli and effective stress . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52


Rock-physics modeling in sand/shale sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Pore shapes and porosity thresholds in sands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
Consolidated sandstones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Unconsolidated and poorly consolidated sands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61


Sandstone diagenesis models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
Conventional shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
Rock-physics templates in siliciclastic sequences . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69
VP-VS and AI-SI relationships in siliciclastics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
The VP-VS relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72
The AI-SI relationship . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 76
Fluid substitution and the AI-SI template . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
4D modeling of sands and sandstones . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80
Chapter 4: AVO Analysis: Rock-physics Basis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
Linearized AVO equations and their features . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
AVO classification in AI-SI space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89
From the AI-SI template to synthetic-gather models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
AVO Class III and the zero-gradient case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91
AVO Class IV . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
AVO Classes I and II . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
VTI anisotropy effect . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
The fluid factor in prospect risk mitigation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
Tuning effects in AVO synthetic modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101
Chapter 5: Simultaneous AI-SI Inversion and N/G Computation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Log editing and the AI-SI template . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
Anisotropy correction of sonic velocities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106
Additional log editing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
AI-SI crossplot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 107
Modeling N/G . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Seismic N/G computation from simultaneous inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
AI-SI inversion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111
Sand volume computation and net/gross mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 114
Effects of lithology and fluid on prestack attributes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118
Chapter 6: Seismic Petrophysics of Unconventional Reservoirs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Petrologic data in unconventional shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Rock composition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121
Organic richness and thermal maturity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 122
Rock texture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126
Log model for unconventional shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Microstructural observations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129
Kerogen-fraction log . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Total porosity and kerogen porosity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131
Water saturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133
Model applications . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 134
Rock physics of unconventional shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Core measurements of velocity and anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 137
Phase velocity versus group velocity . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 139
Intrinsic velocity and anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 141
Effect of kerogen on velocities and anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 144

00-Vernik_FM.indd 4 12-08-2016 20:04:40


Contents v

Stress dependence of velocity and anisotropy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 153


Modeling stress dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 155
Rock-physics model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 158
VP-VS and AI-SI relationships . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 161
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 7: Geomechanics of Organic Shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167


Elastic-property relationships in TI shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167
Can brittleness be estimated from elastic properties? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 168
Static and dynamic elastic parameters in anisotropic mudrocks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 169
Elasticity-based brittleness of organic mudrocks: A controversial notion . . . . . . . . . . . . . . . . . . . . . . . . . . . 171
Shmin stress-gradient estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Uniaxial-strain-based approach . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173
Anisotropy prediction in organic and conventional shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 174
Stress profiling from log data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175
Geomechanics of maturation-induced microcracking . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 176
Chapter 8: Seismic Analysis in Unconventional Shales . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
Reservoir-scale anisotropy estimation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183
Seismic ties . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184
TOC estimation from an AI-SI template . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186
AI-SI Inversion of Prestack Seismic Data . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
TOC mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 191
Shmin stress-gradient mapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 194
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 199

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 209

00-Vernik_FM.indd 5 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

About the Author

Lev Vernik is a geophysics consultant and owner of Seismic


Petrophysics, LLC, in Houston, Texas, USA, and a research profes-
sor of geophysics at the University of Houston. He has previously
held various geoscience positions with Arco, Vastar Resources, BP,
Noble Energy, and Marathon Oil, focusing on seismic petrophysics,
AVO modeling/analysis, prestack seismic inversion, and geome-
chanics. Lev’s long career in subsurface characterization started in
the former Soviet Union, where he was involved in drilling and in-
vestigating the world’s deepest well on Kola Peninsula. His research
continued at the Stanford Rock Physics Project and then in the tech-
nology groups of the aforementioned five oil and gas companies.

vii

00-Vernik_FM.indd 7 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Preface

Recent advances in seismic acquisition and process- discussed in Chapter 3, “Seismic Rock Properties and
ing technologies have led to enormous progress in oil and Rock Physics.” The chapter also addresses certain contro-
gas exploration and reservoir characterization. Quanti­ versial issues, such as parameterization and appropriate
tative seismic interpretation also could have improved model selection. The outcomes of this discussion are sev-
significantly if we had had the internally consistent meth- eral rock-physics templates that feature the key petrophys-
odologies and workflows necessary for seismic modeling. ical groups of siliciclastics in velocity-versus-porosity
Seismic petrophysics plays the key role in this respect. space and acoustic-impedance-versus-shear-impedance
Because the methodologies of petrophysical analysis (AI-SI) space. Those templates naturally incorporate the
lead to uncertainties in estimating key parameters, such as effect of rock compaction and diagenesis and therefore are
rock composition, texture, porosity, permeability, and essential in modern seismic inversion interpretation.
saturation, it is important to (1) understand the rock for- Chapter 4, “AVO Analysis: Rock-physics Basis,” is
mations for their proper characterization and (2) incorpo- the core of the book. This chapter integrates seismic pet-
rate proper rock-physics and geomechanics concepts in rophysics with seismic amplitude analysis — notably,
predicting the seismic signatures of those formations. with amplitude variation with offset (AVO) or angle of
Over the last several years, our understanding of incidence (AVA). The AVO classification scheme and in-
­seismic rock properties and our ability to model those terpretation methodology offered here differ from the
properties — VP, VS, and density — has improved substan- classic scheme of intercept/gradient analysis and invoke
tially. Nonetheless, a certain amount of confusion remains acoustic- and shear-impedance templates instead of inter-
with respect to the petroelastic classification of rocks, the cept/gradient crossplots. That allows a user to streamline
terminology used, the selection of models, the parameter- AVO behavior predictions during seismic petrophysics
ization, and the identification of the primary controls on analysis of the log data early in the workflow. It also pro-
the distribution of rock properties in the subsurface. These vides a robust quality assurance during synthetic forward
issues are discussed in Chapters 1 and 3 of this book. modeling. In turn, that modeling can be used in seismic
Pore-pressure prediction from well-log and seismic processing and gather conditioning for prestack attribute
data is critical in drilling operations and also is important computation and simultaneous impedance inversion of
in estimations of geomechanical parameters and stress- seismic data.
states from seismic data. It is well established that s­ tresses The AI-SI templates and their benefits in prestack in-
in the subsurface have a direct impact on rock physics version of sand/shale sequences are exemplified in
and, hence, on seismic amplitudes. Indeed, knowledge of Chapter 5. Marked mismatches in frequency content be-
effective-stress variations laterally and with depth mark- tween seismic data and log data result in large differences
edly reduces interpretation uncertainty and helps separate in vertical resolution and lead to the argument that limita-
intrinsic effects from extrinsic influences of pore pressure tions exist in our ability to directly invert for certain res-
and stress state on seismic rock properties. An approach ervoir properties (e.g., porosity and permeability) in high-
to pore-pressure and vertical-effective-stress analyses is ly heterogeneous sand and sandstone reservoirs. However,
discussed in Chapter 2. another valuable reservoir parameter — the net-to-gross
The essence of seismic-related rock physics consists thickness ratio — lends itself to fairly accurate estima-
of finding relationships between seismic rock properties tions from seismic inversion in many cases.
and their petrographic and petrophysical parameters, in- Over the last several years, both the exploration and the
cluding their mineralogy, texture, diagenesis, reservoir development of unconventional shales have grown
properties, and effective stress. These problems are ­dramatically and, in fact, many operators have switched to

ix

00-Vernik_FM.indd 9 12-08-2016 20:04:40


x Seismic Petrophysics in Quantitative Interpretation

unconventional plays. It is often thought that these plays Palciauskas, 1994), Seismic Data Analysis (Yilmaz,
also require unconventional approaches to reservoir charac- 2001), Quantitative Seismic Interpretation (Avseth et al.,
terization. In Chapter 6, “Seismic Petrophysics of 2005), and The Rock Physics Handbook (Mavko et al.,
Unconventional Reservoirs,” it is demonstrated that the fun- 2009). Many theoretical models presented here originated
damental workflows typical for conventional reservoirs still from the work of Mark Kachanov (Tufts University) with
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

are applicable in organic shales. However, the complexity various coauthors.


of rock-physics modeling increases in proportion to the ad- Methodology and workflows described in this book
ditional parameters of interest, such as total organic carbon have been successfully applied in many exploration and
(TOC) content and thermal maturity, and their effects on reservoir characterization projects worldwide and result-
organic porosity, overpressure, and the stress sensitivity of ed in significant oil and gas discoveries and accurate re-
seismic velocities. Important geomechanical issues related serve calculations, particularly in the Gulf of Mexico,
to predicting horizontal-well-­completion quality — issues West Africa, North Sea, Norwegian Sea, and Eastern
such as brittleness, least-horizontal-stress gradient, and Mediterranean provinces.
hydrocarbon-­generation-induced microcrack development, Seismic Petrophysics in Quantitative Interpretation is
in organic mudstones — are discussed in Chapter 7. written for oil and gas industry professionals and academi-
Finally, Chapter 8, “Seismic Analysis in Uncon­ cians who are concerned with the use of seismic data in
ventional Shales,” presents examples of seismic forward petroleum exploration and production. In addition, it
modeling and prestack inversion for some of the most le- should prove useful toward thoughtful applications of those
veraging parameters of unconventional shale reservoirs data by geotechnical engineers. Seismic interpretation can
— both the intrinsic parameters (e.g., TOC content) and be made simple and robust by integrating rock-physics
the extrinsic ones (e.g., horizontal stress gradient). It ap- principles with the seismic and petrophysical attributes that
pears that AI-SI templates modified for organic shales can bear on the properties of conventional reservoirs (thick-
be adapted for those evaluations, and the ­organic richness ness, net/gross, lithology, porosity, permeability, and satu-
and fracability of organic mudrocks can be mapped on the ration) and of unconventional reservoirs (thickness, lithol-
basis of simultaneous seismic inversion results. ogy, organic richness, and thermal maturity). Practical so-
This book is, of course, complementary to other pub- lutions can be used to address existing interpre-
lications on related topics, such as Acoustics of Porous tation problems in rock-physics-based AVO analysis and
Media (Bourbie et al., 1987), AVO (Chopra and Castagna, prestack seismic inversion in order to streamline the
2014), Introduction to the Physics of Rocks (Gueguén and workflows in subsurface characterization.

00-Vernik_FM.indd 10 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Acknowledgments

I wish to thank Marathon Oil Corporation for permis- have helped to focus on what is important. Special thanks
sion to publish this book. Special thanks are extended to go to Amos Nur (Stanford University) for the jumpstart he
David Brimberry, Dicman Alfred, Jim Allen, Robert provided me in my early years in the United States. I am
Blanchard, Vladimir Grechka, Glory Kamath, Yulia also enormously grateful to Rebecca Latimer, Tad Smith,
Khadeeva, Scott Koza, Jadranka Milovac, and Chris Konstantin Azbel, and Marco Perez for their thoughtful
Tuttle (all presently or formerly with Marathon Oil), reviews and many helpful suggestions.
Mark Boggards and Niven Shumaker (Noble Energy), Special thanks also go to the SEG publication group,
and Tim Lane and Leon Thomsen (ex-BP). Cooperation led by Susan Stamm, for their diligence in the book’s
with Mark Kachanov (Tufts University) over the last five early review and later composition and printing phases.
years has been very useful in bringing me fresh perspec- Anne Thomas’s help with style editing and making the
tives on theoretical elasticity, whereas discussions with book consistent with the SEG standards is most
Brian Russell, Tad Smith, Kitty Milliken, and Öz Yilmaz appreciated.

xi

00-Vernik_FM.indd 11 12-08-2016 20:04:40


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 1: Petrophysics of Siliciclastic Rocks

This chapter was inspired by the book Quantitative and porosity, and (3) leads to development of theoretical
Seismic Interpretation, by Avseth et al. (2005), in which models that are not applicable to the specific petrophysical
we read that “some of the most important breakthroughs groups of rocks they are designed to simulate.
in rock physics during the past decade have come not Meanwhile, these compositional and textural parame-
from additional mathematics, but from rediscovering the ters are most closely linked to the provenance, original
physics of rock geology” (Avseth et al., 2005, p. xii). I depositional environment, and diagenesis of clastic sedi-
fully agree with that statement — many theoretical mod- mentary rocks (Blatt, 1982; Greensmith, 1989) and are of
els available to us today tend to oversimplify the petro- great and growing importance in oil exploration and reser-
graphic features of the rocks and, as a result, those models voir characterization (Avseth et al., 2005).
are difficult to properly test with experimental data and
downhole measurements.
Admittedly, the science of petrography can be tedious
Petrographic data and petrophysical
in dealing with some details of rock composition and tex- classification
ture that may not be critical for our goal of establishing the To gain better physical insight into the relations
major components that control elastic properties. Therefore, among elastic properties, porosity, and permeability in
the role of petrophysics, and especially that of seismic pet- such complex heterogeneous systems as sedimentary
rophysics, is hard to overestimate in this regard. rocks, the rocks must first be classified according to prin-
Using a diverse database of published measurements ciples that tend to consider petrophysical, compositional,
of elastic-wave velocities under controlled or well-­ and textural parameters simultaneously. In the domain of
established fluid-saturation and pressure/temperature con- clastic sediments and rocks, such principles are largely
ditions complemented by the porosity, mineralogy, and reduced to grain mineralogy, the amount and textural po-
texture of siliciclastic rocks, a straightforward classifica- sition of clay, and finally, the load-bearing structure (clay
tion scheme can be advanced and used as a foundation in matrix versus granular framework) of the rock. Such a
rock-physics model building and quantitative seismic classification helps to guide our understanding of inter-
analysis. relationships among seismic and petrophysical properties.
The first attempts to introduce a petrophysical classi-
fication of siliciclastics (Vernik and Nur, 1992b; Vernik,
Petrophysical classification of 1998) aimed at porosity prediction from sonic velocity
siliciclastics using empirical transforms (e.g., Wyllie et al., 1956;
Raymer et al., 1980; Han et al., 1986) in combination with
Numerous experimental data suggest that relatively Nur’s (1992) critical-porosity concept. The petrophysical
simple relations exist between seismic velocities and such classification proposed was deemed useful for predicting
important rock parameters as porosity and clay content lithologic facies from sonic and seismic data analyses, and
(Wyllie et al., 1956, 1958; Raymer et al., 1980; Tosaya and it included a most desirable link to permeability.
Nur, 1982; Kowallis et al., 1984; Han et al., 1986). A major For this book, a database was compiled with more than
shortcoming of these relations, however, is that they ne- 160 measurements of elastic properties, porosity, and pe-
glect lithologic parameters such as quantitative mineralo- trographic parameters from outcrop samples, cores, and
gy, the textural position of clay, and the distributions of core/log combinations. The majority of those data has been
pore size and shape. Neglecting these parameters (1) re- published (Han et al., 1986; Klimentos and McCann, 1990;
sults in reduced accuracy of the porosity prediction, (2) Domenico, 1977; Strandenes and Blangy, 1991; Issler,
complicates lithology prediction from acoustic velocity 1992; Vernik and Kachanov, 2010). Using VP-versus-porosity

01-Vernik_Ch01.indd 1 12-08-2016 21:03:46


2 Seismic Petrophysics in Quantitative Interpretation

Quartz
Clean
a sandstone is ν cl = ( IGV − φ ) / (1 − φ ) = 0.13. This is
2% arenite considered in more detail below.
Figure 2 shows thin-section photomicrographs of two
sandstones that have approximately the same grain size
and sorting of their clay-free components (predominantly
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Arenite
quartz and feldspars). The sandstone in Figure 2a, with a
net clay volume fraction vcl of 5%, falls into the domain of
arenites in the classification presented; the load in this
rock is supported by the framework grains. The texture of
Wacke the second sandstone (Figure 2b), which falls into the do-
main of wackes because it has a vcl of 15%, is largely clay-
Shale matrix supported and has only sporadic direct-sand-grain
contacts.
Other Clay
Conventionally, both sandstones in Figure 2 would be
12% 30% described as clay-bearing sandstones; however, that
Figure 1.  Petrophysical classification of siliciclastic rocks ­description is not useful in terms of elastic or reservoir
(Vernik and Nur, 1992b). properties. It would be even more confusing to refer to
arenites as “clean sands,” because their mineral composi-
tion includes chemically unstable feldspars and lithic
crossplots in combination with petrographic observations fragments with clay alteration (the so-called structural, or
of rock texture and structural position of clay aggregates, intragranular, clays).
the classification scheme displayed in Figure 1 can be pre- Incidentally, sands and sandstones that contain more
sented (Vernik and Nur, 1992b). than 12% of predominantly structural, load-bearing clay,
Note that even though this scheme uses some of the as is the case for some lithic arenites with abundant shale
major principles adopted in the purely petrographic clas- clasts, do not necessarily fall into the petroelastic catego-
sifications (e.g., Greensmith, 1989), it is distinctly differ- ry of arenites in the classification scheme discussed here.
ent from those schemes in that it focuses on the major Such lithologies are often characterized by mechanical
petrographic variables that affect the elasticity of silici- compaction related to ductile lithic or totally altered feld-
clastic rocks — the amount and structural position of spar clasts being gradually squeezed into the pore space
clay. Independently of depositional facies (deepwater tur- (Füchtbauer, 1967; Blatt, 1982). Ultimately, that process
bidites, and shallow-marine, fluvial-deltaic, or eolean leads to a rock structure similar to the one illustrated in
systems), these two parameters define the texture of si- Figure 2b, although it was quite different at the initial
liciclastics worldwide in a fundamental way: Sands and stages of sedimentation and compaction because of the
sandstones with a net volume fraction of clay (on a solid- minor presence of syndepositionally dispersed clay.
rock basis) that is less than 12% usually are character- In the domain of grain-supported sands and sand-
ized by the grain-supported texture, whereas those with stones, it is useful to distinguish the group of clean aren-
greater amounts of clay quickly transition into the domain ites, which have vcl values smaller than 2% and are almost
of predominantly clay-matrix-supported rocks. exclusively quartz sands and sandstones typical of well-
Just as is the case throughout the earth sciences, the sorted and mature deposits. The reason for separating
boundary of 12% is neither exact nor unique, and it could clean arenites from arenites (which have vcl values from 2
easily be replaced with a transition zone from 10 to 15% to 12%) is based on petroelastic properties discussed later.
or with a value as low as 8% (Kenter et al., 2007). On the other hand, rocks that are supported by a clay
Interestingly, the intergranular pore volume, IGV, in matrix can be subdivided into wackes and shales, with a
many slightly consolidated, clay-rich, moderately sorted tentative boundary at a vcl value of 30 to 35% (Vernik,
sands and sandstones with mostly rigid grains and minor 1998). This boundary is primarily compositional, but it
quartz cementation, typically averages 20% ± 4% (e.g., may also have significant petroelastic implications due to
Blatt, 1982) rather than 35 to 40%, as is suggested by the gradual development, in the clay-matrix-supported
physical modeling studies on artificial sand/clay mixtures rocks, of thinly laminated textures and preferred-
in the laboratory (Marion, 1990). An IGV of 20% can be clay-particle/aggregate alignments that result in substan-
completely filled with clay aggregate that has its own mi- tially greater elastic anisotropy in shales than in wackes.
croporosity of approximately 40%, which yields the total Another important petrographic parameter of siliciclas-
porosity of 8%. It is straightforward to show that, on a tic rocks is sorting (Avseth et al., 2005). Of course, if one is
solid-rock basis, the net-clay-volume fraction vcl that is to include depositional clay particles in the grain-size distri-
required to completely fill the intergranular space of such bution, then wackes would be considered to be more poorly

01-Vernik_Ch01.indd 2 12-08-2016 21:03:46


Chapter 1: Petrophysics of Siliciclastic Rocks 3

a) b)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

100 μm 100 μm

Figure 2.  (a) Medium-porosity, fine-grained feldspathic arenite with a vcl of 5%, and (b) medium-porosity, fine-grained shaly
sandstone (wacke) with a vcl of 15%. Notice that grain contacts are statistically insignificant in the wacke because of sand-grain
insulation, even at this relatively small clay content. Sorting in these two sands on a clay-free basis is approximately the same.
The permeability for the sample on the left is approximately two orders of magnitude greater than that for the sample on the right.

a) whereas the other one is very poorly sorted due to the pres-
ence of coarse-grained clasts of quartz and feldspars “float-
ing” in the moderately sorted background (Figure 3b). As a
result, the poorly sorted arenites fall into the lower portion
of porosity distribution in this formation, although their
mineral composition and quartz cementation are the same
as those in the better-sorted facies.
It is relevant to mention here another useful classifica-
tion scheme introduced by Thomas and Stieber (1975). This
300 μm scheme is based on the crossplot of “shale volume” versus
porosity to describe three trends jointly making a triangle
b) that is supposed to contain almost all possible siliciclastic
rocks, including those with laminar and pore-filling shale.
The term “shale” in this classification is loosely defined and
basically indistinguishable from the clay aggregates in the
matrix. The main difficulty with the Thomas-Stieber plot as
a classification tool is in its requirement to predetermine the
end-member porosity of both pure sand (or sandstone) and
pure shale. This requirement is especially challenging in se-
quences that have undergone the advanced diagenetic pro-
300 μm
cesses of compaction and quartz cementation in an interval
of interest, thereby necessitating the introduction of several
Figure 3.  Photomicrographs of turbidite sandstone cores overlapping triangles and creating detrimental consequenc-
taken 2 m apart: (a) moderately sorted, fine-grained arenite es for uncertainty analysis.
with a porosity of 16.5%, and (b) poorly-sorted, fine/coarse- Arenites comprise the vast majority of siliciclastic
grained arenite with a porosity of 13.6%. Both sandstones reservoirs worldwide and, as mentioned above, can be at-
are characterized by less than 4% quartz cement. Courtesy
tributed to a wide range of depositional environments,
of T. Kosanke. Used by permission.
from deepwater turbidites to fluvial sediments. The prac-
tical significance of clean arenites is reduced to eolian
sorted than arenites are. However, such an inclusion intro- sands — that is, to the most mature dune deposits.
duces confusion in many rock-physics publications because Wackes can be reasonably high-quality reservoirs
the effect of clay on rock elasticity can interfere with the only if (1) they are very poorly consolidated, (2) their
effect of grain sorting. As an example, Figure 3 shows two clay content is in the lower half of the range (i.e., is less
thin-section photomicrographs taken from cores sampled than 20%), and (3) a significant portion of clay occurs
2 m apart in a turbidite sandstone formation. The amount of inside of lithic or feldspar grains as structural clay, there-
clay in these sandstones does not exceed 7%, so both of by leaving most of the pore system unclogged.
them fall into the group of arenites. However, the sandstone Finally, conventional shales with a total organic
shown in Figure 3a is described as being moderately sorted, carbon (TOC) content ranging from less than 1.0% to

01-Vernik_Ch01.indd 3 12-08-2016 21:03:47


4 Seismic Petrophysics in Quantitative Interpretation

1.5% are the nonreservoir lithologies; their significance is illite/chlorite/kaolinite clay aggregate with imperfect par-
reduced to seal capacity because of the extremely high ticle alignment, as reported by Vernik and Kachanov
capillary pressure in these ultra-fine-grained lithologies. (2010), in combination with an assumption of 15% an-
Nonetheless, the dominance of shales in clastic sedimen- isotropy (ε = γ = 0.15 ). As will be demonstrated later in
tary basins makes them as important as the other silici- this book, these values of polar anisotropy (transverse
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

clastic rock groups are, as subjects of seismic isotropy or TI) are typical of shales with porosity range
petrophysics. from 15 to 20%, which is the median range shown in
Figure 4. The isotropic-equivalent matrix bulk and shear
moduli of Km = 40.1 GPa and Gm = 10.2 GPa, ­respectively,
Application to the core/log database obtained this way, compares well with the “isotropic il-
Because velocity- (or impedance-)versus-porosity lite” estimates by Eastwood and Castagna (1987). These
plots lie at the heart of rock-physics modeling, let us start ­matrix-moduli estimates result in a zero-porosity inter-
with the application of the petrophysical classification in cept of 3.85 km/s for the lower bound when a clay aggre-
water-saturated compressional-velocity-versus-porosity gate grain density of 2.70 g/cm3 is used.
or VP -φ space. As Avseth et al. (2005) and Mavko et al. The upper bound selected for the database is the mod-
(2009) suggested, an elegant way to initiate the workflow ified Voigt model, also known as the critical-porosity
is to provide the appropriate bounds to the data for the model (Nur, 1992) because it terminates at φ = 0.4 (the
first level of quality checking. porosity of a moderately-sorted, unconsolidated sand)
Our database, which is also employed extensively in ­instead of at φ = 1.0. This bound is computed from the
other chapters of this book, is superposed with theoretical matrix elastic moduli and density of the isotropic quartz
bounds, as shown in Figure 4. The Hashin-Shtrikman aggregate, which is appropriate for clean arenites.
lower bound is tailored for hypothetical shale with a vcl of The database of measurements at the confining stress
100%, and it assumes isotropy; that is, the isotropic-­ of 40 MPa for sands and 20 to 40 MPa for shales, shown
equivalent matrix bulk modulus of this shale is computed in Figure 4, does not violate the Hashin-Shtrikman lower
first. This computation is based on the bedding-normal bound. Moreover, the fact that this theoretical bound lies
stiffnesses of C33m = 33.4 GPa and C44m = 8.5 GPa for an approximately 0.7 km/s below the shale cloud can be
readily explained by the natural lack of shales with a vcl
value of 100% and the few data points with vcl values
greater than 80%. On the other hand, the upper bound,
being semi-­empirical in nature regarding the selection of
critical porosity, provides a nice cap for data distribution —
it is a good test of data quality, although we should note
that some of the scatter could be accounted for by the
presence of subordinate minerals such as feldspars, micas,
calcite, and the like.
The petrographic data in the database come from thin
sections and X-ray diffraction (XRD) analyses and are bi-
ased toward arenites (Figure 5). Note that net clay content
on a solid-rock basis is used in the histogram, which
shows a wide range of vcl values, from 30 to 90%, for the
group of shales. Those shales have an average vcl of
52% ± 16% (the plus-minus term represents the standard
deviation here and throughout).
Because microporosity cannot be resolved in a thin
section, the only way to estimate this quantity is to com-
Figure 4.  Compressional velocity versus porosity for 164 bine the helium porosimetry of dry rocks, which yields a
samples of siliciclastics, compiled from Han et al. (1986), good estimate of the total porosity φ, with thin-section
Klimentos and McCann (1990), Issler (1992), Domenico point-counting of macroporosity φ m and bulk volumetric
(1977), and Vernik and Kachanov (2010). Nur’s (1992) clay content C. It is important to emphasize that C typi-
critical-porosity model for clean quartz arenites and the cally is estimated from thin-section point-counting analy-
Hashin-Shtrikman (HS) lower bound for shales with a vcl sis and includes microporosity φµ , which is the difference
of 100% provide upper and lower bounds, respectively. between total porosity and macroporosity:
The term Km_iso is the isotropic-equivalent bulk modulus
φµ = φ − φ m . (1)
of an anisotropic clay matrix.

01-Vernik_Ch01.indd 4 12-08-2016 21:03:47


Chapter 1: Petrophysics of Siliciclastic Rocks 5

Table 1.  Density of major siliciclastic rock-


forming components. Data are from Alexandrov
and Ryzhova (1961), Carmichael (1989), and
Schlumberger (2000).
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Solid phase ρm (g/cm3)


Quartz 2.65
Plagioclase 2.62–2.67 
K-feldspar 2.58
Calcite 2.71
Dolomite 2.86
Muscovite 2.81
Illite 2.67
Illite/Smectite 2.52–2.60
Kaolinite 2.62
Figure 5.  Histogram of volumetric clay content (on a solid- Chlorite 2.84–2.92
rock basis) for the compiled data set used in this study, with
identification of the petrophysical groups of siliciclastics and Pyrite 5.01
illustrating the scarcity of petrographic data on conventional Anhydrite 2.97
shales. Zeolite 2.15–2.25
Coal 0.90–1.50
From this, the microporosity of clay aggregates φ cl can be Kerogen 1.10–1.55
estimated as

φ − φm C (1 − φcl )
φ cl = (2) vcl = . (5)
C 1−φ

with adequate accuracy, unless C is a small value. Analysis A simplification for the petrophysical group of shales can
of the global database carried out this way yields an aver- be derived from equation 5 if their macroporosity φm is
age φ cl value of 40% ± 25% for arenites and somewhat reduced to zero, so that the porosity of their clay aggre-
smaller values for wackes, with naturally higher v­ alues of gates φcl is φ / C and the relationship between the bulk clay
φ cl in shallow, higher-porosity rocks and lower values in content and net clay content on the solid-rock basis then
more-compacted rocks. The clay microporosity values in is C = vcl (1 − φ ) + φ .
excess of 40% for arenites are perhaps overestimated be- Alternatively, these clay-content parameters can be
cause the petrographic technique tends to slightly under- verified by XRD data, which can be regrouped in terms of
estimate φ m, although values in excess of 50% have been clay and nonclay fractions given in weight percent. Then,
reported in some high-porosity arenites in the North Sea using individual mineral and solid-phase (matrix) densi-
province (Ehrenberg, 1990). ties, we obtain
Another useful shaly rock parameter in petrophysics
is Vcl, which, if we denote the rock constituents other than ρm
Vcl = vcl (1 − φ ) = wcl (1 − φ ), (6)
clay as Vncl, can be determined from ρcl

Vcl + Vncl + φ = 1, (3) where wcl is the weight fraction of total clay from XRD,
ρm is the solid-phase density of the rock, and ρcl is the
which relates to vcl (the volume of clay for the solid-rock weighted mineral density of the clay composite. The
matrix) as follows: dry mineral density values used in this book are given in
Table 1. Note that the term “dry” with respect to such
Vcl minerals as clays and zeolites refers to the conditions
vcl = . (4)
1−φ under which loose, adsorbed water is excluded and ac-

counted for as a portion of the total porosity of the rock,
Because, by definition, Vcl = C − φµ = C (1 − φcl ) , the but the chemically bound water molecules and hydroxyl
clay fraction of the solid-rock basis equals groups are still considered to be solid-rock constituents.

01-Vernik_Ch01.indd 5 12-08-2016 21:03:49


6 Seismic Petrophysics in Quantitative Interpretation

Tosaya and Nur (1982), Han et al. (1986), and Eberhart- makes equations 7 rather dubious, notably in shale pore
Phillips (1989), among others, studied the effect of clay on pressure prediction from sonic or seismic velocities.
the elastic properties of water-saturated consolidated silici- Han et al. (1986) provide the following linear-­
clastics by using ultrasonic velocity measurements at dif- regression fits (in km/s) to the data set of wet consolidated
ferent confining pressures. The linear-­regression equations sandstones and shales that they measured at a confining
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

derived by Eberhart-Phillips (1989) were based on the mul- pressure of 40 MPa:


tivariate analysis of Han et al.’s measurements:
V P = 6.08 − 8.06φ (C = 0),
1 (8)
VP = 5.77 − 6.94φ − 1.73C 2 + 0.446 [σ − exp( −16.7σ )] , V S = 4.06 − 6.28φ ,
1
V = 3.70 − 4.94φ − 1.57C 2 + 0.361 [σ − exp( −16.7σ )] , and
S
(7) V P = 5.59 − 6.93φ − 2.18C (C = 5 to 55%),
(9)
where velocities are in kilometers per second (km/s) V S = 3.52 − 4.91φ − 1.89C .
and σ is the mean effective stress in kilobars (kbar).
As follows from Figure 6, for the database compiled These empirical relationships are superior to those given
for this book (including the Han et al. data), the Eberhart- by Tosaya and Nur (1982) and Eberhart-Phillips et al.
Phillips predictions increasingly mismatch the experi- (1989) in fitting Han et al.’s (1986) data set of approxi-
mental data as clay content increases from clean arenites mately 85 samples. However, it should be emphasized
to shales. Using equations 5 and 6 for the median shale that the thin-section-based bulk clay fraction C is not a
porosity of approximately 15%, we can see that the petro- unique function of the verifiable clay content vcl. The lat-
graphically determined C of 60% actually corresponds to ter, given by equation 5, is a more robust variable that is
a φcl of 25% and a vcl of 53%, which are the shale-subset required in any accurate petrophysical model building.
averages. However, the shale lines obtained from equa- If we use equation 5, and we assume that φcl = 0.4 and
tion 7 for 20 and 40 MPa obviously overpredict the bed- vcl = 0.12, the porosity-dependent bulk clay volume is cal-
ding-normal P-wave velocity measured in shales, both in culated as 0.2(1 − φ ) . Then, substituting this variable into
situ and in laboratory experiments. This observation Han et al.’s transform for shaly sands, we can compute a
model line that provides a remarkably accurate demarcation
between arenites and wackes in water-saturated VP-versus-
porosity space (Figure 7). Moreover, a similar exercise for
shales, in which, as we have seen above, C = vcl (1 − φ ) + φ ,
strongly confirms the significance of a vcl value of 30 to
35% as another important boundary in the VP -φ space sepa-
rating wackes from shales measured in the bedding-normal
direction. It is also intriguing that the evaluation of equa-
tions 9 for a vcl value of 50% results in reasonably good
linear fits to the shale measurements of S-wave velocity
and, especially, of P-wave velocity. Interestingly, that evalu-
ation for shales has a slope that is almost parallel to the
slope for clean arenites from equation 8 (Figure 7).
It is noteworthy that some of the original measure-
ments on very shaly sandstones from Han et al.’s (1986)
data set, which in the classification scheme presented
above and shown in Figure 1 belong to the group of shales,
had no account of anisotropy. This becomes obvious when
Figure 6.  Compressional velocity VP versus porosity φ comparing Han et al.’s shale velocities with those from
for 152 consolidated siliciclastic rocks, color-coded by the Issler (1992) and Vernik and Kachanov (2010), which
petrophysical groups. All shale data points refer to the were measured by sonic logs in vertical wells in the ap-
bedding-normal direction. Linear models are from Eberhart- proximately bedding-normal direction. To compensate
Phillips (1989), computed for 40-MPa effective stress for all for the discrepancy observed and the fact that many shale
bulk clay contents except C = 0.6 — for which a 20-MPa line core plugs in Han's collection were drilled horizontally, a
is added because most of the log/core measurements in shales minor anisotropic correction was carried out for some of
were associated with a stress range of 20 to 40 MPa. the shale samples reported by Han et al. (1986).

01-Vernik_Ch01.indd 6 12-08-2016 21:03:50


Chapter 1: Petrophysics of Siliciclastic Rocks 7

a) position of clay in those arenites. That inconsistency is an-


other reason why equations such as 7, 8, and 9 have not
gained traction in seismic petrophysics.
Note that the classification of siliciclastics proposed
here discards their compaction, diagenesis, grain size,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

and grain-size distribution; hence, each petrophysical


group includes a wide range, from unconsolidated sedi-
ments to their well-indurated, low-porosity equivalents
with vastly varying grain-sorting characteristics. That fact
makes the scheme especially attractive for investigating
the effects of compaction and diagenesis on the physical
properties of rocks, independently of the complicating
factors such as clay content and textural features.
On the basis of nonlinear extrapolation to zero poros-
ity for the individual petrophysical groups, Vernik (1998)
obtained certain matrix properties that are characteristic
of these groups to simplify the rock-physics modeling
workflow in siliciclastic sequences (Table 2). Indeed,
b) these estimated zero-porosity properties show remarkable
stability worldwide and can be complemented with tradi-
tional Hill’s and Backus’ averaging if the mineralogy data
are available and/or the presence of some exotic mineral
is suspected (Backus, 1962; Hill, 1963; Smith, 2011).
In Table 2, Mm, Gm, and ρm are, respectively, the
P-wave modulus, shear modulus, and density of the rock
with total porosity approaching zero. Referring to the his-
togram of clay content (Figure 5), the low-clay shales can
be characterized by the vcl range from 30 to 45%, whereas
the high-clay shales lie in the range from 45 to 70%.
Subordinate shale types with a vcl value exceeding 70%
are conditionally omitted.
P-wave and shear moduli for shales in Table 2 refer
to the bedding-normal direction in these typically aniso-
tropic rocks (C33m and C44m elastic stiffnesses respective-
ly). The hypothetical shale with a vcl value of 100% and
strong but imperfect clay-particle alignment was derived
Figure 7.  Water-saturated velocities (a) VP and (b) VS versus from C ij = f (φ , v cl ) modeling by Vernik and Kachanov
porosity φ, showing the database with a breakdown according (2010). The somewhat elevated matrix density of shales
to the petrophysical classification scheme. The lines are Han is related to the nonnegligible concentrations of chlorite,
et al.’s (1986) empirical equations for an effective stress of carbonate, and, especially, pyrite in these lithologies
40 MPa (all sands) and of 20 MPa (shales) with the improved (Herron, 1987). In that respect, it is hard to justify the
clay-volume definition. The realizations for net clay contents frequently selected value of 2.65 g/cm3 as a convenient
of vcl values at 12% and 30% provide excellent demarcations choice for shale grain density in standard petrophysics
among arenites, wackes, and shales, whereas the clean arenite and rock-physics modeling.
and shale lines with respective vcl values of zero and 50% The ever so important solid-matrix properties of main
become almost parallel to each other.
lithotypes in siliciclastic rocks will be used extensively in
the part of this book devoted to elastic rock properties and
It is obvious that for a given porosity, elastic-wave ve- rock-physics relationships (Chapter 3).
locities and permeability both decrease considerably in the To summarize,
sequence of clean arenite/arenite/wacke/shale. However,
as Vernik and Nur (1992b) show, that reduction is not con- • The petrophysical classification of siliciclastics is
tinuous. For instance, no strong correlation exists between quite robust in mapping one of the most important
velocity and clay content in the most important group of crossplots in seismic petrophysics — velocity versus
arenites, likely because of the peculiarities in the structural porosity — in terms of lithologic facies.

01-Vernik_Ch01.indd 7 12-08-2016 21:03:51


8 Seismic Petrophysics in Quantitative Interpretation

Table 2.  Solid-matrix properties of major petrophysical groups of siliciclastic rocks.


Petrophysical group vcl Mm (GPa) Gm (GPa) VPm/VSm ρm (g/cm3)
Clean arenites <0.02 96.7 43.4 1.49 2.65
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Arenites 0.02–0.12 79.7 33.0 1.55 2.65


Wackes 0.12–0.30 68.9 25.3 1.65 2.67
Low-vcl shales 0.30–0.45 55.4 16.5 1.83 2.73
High-vcl shales 0.45–0.70 45.9 12.5 1.92 2.73
Pure-clay shale 1.00 33.4 8.5 1.98 2.73

• Bulk clay volume, which includes microporosity, is in- basis. Vsh is introduced as a separate parameter to describe
corporated into some of the empirical velocity-­versus- thinly laminated reservoirs; however, it also is often used as
porosity transforms but is not an adequate ­parameter in a vague petrophysical term that includes clay and silt. A
rock physics. On the other hand, the net clay volume on laminated lithofacies with a layer thickness at least an order
a solid-rock basis, which can be verified by means of of magnitude lower than the minimal log resolution of 0.5
quantitative mineralogy and can be used in petrophysi- to 1.0 m can only be evaluated in the sense of effective rock
cal log-model calibration, is the proper parameter to use. properties — that is to say, properties of the individual lam-
• Published empirical transforms to account for clay- inae can only be estimated by using some effective volu-
fraction effects are not quite accurate as presented. metric property at the log scale. The laminar shale-volume
Surprisingly, however, such transforms can be useful fraction, or Vsh, can serve this purpose as a lithologic vari-
in lithology mapping in VP -φ and VS -φ spaces, if the able, as opposed to Vcl (or vcl) that is a mineralogical one.
porosity-independent, net clay content is used instead There are several ways to compute clay and shale frac-
of the bulk clay content. tions in sedimentary rocks from log data. One of the most
• Matrix properties of the main siliciclastic rock groups popular models is a linear interpolation between the clean
can be further employed in rock-physics modeling, gamma-ray-log (GR) and the shale-gamma-ray-log re-
including modeling of the effects of diagenesis, sponses, which typically are taken as the minimum (GRmin)
where some constraints on the end-member proper- and the maximum (GRmax) values, respectively. In such a
ties must be imposed. case the gamma-ray index G is conventionally computed as:
GRn − GRmin
G = , (10)
Petrophysical model building
GRmax − GRmin

Accurate determination of the key reservoir and seal where GRn is the normalized gamma-ray log. The shale
parameters is imperative not only in reservoir evaluation but fraction can then be computed as Vsh = G (with lower and
also in rock-physics modeling and quantitative seismic upper limits of zero and one, respectively) if a linear
analysis. The workflow typically invoked in petrophysics transform is found to be adequate. Alternative models
when we are dealing with standard log suites is character- (see, e.g., Larionov, 1969) relate Vsh and G in a nonlinear
ized by the combination of fundamental and empirical mod- fashion:
els to compute key parameters necessary in the process of
seismic calibration and reservoir characterization. Seismic Vsh = a (2bG − 1), (11)
petrophysics implies that proper log conditioning and anal-
ysis are performed from the top to the base of the measured where a = 0.33 and b = 2.0, but, in general, are adjustable
data, instead of just for the target zone (Smith, 2011). parameters. Proper selection of the parameters entering
Let us address the workflow step by step, realizing equations 10 and 11 is critical for accurate computation of
that the deterministic log-model code may involve some the shale volume. Unfortunately, shale volume is often
feedback computation with iterations. Again, we will be also referred to as Vcl in the petrophysics literature, and
concerned only with siliciclastic sequences. that is a persistent pitfall. Even more confusion is intro-
duced when lumping all clay mineral constituents togeth-
Lithologic parameters Vcl and Vsh er with loosely-bound water and naming this composite
the shale volume of the rock.
The terms vcl and Vcl represent primarily the clay con- Vernik and Hamman (2009) have introduced a slight-
tents on a solid-phase and bulk-rock basis, respectively. ly modified approach to the definition and computation
Vshale (or Vsh) is a shale-volume fraction on a bulk-rock of Vsh. They define the term as the laminar shale volume

01-Vernik_Ch01.indd 8 12-08-2016 21:03:52


Chapter 1: Petrophysics of Siliciclastic Rocks 9

fraction. In other words, it is a composite lithologic vari- a)


able that is distinct from Vcl, which is the “dry” volume
fraction of the total of the clay minerals over an interval
resolvable on the gamma-ray-log scale. The laminar shale
fraction conveniently defines the continuous net-to-gross
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

curve, which is expressed by Vsand = 1 − Vsh .


Both Vsh and Vcl can be measured accurately on cores
using (1) foot-by-foot counting of cumulative shale thick-
nesses to generate Vsh and (2) XRD analysis on individual,
more or less homogeneous thin beds and laminae in the
same core to determine core-scale Vcl (equation 6), which
then can be upscaled to calibrate the log-derived curve. The
distinction between Vsh and Vcl is important, and the values
across which they vary are quite different: Vsh ranges from
zero to one, and both Vcl and vcl (clay content on the solid-
rock basis) range from zero to values much smaller than
one because the presence of silt particles keeps the total of
b)
clay minerals from reaching 100% of the shale.
To appreciate the difference in computation between
the two variables, one can review a typical normalized GR-
log histogram for a 300-m-thick interval from a well drilled
in a Tertiary sedimentary basin (Figure 8a). The normalized
GRmin to GRmax range is shown to be approximately 30 to
130 gAPI (API gamma units), and the bimodal distribution
accounts for the predominance of shales in the interval.
Assuming there is a minor contribution of nonclay
minerals to the natural GR log (which may not be true in
some cases), the fractional dry clay volume Vcl can be
computed simply as Vcl = G, as given by equation 10. In
doing so, a different GRmax value equal to that of pure clay
(e.g., 210 to 250 gAPI for illite/smectite clay [Appendix
B in Schlumberger, 2000], instead of 130 gAPI) will have
to be selected:

GRn − GRmin Figure 8.  Normalized GR histograms for (a) a typical


Vcl = . (12)
GRcl − GRmin Tertiary siliciclastic sequence, and (b) clay volume and

shale volume fractions, as a function of GRn. The values
The range from 210 to 250 gAPI corresponds to a net clay GRsand = 52 gAPI and GRshale = 93 gAPI were selected on the
volume vcl of 100% (i.e., no silt grains). Alternatively, basis of the histogram. Adapted from Figure 1 of Vernik
core XRD data can be used to calibrate the model by com- (2008). Used by permission.
puting an appropriate GRcl value before its substitution
into equation 12. Most typical XRD analyses in massive GRsand can be picked in such a way that, if the values
medium-porosity shales, after weight/volume conversion, greater than this are discarded, the sand mode of the his-
show a Vcl range from 40 to 60%. togram is preserved with its bell-shaped distribution.
Now, the Vcl parameter computed in this way should Similarly, GRshale can be picked at the left side of the shale
be distinguished from the shale volume fraction Vsh that is mode of the histogram.
calculated empirically from threshold values of GRn that For instance, in Figure 8, GRsand = 52 gAPI and
correspond to the transitions (1) from massive sand to in- GRshale = 93 gAPI. If, for example, GRcl = 210 gAPI, the
homogeneous intervals with minor, but distinct individual corresponding Vcl values will be 12% and 35%, respec-
shale beds (GRsand) and (2) from inhomogeneous intervals tively. If Vcl is less than 12%, Vsh is zero, meaning that
with minor, but distinct individual sand beds to massive there are practically no individual shale laminae in the
shale (GRshale), as shown in Figure 8a. interval. On the other hand, if Vcl exceeds 35%, Vsh is one
In the absence of core data, the following general rule (or 100%), so a massive shale interval is interpreted. The
for selecting those threshold GR values is suggested: Vsh values in between the two thresholds can be linearly

01-Vernik_Ch01.indd 9 12-08-2016 21:03:53


10 Seismic Petrophysics in Quantitative Interpretation

interpolated — that is, by computing G from equation 10, whereas massive arenites from channel facies dominate in
which is the same as using V sh = (V cl − 0.12) (0.35 − 0.12). the first well, they are almost absent in the second one.
Alternatively, Vsh is derived from equation 11 if a nonlin- Note also that clay content in these arenites varies from 2
ear interpolation is sought (Figure 8b). to 12%, whereas their shale content is set to zero, which
It turns out that the shale fraction determined in this is consistent with the geologic core description.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

way in conventional shale/sand intervals is remarkably Core photographs of some of the lithofacies shown in
consistent with core descriptions, especially in deepwater- Figure 9 are presented in Figure 10. The splay-sand inter-
turbidite and shallow-marine sequences, where the vast val with a geologically determined shale count Vsh of 13%
majority of heterolithics are composed of thinly laminated is shown in Figure 10a and is characterized by a largely
sand/silt/shales rather than of wackes with pore-filling thin-bedded texture with crossbeds. A medial levee turbi-
clay. Two well-log fragments from two wells in the same dite facies with a laminar shale count of 54% is shown in
deepwater reservoir are shown in Figure 9, where the Vsh Figure 10b, and Figure 10c shows a distal levee facies
curve of the log model is superposed by the core Vsh curve with a laminar-shale Vsh of 75%. These geologically de-
in the first track. The results suggest that the GR-based termined shale counts can be translated into sand counts
model described above can be remarkably accurate. — that is, into the net-to-gross values of each interval.
It is also instructive to compare Vsh with the Vcl curve In certain cases, when the sand component is rela-
included in track four of Figure 9. One can see that tively “hot” on the GR log because of the presence of

Figure 9.  Two fragments of the turbidite reservoir penetrated by two wells in deepwater Gulf of Mexico. The left panel shows
channel-sand facies (yellow, dotted) with Vsh = 0, as well as proximal (brown) and distal (green) levee facies shown in the
Lithology track. The right panel features splay sand (orange, dotted) interlayered with proximal and distal levees and some
intrareservoir shales (gray). Note the relationship between Vsh and Vcl in both intervals, and calibration of the log Vsh by using
core Vsh from the geologic description.

01-Vernik_Ch01.indd 10 12-08-2016 21:03:54


Chapter 1: Petrophysics of Siliciclastic Rocks 11

a) b) c)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 10.  Core photographs showing very fine-grained sand (brown) and shale (gray) thin beds with varying shale counts
(core Vsh), from (a) 13% in splay channel-sand facies, to (b) 54% in medial levee facies, to (c) 75% in distal levee facies. These
geologic shale counts are consistent with laminar-shale Vsh values derived from the log model.

potassium feldspar or radioactive secondary minerals, the sand in a neutron/density porosity plot. If one positions
approach just described may fail and should be replaced the lines to be tangential to shale and wet-sand data
with neutron/density estimations of Vcl and Vsh (Bateman, clouds, as shown, the two intercepts ash and asd can be
1990). In that workflow, the apparent density porosity φ D defined and used in the alternative Vsh evaluation for
that is related to the total porosity of the rock should be heterolithics:
computed first:
(φ N − bφ D − asd )
ρ − ρb V sh = , (15)
φ D = ma , (13) ash − asd
ρma − ρfa
whereas the shale volume should be set to zero for the
where ρ b is the bulk-density log, and ρma and ρfa are the negative values (positions below the sand line) and to
apparent matrix and fluid densities, respectively. To sim- unity when Vsh is greater than 100% (positions above the
plify this step, the latter two parameters can be equated to shale line).
2.65 and 1.00 g/cm3, respectively. Because a neutron po-
rosity log responds to the hydrogen index, which is parti-
tioned between the porosity and the solid-clay-­mineral Total porosity
volume, the following equation can relate density to neu- Another critical parameter used in evaluating subsur-
tron porosity φ N: face formations is the total porosity of the rock. This
φ N = a + bφ D , (14) property of sedimentary rocks is definitely a link between

log evaluation and seismic interpretation, but it also is the
where the intercept a is the clay component of the hydro- focus of any rock-physics analysis. Due to the peculiari-
gen index in terms of neutron porosity and the slope ties of siliciclastic microstructures, total porosity relates
b = 0.875 is relatively stable in siliciclastics that have po- to the interconnected pore volume (but not necessarily
rosities lower than 40%. The example shown in Figure 11 to the “effective” or hydrocarbon-saturated pore volume).
is realized for two different intercepts and overlain by the Thus, the best test for determining it is the helium-­
schematic ellipses indicating the position of shale and wet porosimetry technique implemented on core samples.

01-Vernik_Ch01.indd 11 12-08-2016 21:03:56


12 Seismic Petrophysics in Quantitative Interpretation

0.7
ρm_sh = 2.76 + 0.001{( ρb − 2) − 230 exp[−4( ρ b − 2)]}, (18)
0.6
Shale where the first term can vary from 2.70 to 2.79 g/cm3 and
ρ b is the bulk density in grams per cubic centimeter.
0.5
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Perhaps this model for estimating matrix density should


0.4 be preferred to that for the constant-shale-matrix density,
because this model accounts for the gradual increase in
φN

0.3 ρm_sh with clay-diagenesis indicators, such as the illite/


Wet
sand
smectite ratio in mixed-layer clays. That type of chemical
0.2 diagenesis in shales is a slow process accompanied by
0.875 mechanical compaction, although evidence also has been
0.1 reported for faster phase-transformation reactions occur-
ring in a relatively narrow depth range (e.g., Dutta, 2002).
0 An even greater uncertainty exists with the fluid-­density
0 0.1 0.2 0.3 0.4 0.5
term entering equation 16. Indeed, even if water saturation
φD
and the fluid composition in the reservoir or aquifer are
Figure 11.  Density-porosity-versus-neutron-porosity plot, known precisely, a challenge remains for predicting the
schematically showing expected positions of massive wet mud-fluid-invasion profile. Because the effective depth of
sands and shales. Two linear trends with different intercepts investigation for a typical density tool does not exceed
should be approximately parallel to each other if porosity is 13 cm (5 in) but the mud-filtrate invasion can create a much
less than 40%; these lines can be selected, as shown, to greater annulus around the hole, it becomes imperative to be
compute Vsh values in heterolithics. able to account for the alteration of fluid density.
Again, if formation properties are well known — in-
The term "effective porosity", used in some petrophysical cluding core measurements for capillary pressure, relative
models, is intentionally discarded in this book because of permeability, and fluid viscosity — then the pressure dif-
the great uncertainty with determining the microporosity ferential during drilling can be used to numerically pre-
of clay aggregates. This uncertainty propagates into and is dict the mud-invasion profile (e.g., Malik et al., 2008).
further amplified in the domain of rock physics, notably Most often, however, those data are not available and phe-
in estimating elastic properties of solid rock constituents, nomenological models may be constructed to estimate the
as will be discussed later. mud-invasion saturation in the volume around the well
In petrophysics, the total porosity is most accurately that is “visible” to the density logging tool:
computed by using the density log:
Sinv_ma x − Sinv_min
Sinv ( R) = + Sinv_min ,
ρm − ρb  R
c
(19)
φ = , (16) 1+  
ρm − ρfi  Ra 

where ρm and ρfi are the matrix and fluid densities, respec- where Sinv_max and Sinv_min are the filtrate saturations at the
tively, of a representative rock volume adjacent to the bore- borehole wall and at some significant distance from it,
hole wall. If the formation is laterally homogeneous on at respectively; Ra is the distance from the borehole wall
least a 0.25-m scale (the depth of investigation of the den- where the average saturation between the minimal and
sity log) and is vertically homogeneous on at least a 0.5-m maximal values is attained, and c is the exponent defining
to 0.75-m scale (the vertical resolution of the density log), the profile shape. It is generally safe to assume that Sinv_min
the matrix density is representative of the layer in situ. tends to approach zero with distance.
In the absence of core data, one can estimate matrix An example of such a profile is given in Figure 12.
density by using Table 2 for homogeneous intervals, with The most probable mud-filtrate saturation value in medi-
the following extension for thinly laminated ones: um- to high-porosity arenites at the borehole wall is ap-
proximately 50% ± 20%. Values in excess of 50% can be
ρm = 2.65 (1 − V sh ) + ρm_sh V sh, (17) attributed to the lower-porosity arenites or medium-­

porosity wacke sandstones and are largely the result of
assuming that ρm_sh = 2.73 ± 0.03 g/cm 3. An empirical less effective mud-cake formation. Simultaneously, the
equation for the matrix density of shale also has been pro- invasion radius increases, which can be controlled by Ra
posed that is designed to avoid the functional dependence and exponent c in equation 19.
on porosity, which is unknown at this stage of model The invasion profile shown in Figure 12 suggests that
building (Vernik and Kachanov, 2010): in the 10- to 15-cm annulus around the borehole wall, the

01-Vernik_Ch01.indd 12 12-08-2016 21:03:57


Chapter 1: Petrophysics of Siliciclastic Rocks 13
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 12.  Mud-filtrate-saturation (Sinv) profile, based on Figure 13.  Effect of water-based mud-filtate invasion on
phenomenological equation 17. The most probable Sinv bulk-density readings from a density logging tool in a
saturation is shown in red, and the two brown dashed lines hydrocarbon reservoir. Three lines were computed with
indicate a plausible range. In situ water saturation Sw is equations 16 and 20 for the porosity range of 15 to 35% and
shown in blue. Shc is hydrocarbon saturation. by setting the other parameters as indicated.

fluid density to be used in the total-porosity equation 16 value of 50% the relative error in bulk density will range
must be estimated as follows: from 1 to 3% and the relative error in porosity will be ap-
proximately 8%.
ρfi = S w ρw + (1 − S w − S inv ) ρhc + S inv ρmf , (20)
Consequently, the relatively simple equation for
computing the total porosity from bulk density measured
where ρw, ρhc, and ρmf refer respectively to the densities by the density logging tool has many pitfalls, which, in
of brine, reservoir hydrocarbon, and mud filtrate. The ef- the absence of core data on helium porosimetry, may re-
fect of mud invasion on fluid density is the greatest when sult in a relative error as high as 10% if the mud-invasion
a hydrocarbon reservoir is drilled with a water-based mud effect is neglected. It is safe to conclude that accounting
system (WBM) or an aquifer is penetrated using oil-based for variations in grain density and fluid density, espe-
mud (OBM). On the other hand, this effect is practically cially in inhomogeneous reservoirs penetrated by con-
nonexistent in shales in which the last two terms in equa- trasting mud systems, is critical for accurately determin-
tion 20 vanish and ρfi is equal to ρf , which is uninvaded, ing porosities, correcting density-log invasions, and,
in situ fluid density. hence, for rock physics and forward modeling of seismic
The term (1 − Sw − Sinv ) in equation 20 represents the amplitudes.
residual hydrocarbon saturation around the borehole.
Because water saturation normally is determined by using
some version of the Archie equation, it is obvious that Permeability prediction in siliciclastics
several iterations will have to be engaged in order to ar-
rive at a reasonable Sw curve. It is worth mentioning that Permeability is one of the key parameters we need to
shallow resistivity logs are not effective in the 10- to estimate during formation evaluation and characterization
15-cm annulus around the well, because of their greater because it provides us with important information about
depth of investigation. flow rates from the reservoir. It is also frequently used in
Let us consider an oil reservoir penetrated by a bore- net-pay prediction from logs. It may seem that its effects
hole drilled with WBM. The effect of the mud-filtrate in- in quantitative seismic interpretation are secondary —
vasion on density-log readings will, of course, depend on after all, there is no straightforward theoretical link be-
the total porosity, as is illustrated in Figure 13. The true tween permeability and elastic wave velocities. However,
formation bulk density in situ is given by the intercepts, permeability impacts certain viscoelastic effects that con-
but because of the invasion of a much denser fluid, the trol velocity dispersion as a function of frequency (Biot,
density tool will read higher. Thus, at our default Sinv 1956; Mavko et al., 2009).

01-Vernik_Ch01.indd 13 12-08-2016 21:03:58


14 Seismic Petrophysics in Quantitative Interpretation

Log-based permeability prediction is a daunting task


that has been addressed in many publications (e.g., Timur,
1968; Nelson, 1994; Herron et al., 1998). Practically all
the models that use the standard log suite, with or without
core calibration, are empirical in nature, although most of
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

them originate from the functional form of the theoretical


model for a bundle of capillary tubes (widely referred to
as the Kozeny-Carman equation; Carman, 1956; Berg,
1970; Van Baaren, 1979). Along those lines, some models
attempt to relate the log-derived clay content to the grain
size in siliciclastics and then use the latter as a proxy for
the pore-throat-size distribution in permeability predic-
tion (Vernik, 2000).
In the following discussion these models are used as
a foundation, with the primary focus being on grain-sup-
ported siliciclastic rocks — arenites and clean arenites.
For simplicity and clarity, the term permeability will refer
to the absolute (or air) permeability.
The Kozeny-Carman equation suggests that in addi- Figure 14.  Permeability versus porosity for a data set of
tion to total porosity φ, another rock property (surface clean arenites (with a quartz content greater than 98%),
area, dominant grain size, or relevant pore dimension) superposed by the optimized Kozeny-Carman model as given
must be incorporated into the permeability prediction: in equation 21 (B = 4).

Dm2 φ 3 data of Beard and Weyl (1973) on unconsolidated sand


k = B , (21)
(1 − φ 2 ) packs that exhibit varying grain size and sorting.
A slightly modifed Van Baaren model could be even
where Dm is the mean grain size in microns (μm), which more useful and applicable to a wider range of porosity
yields permeability in millidarcys (md) and B is the unit- and/or consolidation:
less constant related to the tortuosity. However, B can be
deemed an empirical parameter, because equation 21 k = 30 Dd2φ b , (23)
loses its strict theoretical sense when it is applied to gran-
ular materials rather than to capillary-tube bundles. The with b = 4 in clean arenites and b = 5.1 in arenites, inde-
overall validity of using grain size in conjunction with pendently of consolidation. This model is applied to the
porosity as the controlling influence on permeability has same data set of clean arenites in Figure 15. The figure
been confirmed in many laboratory experiments (e.g., shows a major improvement in fitting the data in a wide
Beard and Weyl, 1973). porosity range, with outliers only at extremely low po-
It is instructive to apply equation 21 to the well-estab- rosities of less than 5% — that is, when porosity intercon-
lished data set composed of clean arenites (Yale, 1984; nectivity becomes questionable.
Bourbie et al., 1987), as is demonstrated in Figure 14. The performance of this modified Van Baaren model
Evidently this model starts to systematically deviate from is further tested using a vast database of arenite data ac-
the ground truth at a porosity smaller than 10% — a prob- cumulated over the last 20 years. In addition to porosity
lem noticed by Bourbie et al. (1987) and for which they and permeability data, the database includes mean grain
suggested the variable porosity exponent to match the size from laser particle-size analyses. Figure 16 shows
data. Interestingly, the porosity exponent in Van Baaren’s this database superposed by the modified Van Baaren
equation, which originates from the Kozeny-Carman for- model realized for arenites (b = 5.1) in the Dm range, from
mula, is allowed to vary with Archie’s cementation expo- coarse silt (31–62 µm) to coarse sand (250–500 µm).
nent m (Van Baaren, 1979): For comparison, lines from the Winland model (e.g.,
Nelson, 1994) also are superimposed onto Figure 16. The
k = 10 Dd2 S −3.64φ 3.64 + m , (22) Winland model takes advantage of the pore-throat radius (in
μm), at 35% mercury saturation, computed on the basis of a
where Dd is the dominant grain size (in μm) and S is the mercury-injection test. The Winland model is accepted as
sorting coefficient, which ranges from 0.7 in very-well- being more indicative of permeability than the grain size is:
sorted sediments to 1.0 in poorly sorted sediments. This 2
model is remarkably consistent with the experimental k = 50 R35 φ 1.47 . (24)

01-Vernik_Ch01.indd 14 12-08-2016 21:03:59


Chapter 1: Petrophysics of Siliciclastic Rocks 15
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 15.  Permeability versus porosity for a data set of Figure 17.  The same data as in Figure 15, color-coded here
clean arenites (with a quartz content greater than 98%), by the mean grain size (μm) and with an added data set of
superposed by the modified Van Baaren equation (equation tight-gas sandstones but without information on Dm (black).
23) with b = 4.

quite limited by the lack of data on R35, notably in log


modeling.
A permeability-versus-porosity crossplot for our da-
tabase, colored-coded by the mean grain size and includ-
ing added data on tight gas sands, is shown in Figure 17.
The added data set is not complemented by the Dm data.
Analysis of Figure 17 suggests that the modified Van
Baaren model for arenites provides an excellent fit to the
vast majority of data, except for tight sandstones with a
porosity of less than 12%. Note that the porosity thresh-
old, below which this model starts to overpredict perme-
ability, is pushed upward from approximately 5% in clean
arenites. A possible explanation of this observation is
likely to be related to the elevated role of clay, even a
small amount of which can have a strong effect on tortu-
osity and pore connectivity at very low porosities.
Further investigations can shed light on whether using
the original relation by Van Baaren (1979) improves upon
Figure 16.  Database of arenites in permeability-versus- the simplification suggested here. One can argue that dra-
porosity space, showing a modified Van Baaren model (blue) matic variations in sorting parameters must be taken into
and a Winland model (red) realized for pore-throat radii of account. However, it seems that the modification present-
R35 = 0.25, 2.5, and 25 µm. ed in equation 23 is good enough to warrant its use, pro-
viding that the distinction is made between the porosity
Here, R35 is the pore-aperture radius at which the nonwet- exponents for clean arenites and those for arenites.
ting phase saturation is 35% of pore–pore throat volume. This statement is exemplified by the crossplot of
The porosity exponent in this model is much smaller than measured permeability versus predicted permeability
those used in previous equations, thereby resulting in the shown in Figure 18, wherein only arenites with porosities
much gentler, “semilogarithmic slopes” seen in Figure higher than 12% are plotted. The correlation is excellent,
16. This model provides an important insight into the re- but most importantly, a desirable prediction accuracy is
lationship between the mean grain size and mean pore- achieved — approximately 25% of the order of magni-
throat size in arenites. However, its predictive power is tude of permeability.

01-Vernik_Ch01.indd 15 12-08-2016 21:04:01


16 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 19.  Mean grain size versus clay fraction for a subset
Figure 18.  Permeability from the modified Van Baaren
from the database. Note that the straight line given by the
model versus measured permeability. Prediction accuracy
power-law model in this logarithmic plot provides a fairly
can be compared with the line that has a slope equal to unity.
good fit to the data for massive arenites (Vcl < 0.10–0.12), but
The root-mean-square (rms) error is 0.257 of the order of
it breaks down in heterolithics as a result of sampling bias.
magnitude of permeability. The coefficient of correlation
R = 0.947.
curvilinear models displayed in the logarithmic plot of
Nonetheless, the question remains: How can we use Figure 19 in this domain are, in fact, linear weighted aver-
this model in standard log-based permeability estimation? ages between Vcl values in the range 10 to 12% (with
A valuable insight can be gained when we are relating Vsh = 0 and Dm given by the model for arenites) on one
mean grain size to the clay content, which, as we dis- hand and Vcl values of 40 and 60% (with Vsh = 1 and
cussed above, can be computed using GR and/or neutron/ Dm = 1 µm) on the other. These are displayed just for il-
density separation. lustration — they are not predictive models.
A subset of approximately 60 core plugs in our data- Substitution of the relationship Dm = f (Vcl ) in aren-
base has measurements of Vcl (derived from XRD) and Dm ites into equation 23, with b = 5.1, results in
(obtained from laser particle-size analysis, or LPSA); all
samples came from deepwater turbidite sequences. This k = 30(2.3 V cl−1.35 )2 φ 5.1. (26)
subset must be evaluated with caution, because it can suf-
fer from sampling bias in thinly laminated sand/shales This equation is evaluated in Figure 20 for a range of clay
(typical of levee facies) and also from uncertainty associ- fractions, from 2 to 20%, with the same database that was
ated with errors in clay fractions measured from XRD used in Figure 17. Again, with the exception of very-low-
data. However, in homogeneous (on the log scale) aren- porosity sandstones, this model is fairly accurate in brack-
ites from channel or subsurface fan facies, finding a rela- eting arenites between two lines with Vcl values of 2 and
tionship between the clay content and grain size is rela- 12%.
tively straightforward. Good correlation between the grain size and clay con-
From the sedimentologic point of view it is almost a tent in the petrophysical group of arenites should not be
common wisdom that the mean grain size of arenites surprising, because both parameters are largely controlled
should correlate well with clay content. Thus, the follow- by the depositional environment, which can range from
ing empirical, power-law model can be advanced: high to low energy in deepwater turbidity currents (Blatt,
1982). However, the relationship presented may need to be
Dm = 2.3 Vcl−1.35 , (25) adjusted for arenites found in depositional environments
other than deepwater turbidites. Another pitfall of using
for Vcl values smaller than 10 to 12% and with Dm values equations 25 and 26 is that they do not account for pecu-
expressed in microns. This value of clay content should liarities of arenite composition above and beyond total
be a bifurcation point, at which thin shale laminations are clay content. The origins of the clay in this group of sands
introduced into the massive sand so that a further increase may vary — from sedimentary (e.g., shale clasts) to
in Vcl values will give rise to nonzero values of Vsh. Two ­diagenetic (e.g., authigenic clay formation) — and those

01-Vernik_Ch01.indd 16 12-08-2016 21:04:02


Chapter 1: Petrophysics of Siliciclastic Rocks 17
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 20.  The same data as in Figure 17, here with the
Figure 21.  Porosity-versus-permeability plot for fine-grained
superposed models using clay volume as a proxy for mean
arenites (with a mean grain size in the range of 125 to
grain size.
250 µm), showing the effect of zeolite mineralization. The
model line superposed is the same as that in Figure 20. Even
origins may affect the clay mineralogy and structural posi- minor zeolite presence is equivalent to doubling the clay
tion. Finally, the presence of other minerals that have little content, in terms of a permeability reduction of nearly an
effect on porosity but that greatly reduce permeability, order of magnitude.
such as zeolites, can result in dramatic overprediction.
The influence of zeolites is illustrated in Figure 21, The permeability model, when properly calibrated
which shows two subsets of upper-fine-grained arenites and applied, is capable of markedly improving the overall
with and without noticeable zeolite mineralization. quality of petrophysical modeling. Besides helping with
Permeability is reduced by almost an order of magnitude in net-pay estimation and reservoir zonation, the parameter
the cores that have a significant amount of zeolite. That ( k /φ )1 2 (also known as the flow-zone indicator, or FZI) is
reduction is not accompanied by variations in grain size or related to the reservoir quality and can be computed as a
clay content, so using equation 26 can grossly overestimate curve and used to constrain resistivity-based water-satu-
permeability. The situation can be predictable if a strong ration calculations. However, the problem of water satu-
influx of volcanogenic material is established in the depo- ration has been extensively discussed in the petrophysical
sitional center. Volcaniclastic fragments can be much less literature and lies outside the scope of the book.
stable than quartz and even feldspars, so such fragments Figure 22 illustrates an example of log modeling to
tend to transform diagenetically into zeolites under the estimate lithology, porosity, and permeability in a deep-
right pressure and temperature conditions (Iijima, 1980). water turbidite reservoir from the Gulf of Mexico. Given
It is necessary to emphasize that the model presented sample bias, heterogeneity, and differing vertical resolu-
here is designed with routine, commonly available log in- tions between the log and core, the porosity and permea-
puts in mind. However, uncertainty can be reduced by de- bility curves match the core measurements closely in this
veloping a careful log model for Vcl, because even small reservoir. At the same time, one can also attempt to use
variations in clay content can result in marked variations downscaling to estimate sand porosity and sand permea-
of permeability. Also, special care should be used when bility curves in the intervals where Vsh exceeds zero.
dealing with thin beds and coarse-grained sands and sand-
stones. Thin beds impose limitations in terms of the log
resolution but can be handled by incorporating image Seismic petrophysics
logs. The presence of mixed-grain, moderately to poorly
sorted sand (e.g., Figure 3b), and pebbly sand can be Contemporary petrophysical analysis, which includes
identified by using porosity logs or from petrographic ob- wireline and “logging while drilling” (LWD) log editing,
servations on cores and cuttings via mud-log analysis. modeling, and reservoir-parameter calculations, becomes
The presence of such lithologies can account for the infre- a truly integrated discipline when combined with standard
quently encountered reversal from a positive correlation and special core analyses. Indeed, when log-model-based
to a negative one between porosity and permeability. computations of clay/shale content, porosity, permeability,

01-Vernik_Ch01.indd 17 12-08-2016 21:04:03


18 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 22.  Log-model interval, showing Vsh, lithology, resistivity, porosity, and permeability compared with the core data.
Channel facies are shown with a light-yellow dotted legend and a Vsh value of zero. PHITsd and PERMsd (red) curves are the
sand-component porosity and permeability, derived by using arithmetic/harmonic modeling via Vsh.

and saturation are complemented and calibrated with qual- defines seismic petrophysics as “purposeful application
ity core data, the results are always more robust and useful of rock-physics theory, as calibrated by laboratory and
in reservoir description. well measurements, to the interpretation of seismic data.”
As a next step, integration of rock physics into petro- Smith (2011) emphasizes the role of petrophysics in
physics allows one to purposefully relate elastic rock the integrated subject of seismic petrophysics. He specifi-
properties (bulk density ρ b, VP, and VS), obtained from cally points out that petrophysics should aim at (1) log
wireline and LWD logs, to reservoir parameters. Ulti­ editing and conditioning, (2) shear-wave velocity predic-
mately, this allows us to use synthetic forward modeling tion, (3) fluid substitution, and (4) porosity modeling. It is
to gain a better understanding of seismic amplitudes in logical to combine Pennington’s and Smith’s definitions
both poststack and prestack domains. and complement them with seismic forward modeling,
In this manner, the science of rock physics, which is which is also an integral part of seismic petrophysics. The
conventionally designed to study the interrelationships primary reason for that combination is that a practitioner
among the physical and petrographic rock properties by should be able to understand the petrophysical pitfalls and
using modeling and experimentation, can be incorporated properly apply rock-physics models and crossplots to
into a broader subject of seismic petrophysics. To avoid check quality and condition the log data. In addition, he/
confusion with existing usage of the terms rock physics, she also should be able to appreciate the impact of those
petrophysics, and the like, Pennington (1997, p. 243) pitfalls, models, and plots on synthetic modeling and

01-Vernik_Ch01.indd 18 12-08-2016 21:04:04


Chapter 1: Petrophysics of Siliciclastic Rocks 19

seismic attribute analysis, especially on amplitude-­ An additional step required for proper application of
variation-with-offset (AVO) interpretation. Often, rock- seismic petrophysics techniques is gather conditioning,
physics modeling allows us to reveal errors in basic petro- including multiple suppression and noise suppression, re-
physical interpretation, which may have adverse effects sidual normal moveout (NMO) correction, and occasion-
on calculations of rock-frame properties. ally spectral balancing, to counteract the NMO stretch at
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

A typical workflow in reservoir characterization, in- far offsets. Synthetic gathers from either well logs or
cluding its critical elements and building blocks, is shown rock-physics-model-based pseudo-wells should be used
in Figure 23. Seismic petrophysics, when defined in the as a guide and reference at this stage. Therefore, synthetic
way suggested above, acts as an interactive link between AVO modeling often becomes a focal integration point,
geologic interpretation and seismic inversion, in the pro- when seismic can help us evaluate and control the quality
cess frequently referred to as quantitative reservoir de- of the log data, and vice versa.
scription, or quantitative interpretation (QI). Combined Even though geologic modeling is separated from
with geologic modeling and interpretation, the QI process quantitative reservoir description in the scheme illustrat-
should be part of the integrated reservoir-characterization ed in Figure 23, it is a critical part of the overall reservoir-
process leading to reservoir simulation (Figure 23). characterization activity. Geologic modeling uses concep-
The scheme in Figure 23 also applies in exploration, tual depositional systems and their subdivision into
but the lower two levels (reservoir characterization and ­lithofacies, sequence-stratigraphic units, and architectural
reservoir simulation) normally are omitted in the explo- elements, and that practice is substantiated by using pet-
ration process. In the absence of adequate well control, rophysical analysis and its upscaling (e.g., Slatt et al.,
seismic inversion for absolute impedance volumes is 2009).
rarely an option and the reservoir description can be re- A typical seismic petrophysics workflow can be sum-
duced to a seismic petrophysics investigation followed marized as follows:
by attribute analysis, geologic modeling, and relative im-
pedance inversion. • quality control of standard petrophysics analysis for
The roles of proper data acquisition and especially of lithology, porosity, permeability, and saturation, with
amplitude-friendly processing are hard to overestimate in additional data conditioning if necessary
any exploration activities, but they are particularly impor- • petrophysical parameterization necessary in rock-
tant in activities focusing on amplitude plays and pros- physics model building (Vcl, Vsh, porosity, saturation)
pects. For instance, anisotropic prestack depth migration • editing of the data on elastic rock properties (bulk
is now a pervasive processing routine that allows com- density ρ b, VP, and VS), including bad-hole and mud-
mon-depth-point (or image-point) gathers to be accurate- filtrate invasion effects, anisotropic corrections in
ly colocated in 3D space using well logs. wells with large relative dips, and VS log calibration
and/or computation
• evaluation of temperature, pore pressure, and at least
Seismic Petrophysics vertical principal stress of the stress tensor
acquisition/processing (log/core analyses)
• application of the lithology-controlled rock-physics
templates, thereby incorporating seismic and reser-
voir properties, and purposeful investigation of the
main diagenetic and local stratigraphic trends
Reservoir description Geologic • AVO behavior prediction using rock-physics
(seismic petrophysics, inversion) modeling templates
• forward model building with the best set of elastic-
rock-property logs, along with comparative analysis
with conditioned gathers around the well location and
Reservoir characterization away from well control
• AVO attribute analysis and prestack simultaneous in-
version of seismic data in terms of impedances, ve-
locity ratio, or any other desirable attributes.

Reservoir simulation The last item, of course, can be a significant activity on


its own and can be considered complementary to the
Figure 23.  Schematic of a reservoir-characterization and ­seismic petrophysics workflow, or vice versa. In fact,
reservoir-simulation workflow. The upper two levels are also some modern software packages for quantitative seis-
applicable in exploration-oriented geoscience workflows. mic interpretation incorporate seismic-petrophysics and

01-Vernik_Ch01.indd 19 12-08-2016 21:04:04


20 Seismic Petrophysics in Quantitative Interpretation

seismic-inversion processes that are intimately interrelat- sands and sandstones. The end-member velocities of sand
ed, thus allowing the necessary calibration loops and and shale can then be combined using the Vsh curve as a
quality controls. weighting factor.
Even though the modified Faust equation 27 is empiri-
cal, it merits some physical basis — the vertical velocity
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Log editing and horizontal resistivity of shale, as measured by respec-


Careful editing and conditioning of well-log data are tive logging tools in vertical wells, are both strong functions
two of the first steps in the seismic petrophysics workflow. of shale porosity, especially in compacting mudstones from
Even moderately inaccurate information about the elastic Tertiary and Cretaceous basins. For that reason, both of
rock properties in situ can generate significant errors in those types of logs are successfully used in pore-pressure
rock physics and, hence, in synthetic seismic models later prediction. For best results, equation 27 should be calibrat-
on, during the quantitative seismic analysis (Smith, 2011). ed using available resistivity and sonic logs and then ap-
Modern software packages offer a range of solutions plied to the missing (typically shallower) log intervals.
that are based largely on empirical equations relating den- It is frequently maintained that sonic-log readings in
sity and/or shear-wave velocity to compressional velocity, gas-saturated rocks are not reliable (e.g., Chopra and
with lithology-specific coefficients (e.g., Gardner et al., Castagna, 2014). In all fairness, this notion was quite ac-
1974; Castagna et al., 1993). The empirical models are curate 20 years ago; however, due to marked changes in
justified in this early stage of data editing because many sonic-log processing over the last two decades, the reli-
other parameters necessary in advanced rock-physics ability of the compressional traveltime log has dramati-
modeling are not yet determined. In this section we will cally improved, even in gas sands (e.g., Cheng, 2015).
discuss some of those models, along with mud-invasion Moreover, the velocity obtained from sonic traveltimes
correction and correction of sonic logs in wells with large can always be tested by the appropriate rock-physics
relative dip angles. models to evaluate possible effects from mud invasion,
heterogeneity, or frequency dispersion.

Sonic log
Density log
Because continuous GR and resistivity logs common-
ly are available from LWD measurements, they can be Density-log conditioning
used to edit or compute sonic velocity in shales. Shale VP Density-log editing in “bad-hole” conditions (e.g.,
values can be calculated most accurately by using a slight- washouts, breakouts, excessive rugosity) is critical and
ly modified version of the Faust equation (Faust, 1951): has been adequately addressed in the petrophysical litera-
ture (e.g., Smith, 2011). The best way to condition the
 1
VP = 0.3048 b ( zml Rsh ) 4  , (27) density log is to replace suspicious intervals with com-
  puted density. The same applies to the shear-velocity log,
which will be discussed in Chapter 3.
where b is a unitless adjustable parameter in the 850–1000 The density log is one of the most critical pieces of
range, zml is vertical depth in feet below the mud line seismic petrophysics information that can be obtained
([zml = TVD – WD – KB], with TVD being true vertical from a well. However, being a pad device, it is very sensi-
depth, WD being water depth, and KB being the kelly tive to poor wellbore conditions, such as washouts and
bushing elevation, all in feet), Rsh is shale resistivity in rugosity. Moreover, unlike GR, resistivity, and even sonic
ohm-meters, and 0.3048 is the conversion factor from feet logs, the density tool is rarely run over extensive inter-
per second to meters per second. Care should be exercised vals, so computing density is sometimes a must.
when using this formula in settings in which a significant Castagna et al. (1993) present a useful compilation of
difference in the effective stress and, therefore, in com- density-velocity relationships, with lithology-specific co-
paction profiles is suspected between the calibrated and efficients that are derived using a popular functional form
edited data sets. given by Gardner et al. (1974),
It is important to realize that equation 27 should be
applied only to shales (i.e., to lithologic intervals with ρ b = aV Pb , (28)

Vsh = 1, using the definition described earlier in this chap-
ter). That application is helpful because shales comprise where VP is in kilometers per second (km/s), a is 1.75 for
the vast majority of the overburden rocks in any siliciclas- shales and 1.66 for sandstones, and superscript b = 0.265
tic sedimentary basin. Rock-physics models discussed in and 0.261 for shales and sandstones, respectively. This em-
Chapter 3 should be used to compute sonic velocities in pirical equation, with a = 1.741 and b = 0.25, also can be

01-Vernik_Ch01.indd 20 12-08-2016 21:04:05


Chapter 1: Petrophysics of Siliciclastic Rocks 21

used to represent an average transform if no information on a)


lithology is available. Density is given in grams per cubic
centimeter and VP in kilometers per second ­everywhere in
this section. As an alternative, Castagna et al. (1985) present
polynomial-form density transforms for sand and shale.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

The equation lines shown in Figure 24a were calcu-


lated using equation 28. It is interesting that the sand-mod-
el line runs through the middle of sands and s­ andstones of
all petrophysical groups, including wackes, from the com-
piled database, but that quite a large density overestima-
tion (0.1 to 0.15 g/cm3) is possible in some subsets — no-
tably in clean arenites and medium-porosity arenites.
To the contrary, the shale line in Figure 24a dramati-
cally underestimates density in shales, except in those
with reduced clay content. The latter were likely preva-
lent in the subset used by Castagna et al. (1993), who
claim that this modified Gardner’s shale transform per-
forms reasonably well. Therefore, although Gardner’s
b)
middle line separates shales fairly accurately from the
other petropysical groups (wackes, arenites, and clean ar-
enites), using it or even its modified version to predict
density distribution in shale-dominated sedimentary ba-
sins is likely to result in unacceptable errors.
Bowers (2001) finds that Gardner’s equation 28 for
shale can serve as the lower bound on density, which is
consistent with Figure 24a. In addition, he proposes an
equation that represents an upper bound:

ρ b = 1.3 + 1.03(V P − 1.46)0.28 . (29)



Model lines illustrated in Figure 24b for shales are given
by a generalized and simplified version of this formula:

ρ b = ρ0 + (V P − 1.5)0.35 , (30)
Figure 24.  Water-saturated bulk density versus VP: (a) a
where ρ0 is the density of the mud (suspension) at the database comprising low-frequency core data at 20 MPa
mud line in offshore applications. It varies from 1.16 to (from Gassmann’s equation modeling) and two well logs
1.40 g/cm3, and VP is in the range from suspension veloc- superposed by the shale, sand, and average models of
ity (1.5 km/s) to the bedding-normal matrix velocity value Gardner et al. (1974); and (b) proposed models for shales
in shales (4.0 to 4.5 km/s; see Table 2). Figure 24b and (four realizations shown as blue lines), soft sands (two
realizations shown as yellow dotted lines with pore shape
Figure 25, which includes massive shales from 15 deep
factors of 15.0 and 25.0, see Chapter 3), consolidated arenites
wells worldwide, support this model and demonstrate that
(red line), and consolidated clean arenites (brown line), the
its applicability range is greater than that of Gardner’s
last two of which are at a 20-MPa-stress reference.
model — almost down to the velocity and density values
for shale of 2.0 km/s and 2.0 g/cm3, respectively, corre- substantial stress unloading is underway, such as in high-
sponding to very poorly compacted muds. ly geopressured intervals, the density-velocity relation-
There are three different controls on ρ0 in equation ship can be quite different from that given by equation 30,
30. The first is quite intuitive: the clay content in the because of the dramatic velocity reduction in high-density
shale. The second is apparently related to the geothermal shales (Bowers, 1995). In Figure 25 this effect is observed
gradient and, hence, to clay diagenesis and mineralogy. as a noticeable bulge of the high-probability cloud to the
Finally, the third may be due to mild shale unloading that left; that is, as a shift toward lower VP values at the same
is accompanied by velocity reversal without simultaneous density of approximately 2.5 g/cm3.
density reversal, indicative of shales that are greatly over- Two consolidated sandstone lines are added to the tem-
pressured due to stress unloading. In zones where plate in Figures 24b and 25 for reference. Their functional

01-Vernik_Ch01.indd 21 12-08-2016 21:04:06


22 Seismic Petrophysics in Quantitative Interpretation

2.80 consolidation level. A departure of data points from this


Vsh = 1 Arenite line toward the shale models in Figure 24b can, in fact, be
indicative of overpressure-related stress reduction in con-
2.60 solidated sandstones, the degree of sand consolidation, or
changing composition of the reservoir sands.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Clean Interestingly, such a departure toward shale models is


ρ b (g/cm3)

2.40 arenite
often the case in poorly consolidated sequences under low
effective stress (<10 MPa) when sands and shales super-
2.20 pose in VP-versus-bulk-density space and occupy the
lower end of the distribution shown in Figure 25. The rea-
sons for all these o­ bservations are discussed and account-
2.00 ed for in Chapter 3, on more theoretical grounds.

1.80
Mud-invasion correction
1.5 2.5 3.5 4.5 5.5 6.5
Because of the relatively shallow depth of investiga-
VP (km/s) tion, with 90% of the signal coming from the 5-in (13-cm)
annulus around the borehole, density-tool readings suffer
1 132 from the mud-filtrate-invasion effect. The best way to
Color: Frequency
minimize that effect is to run a log model code that is built
Figure 25.  Twenty-five wells from a worldwide data set, filtered to include fluid-density contamination, as was discussed
to include only data points with Vsh = 1 (massive shales) and earlier. As a result, the total-porosity curve is generated
plotted on the P-wave velocity-density template of Figure 24. and can be combined with the water-saturation curve to
recompute the true formation density:
forms are motivated by the following simple empirical con-
struct for brine-saturated velocity (e.g., Issler, 1992): ρ = ρm (1 − φ ) + ρf φ ,
(33)
ρ f = ρw S w + ρhc (1 − S w ),
VPb = Vm (1 − φ ) f , (31)

where VPb is the brine-saturated P-wave velocity, Vm is the where ρ hc is the hydrocarbon density. Because water satu-
velocity of the matrix, φ is the total porosity, and f is the ration Sw is calculated on the basis of the deep-­reading re-
empirical exponent that may be related to clay content sistivity logs, it should be relatively immune from the inva-
when applied to conventional shales. This equation also sion effects. Notice that unlike the case in equation 20, the
can be qualitatively applied to brine-saturated consoli- fluid density in equation 33 is the true fluid density in situ.
dated sandstones: in combination with ρ b = f (φ ), it re- It is obvious that the magnitude of the density correc-
sults in brine-saturated sandstone density ρ bb: tion for mud invasion depends on the total porosity and
the fluid-density contrast between the mud filtrate and the
1 formation; the largest invasion effect is predictable in gas
V  f (32) reservoirs drilled with a WBM system. The accuracy of
ρ bb = ρf + ( ρm − ρf )  Pb  ,
 Vm  the density correction is based on the accuracy of the to-
tal-porosity estimation, which, in turn, depends on the
where the exponent f can be selected to be 1.45 for aren- invasion profile. As is often the case in the absence of
ites and 1.55 for clean arenites, and where ρf and ρm are core data, several iterations with equations 16, 20, and 33
the fluid and matrix densities, respectively. It is notewor- are required to converge on the final, most accurate densi-
thy that these lithified sandstone transforms are not ap- ty-log restoration to the in situ saturation condition.
plicable to unconsolidated or poorly consolidated high- A most dramatic manifestation of the mud-filtrate-
porosity arenites and clean arenites, which in Figure 24b invasion effect on a density log is given in Figure 26 for
are shown to have P-wave velocities lower than 3.0– a gas-sand reservoir. Due to significant water invasion
3.5 km/s. Soft, poorly consolidated sand model realized into the annulus probed by the density tool (Sinv = 0.6 was
for pore shape factors of 15.0 and 25.0 and included in obtained by iterations), a density reading approximately
Figure 24b will be explained in Chapter 3. 0.15 g/cm3 higher than the actual in situ density of this
Note also that the empirical equation 32 should only gas sand was observed on the density log. Needless to
be used with caution and preferably as a reference, be- say, even a much less dramatic distortion can impact the
cause it fails to account for either effective stress or sand impedance, the shale-over-sand interface reflection

01-Vernik_Ch01.indd 22 12-08-2016 21:04:08


Chapter 1: Petrophysics of Siliciclastic Rocks 23
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 26.  Bulk-density correction for a WBM-filtrate invasion into a gas-saturated sand reservoir in the North Sea.
Sinv = 0.75, ρg = 0.08 g/cm3, and ρw = 1.03 g/cm3. Note the dramatic effect on the density-log (RHOB, black) reading between
the fully wet curve (RHOBB, blue) and the gas-saturated sand density in situ (RHOBG). The sonic log (VP, black) is reading
a gas-saturated velocity (VPg, red) obtained from fluid-substitution modeling.

coefficient, and, hence, the seismic amplitudes generated indistinguishable from the original sonic-log curve. If, on
by the synthetic forward modeling. the other hand, either the sand reservoir itself or its satu-
It should be mentioned that the effect of mud-filtrate ration is not homogeneous in the 50-cm annulus around
invasion on VP is basically negligible in this case, which is the well, some mud-invasion impact on compressional
a symptomatic case in general. The depth of investigation velocity measured by the sonic device can be expected
(DOI) of modern sonic tools, given by the dominant and predicted by using VP-VS templates (discussed later)
wavelength of the received signal, typically is greater in combination with fluid-substitution modeling with
than the size of the annulus of the most significant mud Gassmann’s equation (e.g., Smith, 2011).
invasion into the borehole wall. In general, depending on
the frequency and the formation velocity, the DOI of the Anisotropic correction of sonic logs
sonic tool can be at least three times that of the density
tool — that is, on the order of 45 to 50 cm. Sonic velocities in wells with large relative dips pen-
Referring to Figure 12, the invaded-mud-filtrate satu- etrating anisotropic formations are known to be non­
ration Sinv affecting the sonic log is typically much lower unique because, in addition to the well-documented ef-
than the Sinv value that is visible to the density log. If the fects of lithology, porosity, and stress, they are also af-
sand reservoir is more or less homogeneous and, due to fected by the relative dip angle. The anisotropic effects
various fluid-mixing mechanisms, a homogeneous satura- are related to the intrinsic anisotropy of shales and thin-bed
tion in the effective annulus is set, then on the basis of laminations of sand, silt, and shale (e.g., Vernik et al., 2002;
Gassmann’s equation only a minor increase in VP can be Hornby et al., 2003).
predicted. Indeed, as is shown in Figure 26, the gas-satu- Seismic-depth ties, low-frequency models for seismic
rated velocity corrected for mud invasion is almost inversion, and pore-pressure prediction in these cases are

01-Vernik_Ch01.indd 23 12-08-2016 21:04:08


24 Seismic Petrophysics in Quantitative Interpretation

complicated and require some “verticalization” of the ve- where Cij are elements of the composite stiffness tensor,
locities prior to their use. Additionally, standard seismic and angles are measured from the symmetry axis
forward modeling and inversion applications require C11 = ρVP (90°)2 , C33 = ρVP (0°)2 , C44 = ρVS (0°)2 , C 66 =
VP(0°) curves, which can be viewed as being verticalized ρV SH (90°)2, ρ is the bulk density, and the triangular brack-
velocities if the formation dip angle is relatively mild ets denote averaging of local layer properties of sand and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

(say, less than 10°). shale end members by using Vsh and (1 – Vsh) terms. Note
It is common knowledge that most sedimentary rocks also that both local and effective elastic stiffnesses c13 and
are anisotropic (Thomsen, 1986; Vernik and Liu, 1997; C13 can be expressed via the bedding-normal stiffnesses
Wang, 2002a, 2002b); the degree of anisotropy can vary and the anisotropy parameter δ (Thomsen, 1986):
from quite low (less than 5%) in sandstones and carbon-
ates to more than 30% in organic or clay-rich low-porosi- C13 = (C 33 − C 44 )2 + 2 δ C 33 (C 33 − C 44 ) − C 44 . (35)

ty shales. Sonic logs that are acquired in deviated wells,
or, more generally, in wells with large relative dip angles, For the two model calculations presented below, it is as-
and that penetrate the same anisotropic formation at the sumed that the sand layers (laminae) are quasi-isotropic,
same vertical depth range, will always measure velocities so that c11 = c33, c44 = c66, ε = δ = γ = 0, c13 = c33 – 2c44,
which differ from those in straight holes. This situation is and hence their velocity values are VP(0°) = VP(90°) and
most frequently encountered in deepwater exploration VS(0°) = VSH(90°).
and development, when several deviated appraisal wells Permeable sands may exhibit significant velocity
may have to be drilled from the same platform. dispersion (Winkler, 1983), whereas the dispersive ef-
In this section a methodology is presented to “numer- fects on shale can be negligible. Local sand velocities
ically rotate” sonic logs in wells with large relative dips should be derived either from sonic logs sampled at a Vsh
by using approximated-phase-velocity functions of the of zero, or from dry core measurements subjected to the
dip angle; such functions can be computed from the well desired fluid saturation from Gassmann modeling (to be
deviation and the formation dip angles. A continuous ve- discussed later). Because calculated Vsh values and veloc-
locity-log correction will be demonstrated to be a func- ity logs both have similar vertical resolutions of 0.6 to
tion of the formation shale content Vsh on the log scale, as 1.0 m, the Backus averaging is internally consistent.
defined above. The methodology largely concerns layer- Nonetheless, uncertainty remains, due to the assump-
ing-induced anisotropy and intrinsic anisotropy of shale tions that
and shale-sand laminations — that is, transversely isotro-
pic (TI) formations with a bedding-normal symmetry • sand/sandstone and shale properties in thin, unre-
axis. This kind of anisotropy is also known as polar an- solved laminae (where 1 > Vsh > 0) are similar to
isotropy. Stress-induced anisotropy in sands is not consid- those measured in thicker, homogeneous sand layers
ered but can be incorporated. (with Vsh = 0) and massive shales (with Vsh = 1); and
Let us assume that the anisotropy parameters ε, δ, and • sand/shale laminations have an average thickness that
γ (Thomsen, 1986) of shale and sand, as well as other is less than approximately 0.1 m, to satisfy the effec-
local elastic rock properties of the two-phase medium, tive-medium constraint.
such as VP, VS, and density, are known and vary little over
a limited depth range of 500 m in deepwater turbidites. The second assumption will result in predictable
The Backus averaging technique (Backus, 1962) can be e­rrors when average thicknesses are in the 0.1- to 1-m
employed to compute the effective density (denoted by range — that is, when individual layers cannot be resolved
ρ*) and the effective elastic constants (denoted by capital by either GR or sonic logs and their composite properties
Cs in the following system of equations): cannot be accurately computed using equations 34.
The three anisotropy parameters of the effective
−1
media on the log scale can be computed as follows
−1
C 33 = c33 , (Thomsen, 1986):
−1 2
2 −1 −1 −1
C11 = c11 − c13c33 + c33 c13c33 ,
C11 − C33
−1
−1
−1 ε = ,
C13 = c33 c13c33 , (34) 2 C33
−1
−1 C66 − C44
C 44 = c 44 , γ = , (36)
2 C44
C 66 = c66 , and (C13 + C44 )2 − (C33 − C44 )2
* δ = .
ρ = ρ , 2 C33 (C33 − C44 )

01-Vernik_Ch01.indd 24 12-08-2016 21:04:09


Chapter 1: Petrophysics of Siliciclastic Rocks 25

Table 3.  Rock properties of sands/sandstones and shale a)


used in anisotropy modeling.
Rock properties Shale Sand 1 Sand 2
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Vsh 1.00 0 0
Porosity 0.17 0.25 0.30
Saturating fluid brine brine oil
Bulk density ρb (g/cm )
3 2.45 2.25 2.12
VP (0°) (km/s) 2.95 3.25 2.60
VS (0°) (km/s) 1.45 1.75 1.45
ε 0.12 0 0
δ 0.07 0 0
γ 0.15 0 0

b)
The functional dependence of the effective anisotropy pa-
rameters on Vsh is implicit in equations 36 because each of
the effective elastic constants computed using equations
34 depends on Vsh.
Both of the examples of the anisotropic effective-me-
dium computations described here share the same shale
properties but different sand properties (Table 3). Note
that the anisotropic properties of shale used in the models
come from specific core measurements and may not be
representative in general. The rest of the required param-
eters can be derived from sonic and density logs. By using
equations 34, the effective elastic constants and bulk den-
sity are computed first, followed by the effective anisot-
ropy parameters estimation from equations 36.
The results for the two cases are presented in Figure
27. It is noteworthy that the shapes of the ε(Vsh ), δ (Vsh ), Figure 27.  Anisotropy parameters versus Vsh in (a) softer
and γ (Vsh ) composite functions, which are determined by sand and (b) stiffer sand. Bulk density is given in grams per
the relative stiffnesses and rigidities of the thin layers, are cubic cemtimeter and velocities are in kilometers per second.
different. Specifically, the anisotropy parameter lines are Adapted from Figure 2 of Vernik (2008). Used by permission.
concave downward for the softer sand/sandstone (Figure
27a), whereas they are concave upward for the softer
group velocities in anisotropic media can differ signifi-
shale (that is stiffer sand/sandstone) layers (Figure 27b).
cantly (Thomsen, 1986; Vernik and Liu, 1997).
Moreover, there may exist more exotic layer-property
For weakly anisotropic rock, Thomsen (1986) de-
combinations, not captured by these two cases, wherein
rived simplified approximations to the quasi-P-wave (qP-
the maximum value of ε or δ can occur at Vsh values
wave) phase velocity surface, and to an exact SH-wave
smaller than one (Melia and Carlson, 1984). The proce-
velocity surface, as follows:
dure outlined in equations 34 and 36 to estimate depen-
dence of the anisotropy parameters on shale volume Vsh is VP (θ ) = VP (0 )(1 + δ sin 2 θ cos 2 θ + ε sin 4 θ ),
a necessary element of any anisotropic correction to sonic (37)
logs. VSH (θ ) = VS (0 )(1 + 2γ sin 2 θ ),
Full-wave processing of dipole sonic logs by using
slowness time coherence was recently demonstrated to where θ is the relative dip angle. The maximum error
yield anisotropic phase velocities rather than group ve- when one is using the approximation for qP-wave anisot-
locities of compressional VP and shear VS waves (Sinha ropies less than 20% (ε < 0.2 ) does not exceed 2% for
et al., 2006). This finding is crucial to the problem dis- relative dips of less than 60° and can be neglected. The
cussed here, because the angle dependence of phase and angle θ in these formulas is measured with respect to

01-Vernik_Ch01.indd 25 12-08-2016 21:04:11


26 Seismic Petrophysics in Quantitative Interpretation

a) be computed from the following generalized equation


(Moran and Gianzero, 1979):

θ rad = sin(α )cos( β )sin(ϕ )cos(ω )


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

+ sin(α )sin( β )sin(ϕ )sin(ω )


(38)
− cos((α )cos(ϕ ),
θ = a cos(θ rad ),

where θ r ad is the relative dip angle in radians, α is the bed


dip, β is the bed dip direction (azimuth), ϕ is the well
deviation from vertical, and ω is the well deviation azi-
muth (all in radians).
Given the well-deviation survey data and the struc-
tural dip angle and azimuth, the relative dip can be com-
puted easily from equation 38 and then used to invert for
b) the rotated sonic velocities VP(0°) and VS(0°) from equa-
tions 37. In the relatively flat subsurface with a vertical
symmetry axis, these rotated velocities should be equal to
those measured in vertical wells, assuming that the nature
of the geology has not changed significantly among the
wells.
Figure 29 shows an example of anisotropy correction
of sonic logs for a highly deviated deepwater Gulf of
Mexico well interval that contains an oil reservoir with
encasing shales. It should be pointed out that continuous
log corrections are only as accurate as the Vsh-dependent
anisotropy parameters given in equations 36. Obviously,
sand and shale properties and anisotropy all vary with
depth, making it difficult to compute accurate ε(Vsh ),
δ (Vsh ), and γ (Vsh ) functions over significant depth inter-
vals (e.g., greater than 500 m). Therefore, it is advisable
Figure 28.  (a) Phase-velocity values versus relative dip
to use log segmentation (blocking) followed by block-by-
angle in (a) softer sand and (b) stiffer sand. Adapted from
block velocity correction that uses constant rock-property
Figure 3 of Vernik (2008). Used by permission.
inputs within each block.
The case illustrated in Figure 29 shows isotropic sand
beds and an anisotropic, massive shale (Vsh = 1), along
the symmetry axis of the TI formation. Figure 28 shows with sand/shale laminations (1 > Vsh > 0) that are typical
the behavior of qP- and SH-wave velocities, as a function of overbank levee facies in deepwater turbidites. The end-
of θ, for the massive shale with a Vsh of 100%, and for the member sand and shale properties are similar to those
laminated composite rock with a Vsh of 50%, based on the presented for the first model in Table 3. The amount of
model examples presented above. Note that for the softer correction depends on the effective anisotropic parame-
sand/sandstone case the P-wave-velocity surfaces at ters obtained from Vsh. The relative correction of shear
Vsh = 1 and Vsh = 0.5 are distinctly separated from each velocity in this field case is greater than that for VP be-
other (Figure 28a), whereas for the stiffer sand/shale case cause, on the basis of the core measurements, γ > ε .
they gradually approach and cross over at an angle of 67° The least-constrained input data necessary for
(Figure 28b). anisotropic correction are the shale anisotropy param-
The angle θ generally can be referred to as the rela- eters and their variation with depth. However, by wir-
tive dip angle in subsurface logging applications, mean- ing the set of equations discussed in this chapter into a
ing that it is the angle between the wellbore trajectory and dedicated s­ oftware program, it is possible to iteratively
the direction normal to the formation’s bedding plane. If adjust shale-anisotropy parameters to minimize the
the formation’s dip angle is zero, the relative dip angle θ correction errors when one is comparing the result to
is reduced to the well’s deviation angle. Otherwise, it can the sonic velocities recorded in nearby vertical well(s)

01-Vernik_Ch01.indd 26 12-08-2016 21:04:12


Chapter 1: Petrophysics of Siliciclastic Rocks 27
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 29.  Log plot from a well with a large relative dip, showing an anisotropy correction (in red) carried out for deep-
resistivity (not discussed in this book), VP, and VS logs. Lithology color-coding: yellow is channel-sand facies, green is levee
facies, gray is shale, and purple is siderite-rich shale.

or the interval velocities from vertical seismic profiling Shumaker and Vernik, 2008) and thus are comparable
(VSP) or checkshot surveys. Anisotropy parameters of to some organic-rich shales (discussed in Chapter 6),
conventional shales vary with clay content, clay whereas most of the Tertiary Gulf of Mexico and Gulf
­m ineralogy, and the clay’s preferred orientation. Shales Coast shales with moderate clay content are character-
with high clay content offshore equatorial West Africa ized by ε values smaller than 0.15 to 0.20 (Vernik et al.,
can have ε values of 0.25 to 0.30 (Toldi et al., 1999; 2002).

01-Vernik_Ch01.indd 27 12-08-2016 21:04:13


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 2: Pore Pressure and Stress State

This chapter is important for our topic because (1) of effective-stress computation and constraints on the
stress state and pore pressure, often viewed as subjects of stress tensor in the subsurface. Because effective stress is
geomechanics, are also principal considerations for seis- a primary control on elastic rock properties, the effec-
mic petrophysics, and (2) rock-property dependence on tive-stress concept is used throughout the rest of this
effective stress has been documented in multiple publica- book.
tions and is common knowledge. Of course, these sub-
jects remain mostly geomechanical in nature (Zoback,
2009), but their impact is enormous in our understanding Shale-compaction model
of rock properties and, hence, of the controls on seismic
amplitudes in the subsurface (Sayers, 2010). It is impera- Porosity reduction
tive for the integrated teams of geoscientists and reservoir
Shale compaction in Tertiary and Cretaceous sedi-
engineers to address pore pressure and stress state early in
mentary basins worldwide lies at the foundation of stress
seismic exploration or reservoir characterization process
and pore-pressure estimations in the subsurface. Indeed,
to avoid potential pitfalls down the road.
mechanical compaction of clay-rich mud, which often
Formation fluid pressure in excess of hydrostatic
comprises as much as 80% of the siliciclastic sequences,
pressure has been routinely observed in deep hydrocar-
is known to be a function of vertical effective stress.
bon-exploration wells worldwide. Overpressure develop-
Vertical effective stress is often approximated by the dif-
ment in clastics-dominated sedimentary basins is most
ference between the overburden stress and the pore pres-
often attributed to shale-compaction disequilibrium —
sure at any given point in the subsurface:
that is, to an undercompaction mechanism caused by in-
hibited dewatering of shales (Sharp and Domenico, 1976;
σ v = Sv − PP , (1)
Bowers, 1995; Hart et al., 1995).
Other sources of overpressure, such as clay diagene-
where σv is the vertical effective stress, Sv is the overbur-
sis, organic-matter maturation, and thermal expansion are
den stress, and PP is the pore pressure (Terzaghi and Peck,
invoked locally and are referred to as unloading mecha-
1948; Skempton, 1969; Aplin et al., 1995).
nisms (e.g., Hart et al., 1995; Dutta, 2002). Numerical pre-
Aplin et al. (1995) derived the following relationship
dictions of unloading-related overpressure meet with vary-
between porosity and vertical effective stress:
ing success due to the nonuniqueness of the sonic-veloci-
ty-versus-effective-stress relationship during unloading.
This chapter presents basic concepts that are helpful  e − e
σ v = σ v0 exp  0 , (2)
in constraining variations in stress and pore pressure in  τ 
sedimentary basins. In turn, pore-pressure variation can
be used in sophisticated rock-physics models that incor- where e = φ/(1 - φ) is the void ratio, τ is the compression
porate effective stress and, subsequently, in synthetic for- coefficient, and σv0 is the reference stress value that cor-
ward-model building for quantitative seismic analysis. responds to the initial void ratio e0 considered in the
First, shale-compaction models are discussed, and model. This model has been supported by multiple uni-
that discussion leads to development of a conceptual den- axial strain tests in soil mechanics, but it may have limit-
sity model that can be complementary to velocity/density ed applicability for shale compaction at depths exceeding
transforms (discussed in Chapter 1) for computing over- a few hundred meters.
burden stress. Next, sonic-velocity-based pore-pressure Rubey and Hubbert (1959) suggested that at greater
prediction in shales is presented, followed by discussions depths, the dependence of porosity on effective stress is

29

02-Vernik_Ch02.indd 29 12-08-2016 20:03:01


30 Seismic Petrophysics in Quantitative Interpretation

exponential, rather than logarithmic as is implied by a)


equation 2:

φ = φo exp( −cσ v ), (3)


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

where φo is the mud porosity at a relatively low stress and/


or a shallow reference depth generally greater than the ini-
tial stress (and depth) used in equation 2. To prevent any-
one from mislabeling the parameter c as shale compress-
ibility (Flemings et al., 2002), and noting that shale com-
paction is essentially a ductile process, c can be replaced
with the inelastic compaction modulus Cm (Vernik, 2011):

 σ 
φ = φo exp  − v  , (4)
 Cm 

with Cm typically being in the range from 24 to 31 MPa


(3500 to 4500 psi) and averaging 27.5 MPa (4000 psi) for b)
normally compacting sediments below sea bottom (mud
line), whereas φo is normally in the 40 to 43% range.
Higher-porosity muds are very soft sediments that barely
support shear stress, so the model quickly becomes inad-
equate at very shallow depths of burial.
An alternative relationship between porosity and ver-
tical effective stress is suggested in the following power-
law expression of Sayers (2010):

n
 φ (5)
σ v = σ vmax 1 − φ  ,
c

where φc is the critical porosity above which the vertical


effective stress is assumed to vanish, and σvmax is the effec-
tive stress corresponding to zero porosity. It is interesting
that with a σvmax value of 107 MPa, a φc value of 43%, an n Figure 1.  Shale porosity as a function of vertical effective
value of 1.9, and a σv value of less than 70 MPa, Sayers’s stress, from well data and models on (a) smaller and (b) larger
power-law model in equation 5 basically coincides with the depth scales. The well data in (b) are from formation-pressure
porosity/effective-stress relationship implied by another tests in sands and mean porosity in encasing shales. Four
popular power-law model, developed by Traugott (1997), exponential-compaction-model realizations (red), with φo
that relates bulk density of the siliciclastic sediments to the reduced to 40%, are superposed in (b). Adapted from Figures
true vertical depth below the mud line (zml = TVDml). The 1 and 2 of Vernik (2011). Used by permission.
Traugott model, which is discussed later, implies a porosity
of 43% at the mud line — a value consistent with the criti- the exponential compaction model at σv values of 3 to
cal porosity used to evaluate equation 5. 8 MPa (400 to 1200 psi), which may correspond to the tran-
Figure 1a shows the vertical-effective-stress-versus- sition from soft muds to more-consolidated mudstones.
porosity models as given by equations 2, 4, and 5. Figure 1a also reveals a significant difference between
Respectively, the Aplin shallow mud-compaction line the exponential compaction model of Rubey and Hubert
(equation 2) is computed with σv0 = 0.1 MPa, φo = 0.55, (1959) and the power-law models of Traugott (1997) and
and τ = 0.17; the exponential compaction model of equa- Sayers (2010), in the entire stress range, with the crossover
tions 3 and 4 is realized with φo = 0.43 and Cm = 27.5 MPa; at approximately 9 MPa. The difference is even more pro-
whereas the power-law model of equation 5 is evaluated nounced at the greater effective-stress values included in
using the parameters indicated above. In other words, the Figure 1b. Notice that the compaction modulus Cm in the
last two models are made to share the same mud-line po- exponential model may have to be slightly reduced, from
rosity for better comparison. With this parameterization, 27.5 MPa to 24 MPa for some undercompacted mudstones,
the shallow mud-compaction line seamlessly merges with to match the data, which were acquired at greater depths

02-Vernik_Ch02.indd 30 12-08-2016 20:03:03


Chapter 2: Pore Pressure and Stress State 31

and/or higher overpressure and which will be further de- accounted for by equation 4 with compaction modulus
scribed below. Such a reduction in Cm results in more sepa- values greater than approximately 20 MPa.
ration between the two types of models. Consequently, on the basis of the modeling results
A part of this discrepancy can be attributed to the fact shown in Figure 1, the more optimal normal-hydrostatic-
that the Traugott power-law model was developed from compaction-trend (NCT) model should be a combination
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

density logs of siliciclastic sequences, including sands, of equations 2 and 4 spliced around 5 MPa. However, if
and without taking the pore-pressure variation into ac- one is less interested in geotechnical applications and
count. The abundance of those sands is not quantified in more concerned with greater current or burial depths,
the model, nor is their contribution to the average density adding the Aplin model is not necessary because the
values. By contrast, the exponential model, with Cm in the ­errors associated with omitting it can be neglected.
24.0- to 27.5-MPa range, is designed for normally pres-
sured and undercompacted shales and was demonstrated
to be quite accurate at intermediate to greater depths
Density model for stress computation
(Flemings et al., 2002). Even though it has been mentioned in some publica-
Well data shown in Figure 1b were derived from 35 tions that the mean effective stress is a better control on
formation-pressure measurements in sands/sandstones shale porosity than its vertical component alone is, it is
and complemented by the log-model-porosity averages safe to say that the errors from using σv as a proxy for the
over 15- to 30-m intervals of encasing shales in the depth mean effective stress should be minimal, at least in slight-
range from 1.1 to 7.1 km TVDml. The pore-pressure gradi- ly compacted sediments in relaxed sedimentary basins.
ents in this data set range from 11.7 to 20.2 kPa/m (0.52 The main reason for this is that in extensional-tectonics
to 0.90 psi/ft); that is, from slight overpressure to geo- basins, at depths below mud line that are greater than
pressure approaching the least principal stress. 3000 m and are of interest for hydrocarbon exploration,
Following Bowers (1995) and Traugott (1997), extra the minimum horizontal effective stress σhmin commonly
care was exercised in the data-set selection to avoid ex- is within 15 to 20% of the effective overburden stress σv.
tensive sand bodies with significant structural relief and Moreover, as will be demonstrated later in this chap-
the associated possibility of lateral pressure transfer (the ter, this ­effective-stress anisotropy decreases with depth.
so-called “centroid” effect; Traugott, 1997). This effect Indeed, there are worldwide observations of stress an-
often results in a mismatch between shale pressure and isotropy reduction at greater depths related to overpres-
sand pressure. However, minor manifestations of the cen- sure generation in combination with the condition of fric-
troid effect could not be entirely ruled out. tional equilibrium on pre-existing faults optimally ori-
The pore-pressure and shale-porosity/velocity data ented with respect to the maximum stress in situ (Zoback
set included in Figure 1b will be interrogated throughout and Healy, 1984). Note also that uniaxial compaction
this chapter to distinguish between the most important tests with zero lateral strain, which are designed to model
mechanisms generating pore pressure, such as compac- the overburden loading, typically are performed in soil
tion disequilibrium and stress unloading. These mecha- mechanics (Wood, 1990).
nisms are anticipated to have different effects on seismic To compute Sv or the overburden gradient OBG
rock properties. (OBG = SvG) as a function of depth, an accurate forma-
In general the majority of well-data points in tion-bulk-density log is required that extends all the way
Figure 1b, regardless of the pore pressure, are consistent to the surface. As mentioned above, the model by Traugott
with the exponential compaction model in which Cm is in (1997) has been widely used to restore inaccurate or miss-
the range from 21.0 to 27.5 MPa. It is also evident that the ing density-log intervals (e.g., Sayers, 2010):
data do not match the power-law compaction model with
b
any logical input parameters, such as those described ρ ( zml ) = ρo + a zml , (6)

above.
Several noticeable outliers in this plot are quite where zml = TVDml and ρo is the mud density close to the
meaningful: Those outliers suggest that the compaction- mud line. Not surprisingly, Traugott’s power-law expres-
dominated stress-porosity relationship can break down in sion turns out to be equivalent to another power-law ex-
zones subject to stress unloading that is typical of areas pression — that of Sayers, in equation 5 — in the stress
with (1) an elevated temperature gradient, (2) a greater range from 0 to 70 MPa (10,000 psi), when the model
rate of clay-mineral-transformation reactions (with mini- parameters are ρo = 1.96 g/cm3, a = 1.53e − 3, and b = 
mal impact on shale porosity), and (3) structural uplift/ 0.6, and porosity is derived from φ = (ρm − ρ)/(ρm − ρf) by
erosion processes. It should be emphasized that any out- using the values ρm = 2.65 g/cm3 and ρf = 1.05 g/cm3
liers identified in effective stress versus porosity space (Sayers, 2010).
are indicative of pore-pressure-generation mechanisms As was previously mentioned, the better way to deter-
other than compaction disequilibrium, if they cannot be mine the density model, including the density NCT, is to

02-Vernik_Ch02.indd 31 12-08-2016 20:03:03


32 Seismic Petrophysics in Quantitative Interpretation

take advantage of equation 4 if small errors due to shal- 0


low mud effects can be neglected. From the linear poros- Cm = 27.5 MPa
ity-density dependence and equation 4, we have
5000
 σ z 
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ρ ( z ml ) = ρm − ( ρm − ρf ) φo exp  − vG ml  , (7)
 Cm
10,000

TVDml (ft)
where σvG = OBG – PPG and is the vertical-effective-stress
gradient. Unfortunately, in reality both the effective-stress
15,000
gradient σvG and the shale matrix density ρm are also depth-
dependent functions, as a result of gradual compaction and
a decrease in the smectite interlayer ratio in illite/smectite
20,000 NCT proposed
clays. An empirical model that attempts to incorporate these
NCT Traugott
processes for shale bulk densities greater than 2.0 g/cm3 is Well data
suggested in the following (Vernik and Kachanov, 2010): 25,000
1.6 1.8 2 2.2 2.4 2.6 2.8
ρ = ρmmax + 0.001{( ρ − 2) − 230 exp[ −4( ρ − 2)] , (8)
m
} Bulk density (g/cm3)

Figure 2. Bulk-density-versus-TVDml models and log averages


where ρmmax is an asymptotic matrix-density value that from the formation-pressure data set. Note that the Aplin
should be selected from the 2.73- to 2.79-g/cm3 range and (shallow mud compaction) model was omitted at shallow
should average 2.76 g/cm3 (see Chapter 1). This equation, depths. The density model proposed (equation 9) is treated as
as already mentioned in Chapter 1, can be solved itera- an NCT density expression for a Cm value of 27.5MPa.
tively with φ = (ρm − ρ)/(ρm − ρf) to derive the bulk-densi- Adapted from Figure 3 of Vernik (2011). Used by permission.
ty/depth model that is in agreement with equation 4.
Alternatively, the process of replacing vertical the ef-
which was not designed to differentiate between normally
fective stress with the depth below mudline can be greatly
pressured and substantially overpressured sediment densi-
simplified by introducing the following approximation
ties. Therefore, the bulk-density model of the compacting
(Vernik, 2011):
mudstone proposed here may be a viable alternative, at
least for the density NCT: It is more consistent with the
 λz  formation-pressure measurements in situ.
ρ ( zml ) = ρmax − exp  − ml  , (9)
 Cm  Finally, if a reliable bulk-density log is available below

a certain depth that is above the onset of overpressure, it
which is consistent with equations 7 and 8 and also fits can be spliced with the prediction from equation 9. The
the iterative solution for the density NCT fairly well when calculation should be adjusted by varying the parameters
the empirical parameter λ = 13.3 MPa/km (0.59 psi/ft), ρmax, λ, and Cm to meet the density log at the merger depth.
ρmax = 2.76 g/cm3, and C = 27.5 MPa.
Figure 2 compares this hydrostatically compacting
shale bulk density (i.e., the density NCT) with the one Vertical effective stress
computed from Traugott’s (1997) power-law model (or
The total vertical principal stress Sv is the foundation
the similar model of Sayers, 2010). In light of the results
for stress-state and pore-pressure analyses in a well, sedi-
shown in Figure 1, the substantial difference between the
mentary structure, or basin. Stress is a second-rank tensor,
shale-density NCT model derived from equation 9 and the
with Sv being one of the three principal stresses in the ten-
power-law-based model seen in Figure 2 is to be expect-
sor. Another two components are Shmin and SHmax — the min-
ed. Note also that the density model as presented in Figure
imum and maximum horizontal total principal stresses.
2 ignores the shallow mud-compaction curve (from the
Commonly it is assumed that the vertical stress is the
Aplin model). However, the latter does have to be consid-
easiest to calculate. Indeed, if we know the bulk-density
ered in geotechnical applications, such as in stress esti-
variation with depth, then a simple integration of the bulk
mates at depths smaller than 1000 ft (300 m).
density from the surface to the depth of interest yields the
Figure 2 also indicates that measurements of shale
vertical principal stress, which is often also referred to as
density below 18,000 ft TVDml (5500 m) are quite accu-
the overburden stress:
rately matched by the Traugott model. Given the fact that
all sand/shale couples from the well-data set are substan- z

tially overpressured at these depths, this match simply em-


phasizes the empirical nature of the Traugott expression,
Sv = ∫ ρ gdz ,
b
(10)
0

02-Vernik_Ch02.indd 32 12-08-2016 20:03:04


Chapter 2: Pore Pressure and Stress State 33

where z is the true vertical depth (TVD) and g is the grav-


ity constant. Working in the metric system with depth in
meters, density in 103 kg/m3 = 1 g/cm3, and g = 9.81 m/s2,
one obtains the stress value in pascals, which is more con-
veniently expressed in megapascals (MPa). Note also that
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Sv (MPa) ≈ 145Sv (psi), in case one prefers to use the


English system with stress expressed in pounds per square
inch (pounds/inch2 or psi).
Of course, the main uncertainty in equation 10 is the
variation in bulk density with depth. As was discussed in
Chapter 1, density logs are not routinely run in well sec-
tions above the main target interval. Sonic logs, on the
other hand, are more frequently acquired in those shal-
lower sections, notably when no check shot or vertical
seismic profile (VSP) acquisition is planned.
A solid workflow in this situation is to build and run a
petrophysical model, including Vcl and Vsh computations,
followed by a density estimation from the sonic velocity
by using equations 30 and 32 of Chapter 1. Generation of Figure 3.  Normal vertical-effective-stress gradient as a
function of TVDml, computed using density equation 9 and
the regional and local trend plots for velocities is quite
assuming a normal PpG value of 10.4 kPa/m (0.46 psi/ft).
helpful prior to this calculation. Additionally, the modified
Adapted from Figure 5 of Vernik (2011). Used by permission.
Faust equation (equation 27 of Chapter 1) can be consid-
ered to compute shale velocity from resistivity. It should reduction and dewatering. A thick salt body, encountered in
also be kept in mind that the density NCT realized from well B, pinches out at the location of well A.
equation 9 should always be the upper bound for any re- For simplicity, the siliciclastic section is assumed to
sultant density profile. These steps naturally lead to the be 100% shale that is undergoing compaction, with den-
well-constrained, Vsh-weighted density-log construction sity increasing according to equation 9. The water d­ ensity
that is to be merged with the actual density-log curve. is assumed to be constant at 1.06 g/cm3, and the salt-body
Integrating the merged density curve to obtain Sv and density is 2.15 g/cm3. The resultant density and total-ver-
then subtracting the pore pressure results in an effective-­ tical-stress-gradient profiles (OBG = Sv/z in kPa/m) are
vertical-stress or effective-overburden-gradient profile shown in Figure 5. For reference, an ­interval with con-
(Terzaghi and Peck, 1948). A hypothetical profile, shown in stant density of 1.0 g/cm3 corresponds to a stress gradient
Figure 3, was computed by integration over depth of the den- of 8.33 lb/g = 0.433 psi/ft ≈ 10 kPa/m.
sity NCT model from the appropriately parameterized equa-
tion 9, with the assumption of a normal pore-pressure gradi-
ent PpG of 0.46 psi/ft (ρw = 1.06 g/cm3). The highly nonlinear
effective-overburden-gradient model is the direct conse-
quence of the exponential normal-compaction curve shown
in Figure 2. In deepwater applications, one will have to use
available industrial software to compute water density as a
function of salinity, pressure, and temperature. Obviously,
the deeper the water is, the more significant is the effect of
the water column on the overburden calculation. Exotic li-
thologies in predominantly sand/shale sequences — such as
salt, anhydrite, carbonates, coal, and the like — are intended
to be taken into account by creating constant-density layers
corresponding to those lithologies (see Table 1 of Chapter 1),
if they are not covered by density logs.
Ignoring all these pitfalls may result in noticeable er-
rors in computing the vertical stress; those errors in turn
propagate into the estimates of effective stress and pore
pressure. Figure 4 presents a schematic cross section be- Figure 4.  Schematic geologic section showing two
tween two pseudowells (A and B) drilled in different water pseudowells, A and B, drilled to 18,000 ft (5.5 km). The two
depths in a sedimentary basin that is predominantly filled wells have markedly different profiles of density and
with shale and is undergoing compaction due to porosity overburden stress (see Figure 5).

02-Vernik_Ch02.indd 33 12-08-2016 20:03:04


34 Seismic Petrophysics in Quantitative Interpretation

• pore-pressure prediction that is based on this unified


DEPTH DEPTH
RHOB1 OBG1 concept is robust and has been verified in multiple
g/cm3 kPA/m
(FEET) (METERS) 0 3 10 25
projects.
RHOB2 OBG2
0 g/cm3 3 10 kPA/m 25
1600.0
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

1000
Sonic velocity versus effective stress in shales
5000 As Vernik and Kachanov (2010) recently demonstrat-
2000
ed, a heuristic model can be created that relates bedding-
normal velocity and porosity in shales preconsolidated to
porosity values of 40 to 45%. This model (1) incorporates
10,000 3000 the key lithologic variable, clay content, and (2) is consis-
tent with the gradual compaction caused by the increase
in effective stress:
4000

15,000 C33m (1 − φ )k *
VP (0 ) = , (11)
5000 ρm (1 − φ ) + ρf φ

19,400.0 5913.1
where C33m is the bedding-normal elastic stiffness of the
shale matrix, k* is the clay-content-dependent empirical
Figure 5.  Density profiles and overburden gradients (OBG), exponent k* = 5.3 − 1.3vcl, φ is the total porosity, and ρm
in kPa/m, for pseudowells A (red) and B (blue), which are and ρf are the matrix and fluid densities, respectively.
shown in Figure 4. This model assumes bimodal shale composition with
At the 5.5-km depth in well A, Sv is 98 MPa, whereas in aligned/laminated clay platelets and nonclay components.
well B it is 114 MPa. Normal (hydrostatic) pore pressure at Therefore, the simple Reuss averaging calculation of the
this depth is 57 MPa, so the normal vertical effective stress bedding-normal matrix stiffness on the solid-rock basis is
is 40 MPa in well A and more than 40% greater (57 MPa) in an adequate option:
well B. The difference could be even more dramatic if a re-
−1
alistic overpressure scenario is considered. For instance, if  v 1 − vcl 
C33 m =  cl + , (12)
the pore-pressure gradient is 15 kPa/m (0.65 psi/ft), which  c33 cl M ncl 

is the median value in the well data set, then PP is 81 MPa
and the effective stresses are 17 MPa (2450 psi) and 33 MPa where the subscripts m, cl, and ncl stand for the mineral-
(4800 psi) in wells A and B, respectively. It is quite obvious matrix, clay, and nonclay (predominantly quartz) compo-
that a difference of this magnitude may have a dramatic ef- nents of shale and vcl denotes the clay content on the sol-
fect not only on prediction of pore pressure but also on the id-mineral basis. The overall performance of this model,
entire exploration or reservoir-characterization project. when compared with the extensive database of sonic and
density logs worldwide, is elaborated upon in Chapter 3.
However, it is important to note here that this model
Pore-pressure prediction from works best with the shale grain mineral density implicitly
given by equation 8.
shale velocity When shale porosity does not exceed 40 to 45%, the
Despite the importance of understanding the origin of coupled equations 11 and 12 can be adequately replaced
anomalously high pore-pressure gradients, relatively few with a simplified empirical alternative:
publications explore the relationships among shale poros-
ity, sonic velocity, and effective stress. Moreover, several V P (0 ) = V Pm (1 − φ ) k ,
schemes and algorithms for pore-pressure prediction lack V Pm = a − b v cl + c v cl2 , (13)
a common principle and rely on contradictory models for
these property relationships. As a result, such schemes k = d − f v cl .

may yield unstable pore-pressure predictions, notably
when made without local calibration (Hart et al., 1995; In these equations, a = 5.69 km/s, b = 3.56, c = 1.42,
Katahara, 2003). The intent here is to show that d = 2.302, and f = 0.646 (all unitless). In the dynamic
range of porosities from 0 to 40%, this model produces
• such a common principle can be found, practically the same VP-versus-porosity model lines as
• various schemes of pore-pressure prediction can be those presented by equations 11 and 12. It is also consis-
organically unified, and tent with a similar power-law model that is used in Issler

02-Vernik_Ch02.indd 34 12-08-2016 20:03:05


Chapter 2: Pore Pressure and Stress State 35

et al. (1989). As shown in Figure 6 of Chapter 1, the latter


basically misses most of the core and log measurements
of velocity due to overprediction. A more detailed analy-
sis of our shale-velocity model is provided in Chapter 3.
By substituting equation 4 into equations 13, we ob-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tain the expression for bedding-normal sonic velocity ver-


sus vertical effective stress, the expression that is consis-
tent with the exponential shale-compaction model:
k
  σ 
VP (0 ) = VPm 1 − φo exp  − v   . (14)
  Cm  

This formula, with an “initial” porosity φo of 40% and a
Cm value of 27.5 MPa (4000 psi) as default values, can be
treated as the velocity NCT, if the normal hydrostatic
pore-pressure gradient is plugged in for the vertical-effec-
Figure 6.  Results from the shale VP-versus-porosity model tive-stress derivation (equation 1). As is the case with the
of equations 13 with the well-data set shown in Figures 1b density NCT, presented above, this velocity model ignores
and 2. Almost no core data are available in the well-data set, unconsolidated mud compaction, so the initial porosity in
so the petrophysical classification must be applied to qualify equation 14 describes a mudstone that has been precon-
all log-data intervals adjacent to the formation-pressure-test solidated. A result of this preconsolidation is porosity re-
measurements as being a massive homogeneous shale with duction, from values of 55 to 80% immediately below the
Vsh = 1. mud line to porosities of 40 to 45% several hundred me-
ters below the mud line.
(1992), but this alternative is more robust than Issler’s However, in general, equation 14 presents an explicit
model because it takes clay content into account. relationship between bedding-normal P-wave velocity
Figure 6 illustrates the performance of this VP-versus- and effective stress in mudrocks, a relationship that can
porosity model against the same shale data set used in be applied in pore pressure prediction and velocity model
Figures 1b and 2, which includes in situ pressure tests and building. Note also that clay content in this velocity-ver-
vertical-effective-stress calculations. Whereas the majority sus-effective-stress model is accounted for by the matrix
of data points corresponding to the homogeneous shale in- velocity term VPm and the exponent k, as given in equa-
tervals align themselves with the model lines realized for a tions 13. An additional flexibility of this model is pro-
wide range of clay contents (vcl values from 40 to 80%), vided by the compaction modulus Cm that can vary from
there are several low-velocity outliers. 15 to 30 MPa depending on the relative contributions of
Interestingly, these outliers in Figure 6 belong to a sub- the d­isequilibrium compaction- and unloading-related
set of well intervals in which the dominant pore-pressure pore pressure generation mechanisms.
mechanism changed from differential compaction to unload- Figure 7 compares this model, computed for a vcl of
ing, likely as a result of thermal expansion or accelerated 50%, with the various power-law models (those of Bowers
clay-mineral-transformation reactions. It must be empha- [1995] and Sayers [2010]). Again, quite predictably, the re-
sized that the shale porosity is barely affected, whereas the sults of the exponential-law-based model (of Vernik [2011])
velocity is markedly reduced during such unloading. Those are strikingly different from the results of the power-law
changes in shale properties have been traditionally attributed models, including those using the straightforward Bowers
to microstructural alterations in their pore space, including equation VP = VPo + Aσ vB , in which A = 1.5 km/s and
microcrack extension (Bowers, 1995). Three out of five very B = 0.68. Note also that the relationships of Traugott (1997;
noticeable outliers come from the Gulf Coast province, not shown in Figure 7) and of Sayers (2010) are virtually
USA, and two come from West Africa. identical and do not explicitly allow velocity-model com-
Unfortunately, absolute clay content in the data set putation; these relationships must be combined with some
plotted in Figure 6 could not be adequately calibrated be- kind of density/velocity transform, thereby making the pro-
cause of the lack of core data. However, the majority of cess more cumbersome and perhaps less accurate.
the samples are aligned near the vcl = 0.6 model line, Figure 8 plots sonic P-wave velocity versus vertical
which is deemed to be representative of average shales effective stress, as given by equation 14, using a vcl value
worldwide (Figure 5 of Chapter 1). When using clay con- of 50%, and for a broad range of compaction moduli Cm
tent range from 40 to 60%, as in Figure 6, our velocity values from 17.0 to 31.0 MPa. Effective-stress estimates
model is seen to much more accurately predict the effect from the formation-pressure tests are compared with the
of shale porosity than the model of Eberhard-Phillips sonic-log-based pressure estimates from adjacent shales.

02-Vernik_Ch02.indd 35 12-08-2016 20:03:06


36 Seismic Petrophysics in Quantitative Interpretation

the unloading process, which is not accompanied by any


noticeable porosity or density changes but results in sig-
nificant bedding-normal velocity reduction associated
with the opening of bedding-subparallel microcracks.
Similarly to the effective-stress-versus-shale-porosity
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

plot in Figure 1b, Figure 8 reveals that the majority of


data points lie between the model lines with Cm values of
20.5 and 27.5 MPa. Also in a similar manner, some veloc-
ity/effective-stress measurements in situ fall out to the
left, along the trajectories defined by the experimental
measurements of VP = f(σ). These clear outliers cannot be
accounted for by using the VP(σ) models with a compac-
tion modulus greater than 21 MPa. The gentler experi-
mental unloading curves naturally follow the apparent
unloading of shales that reached an irreversible porosity
reduction during loading history and compaction.
Figure 7.  Bedding-normal compressional-wave velocity models Clearly, the most logical explanation for the outliers in
as a function of vertical effective stress. Note the difference Figure 8 is an unloading-related reduction in effective
between the power-law models (of Bowers and of Sayers) and stress, rather than disequilibrium compaction, as the domi-
the exponential-law model (of equation 14 and Vernik). The nant pore-pressure-generating mechanism. A similar con-
exponential-law model was realized for vcl = 50% and three clusion was reached above regarding the porosity-versus-
different compaction moduli Cm values: 24.0, 27.5, and effective-stress data and models presented in Figure 1b.
31.0 MPa, plotted here from top to bottom, respectively (red). Therefore, both plots are complementary to each other and
Adapted from Figure 7 of Vernik (2011). Used by permission. quite instructive in establishing the primary mechanism for
generating pore pressure (Bowers, 1995). However, most
importantly, the two plots provide clues to a possible unifi-
cation of pore-pressure-prediction formalism that is solely
based on the more stable exponential compaction model.

Pore-pressure prediction
The vertical effective stress in shales with variable
clay content can be derived from equation 14 if an under-
compaction mechanism (i.e., compaction disequilibrium
between the rate of loading and the rate of dewatering in
low-permeability shales) is expected:

 
 φo 
σ v = C m ln  1 , (15)
1 − (V P /V Pm ) k 
 
Figure 8.  Bedding-normal compressional velocity in shales
versus vertical-effective-stress models, as given by equation whereas the vertical principal stress Sv can be computed by
14, superposed by (1) a data set of reservoir-pressure-test- the merged model/log-density integration over depth. The
based σv and adjacent shale VP log averages (blue data difference between them, by definition, amounts to the
points), and (2) VP-versus-stress measurements on three shale pore pressure (Terzaghi and Peck, 1948). The clay
select shale cores (black curves) that can be viewed as content in shale enters the σv computation through the pa-
unloading curves from the three different levels of rameters VPm and k as defined by equations 13. Consequently,
compaction (24.0, 27.5, and 31.0 MPa, as seen in Figure 7). accurate estimation of clay content in shale, which typi-
Adapted from Figure 8 of Vernik (2011). Used by permission. cally is derived from either the GR log or analysis of neu-
tron and density data, is vital in pore-pressure prediction.
Finally, also superposed on the plot are three velocity- Even though the compaction modulus Cm value of
stress curves (black curves) obtained from laboratory ex- 27 ± 3 MPa is attributed to normal shale compaction and
periments on shale cores taken from various levels of undercompaction, reduced values of Cm can be used to
shale compaction. Those curves are designed to display achieve better accuracy in pore-pressure estimates, when

02-Vernik_Ch02.indd 36 12-08-2016 20:03:08


Chapter 2: Pore Pressure and Stress State 37

unloading scenarios are suspected in localized depth in- less stable model when the unloading mechanism is iden-
tervals or below a certain depth. For instance, if a tem- tified (Bowers, 1995; Katahara, 2003).
perature effect on fluid expansion or clay diagenesis is It should also be emphasized that the pore-pressure
anticipated, the effective compaction modulus Cme can be estimation from equation 15 is much more stable over in-
a function of temperature: tervals greater than 1000 m, compared with calculations
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

from yet another power-law model introduced by Eaton


C me = C m − b (T − Tc ), (16) (1975):
3
where T and Tc are the current and critical temperatures, V 
respectively. The coefficient b (0.18 to 0.31 MPa/°C) is σ v = σ vn  P  , (17)
 VPn 
the rate of unloading that is set after the critical-­
temperature characteristic of the onset of unloading is where σvn is the vertical effective stress calculated by
reached (e.g., 130°C). No matter which compaction-mod- using normal hydrostatic pore pressure, and VPn is the ve-
ulus-perturbation model is used, the pore-pressure predic- locity NCT.
tion is performed by using the same unified-model Figure 9 displays a geomechanical model, including
­expression 15, instead of by introducing an alternative, pore pressure, computed with the unified approach

FG Vp
8 PPG 20 0.5 KM/S 5.5
PpG upscaled Vp NCT
8 PPG 20 0.5 KM/S 5.5
GR DEPTH TVD PpG RHOB Log
30 GAPI280 (FEET) (METERS) 8 PPG 20 0 G/CM3 3
GR Resistivity MW RHOB Model
280GAPI 30 0.1 OHMM1000 8 PPG 20 0 G/cc 3
0.0 0.0
ϕo = 0.40
Cm = 27.5 MPa

5xxx

2xxx RHOB

10xxx
S VG

4xxx MW
Vp
15xxx

PpG

6xxx
20xxx

FG

Figure 9.  Pore-pressure, fracture, and overburden gradients estimated in a deepwater well, Gulf of Mexico, showing the entire
section (top) and the target interval of 4500–7400 km (base). Note the mud weight (blue) providing an upper bound to the
upscaled pore-pressure-gradient curve (orange), which is consistent with downhole pressure measurements (red dots). The
curve for the minimum-horizontal-principal-stress (or fracture) gradient ShminG ≈ FG was calibrated using data from lithology-
dependent stress coupling and LOTs in shale. All data and computations are in pounds per gallon (ppg). Adapted from Figure 9
of Vernik (2011). Used by permission.

02-Vernik_Ch02.indd 37 12-08-2016 20:03:10


38 Seismic Petrophysics in Quantitative Interpretation

FG Vp
8 PPG 20 0.5 KM/S 5.5
PpG upscaled Vp NCT
8 PPG 20 0.5 KM/S 5.5
GRn DEPTH TVD PpG RHOB Log
30 GAPI280 (FEET) (METERS) 8 PPG 20 0 G/CM3 3
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

GRn Resistivity MW RHOB Model


280GAPI 30 0.1 OHMM1000 8 PPG 20 0 G/cc 3

SVG
15xxx

5xxx

6xxx
20xxx

7xxx

Figure 9. (Continued)

(equations 15 and 16) for a deep Gulf of Mexico well Quite a different case is presented in Figure 10.
drilled in very deep water. Note that the calculation of the This well was drilled in the area of an elevated geother-
overburden gradient is also based on the exponential com- mal gradient of 42°C/km on the ocean shelf, offshore
paction model for density (equation 9) merged with the West Africa. The well is highly deviated, so the sonic
density log. Because sonic and model-based clay-content velocity had to be corrected for anisotropy (see Chapter
logs were both available, a continuous-pressure-gradient 1), which generally is quite significant in this province.
profile was built on a meter-by-meter basis using, by de- The shale pore-pressure-gradient profile was computed
fault, a Cm value of 27.5 MPa and a φo value of 40% as the in the TVD interval of 1.8 to 4.0 km, by using the
model parameters. The shale pressure profile PpG can be ­conditioned sonic velocity and the same initial combi-
upscaled (i.e., smoothed; orange curve in the figure) as nation of model parameters (Cm = 27.5 MPa and
needed and compared with the reservoir-pressure mea- φo = 0.4) as in the deepwater Gulf of Mexico example.
surements and the mud-weight gradient. The onset of unloading occurs at the depth of approxi-
Figure 9 shows an excellent match among these vari- mately 3 km.
ous data sources. The geothermal gradient in the target A good agreement is evident between the shale pore-
interval of this well is 27°C/km, so, assuming we have a pressure estimates and the sandstone pressure measure-
critical temperature of 130°C for the onset of unloading, ments in the reservoirs encountered at depths from 3.7 to
and the process has a rather low rate, with b being 3.9 km. The effective stress of approximately 14 MPa is
0.2 MPa/°C, only a slight decrease in the compaction computed for reservoirs and encasing shales. The effec-
modulus can be expected below the onset-of-unloading tive compaction modulus Cme at this gross reservoir in-
depth of 6.7 km. At approximately this depth, the mud terval, which is included in the data set shown in Figure
weight had to be increased by 1 ppg (pound per gallon) to 8, is approximately 15.5 MPa (see equation 16, with
safely drill the last several hundred meters of the well. Tc = 130°C, T = 170°C, and b = 0.3). This value of Cme

02-Vernik_Ch02.indd 38 12-08-2016 20:03:12


Chapter 2: Pore Pressure and Stress State 39

FG Vp
8 PPG 20 0.5 KM/S 5.5
PpG upscaled Vp NCT
8 PPG 20 0.5 KM/S 5.5
GR TVD TVD PpG RHOB LOg
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

30 GAPI280 (FEET) (METERS) 8 PPG 20 0 G/CM3 3


GR Resistivity MW RHOB Model
280GAPI 30 0.1 OHMM1000 8 PPG 20 0 G/cc 3
0.0 0.0
φo = 0.40
Cm = 27.5 MPa

1xxx S vG

FG
5xxx
PpG
MW

2xxx

10xxx 3xxx

Figure 10.  Pore-pressure, fracture, and overburden gradients estimated in a well drilled in shallow water, offshore
West Africa. The mud weight (blue) again provides an upper bound, here on the smoothed pore-pressure-gradient curve
(orange). The minimum-horizontal-stress-gradient ShminG curve was calibrated using data from lithology-dependent stress
coupling and LOTs in shale.

is consistent with the equivalent undercompaction mod- overlying this reservoir was penetrated by using markedly
els presented in Figure 8, thus implying a strong contri- overbalanced mud weight, with almost no consequences
bution from the unloading mechanism to pore-pressure for drilling operations.
generation. One should always keep in mind that the main pitfall
In contrast, the estimated shale pressure gradient in of the techniques for estimating shale-velocity-based (or
the 3.0- to 3.2-km-depth range is larger than that tested in resistivity-based) pore pressure and effective stress is in
thin reservoir sandstones, which turned out to be almost the possibility of inadequate equilibration between per-
normally pressured and subjected to a σv value of 36 MPa. meable reservoir rocks and very-low-permeability shales
The reduction in the effective stress at the 3.3-km-deep encasing them. It has been well documented that the
pore-pressure ramp — from 36 to 14 MPa — is quite dra- lateral-pressure-transfer effect can be bipolar; that is,
matic, and it obviously is expected to manifest itself in there is always a possibility that this mechanism either
elastic velocities, impedances, and velocity ratio. differentially drains sand bodies through faults during
The relatively thin sandstones encountered in the 3.0- tectonically active periods (as in Figure 10), or overpres-
to 3.2-km depth interval were later interpreted to be in surizes tilted sand beds with respect to encasing shale as
lateral communication with a much thicker reservoir, a result of pressure pumping from their extension down-
which experiences a significant pressure-gradient regres- dip (the so-called centroid effect suggested by Traugott
sion due to the hydraulic connection with shallower inter- (1997).
vals via permeable faults. It is also noteworthy that the Nonetheless, the situation exemplified in Figure 9,
1500-m-thick, virtually impermeable shale interval where the full equilibration of pore pressure across sand/

02-Vernik_Ch02.indd 39 12-08-2016 20:03:13


40 Seismic Petrophysics in Quantitative Interpretation

shale boundaries is established, is more the rule than an where β is the Biot parameter given by
exception (Bowers, 1995; Traugott, 1997). Therefore, the
shale-pressure-based methodologies for pore-pressure Kd
β =1− , (22)
prediction and effective-stress estimation may be ade- Km

quate in the majority of cases, at least until disqualified
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

by formation-pressure test results in permeable reservoirs with Kd being the drained bulk modulus, which is equal to
that are subjected to both lateral and vertical pressure the dry frame modulus of the rock, and Km being the solid
transfers. matrix bulk modulus. Accurate measurements of the Biot
parameter are challenging, but for most cases of silici-
clastic rocks that are of interest in conventional hydrocar-
bon exploration high- to moderate-porosity shales, it is
The effective-stress tensor and assumed that β does not deviate significantly from unity.
the Shmin gradient This means Terzaghi’s law should hold. A possible excep-
tion to that rule of thumb is with the domain of very-low-
Because the effective-stress state in the earth is the porosity mudrocks under a very high stress state, with Kd
second-rank tensor σkl (in the Einstein summation con- approaching Km, so that β → 0.
vention), Hooke’s law can be written as An argument frequently is made (e.g., Eaton and
Eaton, 1997; Katahara, 2003) that the uniaxial strain con-
ε ij = S ijkl σ kl , (18)
dition is fulfilled in relaxed sedimentary basins more
often than not, even though the stress coupling could be
where εij and Sijkl are the second-rank strain tensor and the poroelastic or even inelastic. In the inelastic case, the
fourth-rank elastic-compliance tensor, respectively. The stress coupling is commonly estimated with more empiri-
second-rank stress tensor σkl can be written in terms of the cally based values of pseudo-Poisson’s ratio ν*. Indeed,
three orthogonal stresses σ1, σ2, and σ3 as its diagonal much larger values of K0 than those derived from the true
matrix elements, with σ1 > σ2 > σ3. If one of these three is Poisson’s ratio (either static or dynamic) typically are re-
vertical, then by definition the other two are horizontal. In quired to match hydraulic fracturing data and leak-off
geomechanics it is useful to describe the stress state with tests (LOTs) that have been acquired at depths greater
σhmin, σHmax, and σv, which, using Terzaghi’s principle, are than 1.0 to 1.5 km in sedimentary basins undergoing
compaction.
σ v = S v − Pp , From equations 19 and 20, we can obtain the mini-
σ hmin = S hmin − Pp , (19) mum total-horizontal-stress gradient ShminG:
σ Hmax = S Hmax − Pp . S hminG = K 0 (S vG − PpG ) + PpG . (23)

For an isotropic elastic solid medium, the pore-pres- (Note that G in the subscripts signifies “gradient.”) This
sure term vanishes and the effective-stress tensor becomes stress measure is alternatively referred to as the forma-
the total-stress tensor. Thus, when it is uniaxially loaded tion-fracture gradient. The stress-coupling factor is forced
in the vertical direction with zero lateral strain, this solid to be depth-dependent or shale-compaction-dependent,
medium develops coupled horizontal stresses given by because it is computed by using the functional form of
equation 20 but replacing the Poisson’s ratio with an em-
σ h = σ H = K 0σ v , pirical pseudo-Poisson’s ratio v* (Eaton and Eaton, 1997;
(20) deepwater Gulf of Mexico, TVDml > 1500 m)
υ
K0 = ,
1−υ ν * = 0.426 + 7.295e − 6 zml − 1.882e − 10 zml
2
(24)

where K0 is the stress-coupling factor and v is the Poisson’s or (M. McLean, personal communication, 2001)
(or strain) ratio of the differentially loaded medium.
Interestingly, for an isotropic porous elastic material, Biot υ * = 0.125 log (200 + b zml ) , (25)
(1941) has demonstrated that equation 20 holds, but with
the following definitions of the effective stresses: where b = 0.5 ± 0.1 (in feet−1) and zml is the TVDml in feet.
It is well documented that in compaction-dominated
σ v = S v − β PP , Tertiary and late Mesozoic basins, both the static (from
σ hmin = S hmin − β PP , (21) vertical core plug tests) and dynamic (from vertical well
logs) Poisson’s ratios for shale decrease with increasing
σ Hmax = S Hmax − β PP , depth (TVDml). As will become evident from the more

02-Vernik_Ch02.indd 40 12-08-2016 20:03:14


Chapter 2: Pore Pressure and Stress State 41

detailed discussion of this subject with respect to uncon- A very shallow water depth and a normal pressure gradient
ventional shales (Chapter 6), the static-versus-dynamic dif- of 10.3 kPa/m are assumed for simplicity.
ference in Poisson’s ratio-based computation of K0 is rela- Because the Poisson’s ratio for wet shale decreases as
tively small if the shale’s anisotropy is weak or moderate. the depth of burial increases, K0 is also expected to de-
Therefore, we can rely on using the VP-VS relationship for crease, thereby giving rise to a reduction in fracture gradi-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

conventional shales (Vernik and Kachanov, 2010) to (1) es- ent with depth. This phenomenon, of course, is grossly in-
timate the dynamic Poisson’s ratio for a range of shale po- consistent with numerous in situ estimates of ShG that have
rosities, from 35 to 5% (at respective TVDml depths from been made in wells using fracture closure pressure mea-
0.3 to approximately 4.6 km, assuming normal compac- surements and LOT data, which commonly cluster around
tion), and to (2) derive K0 as a function of depth by using the empirical models given by equations 24 and 25. Both
this true Poisson’s ratio model in equation 20. models gradually approach the overburden gradient at great
The minimum-horizontal-effective-stress-gradient pro- depths. This observation unambiguously implies that the
files in shales calculated on the basis of true and pseudo- theoretical uniaxial compaction model of stress coupling is
Poisson’s ratios are compared in Figure 11. All calculations hardly rational in geologic basins undergoing compaction,
are based on equation 23, with the stress-coupling factor K0 because it depends on the true Poisson’s ratio of shale and
computed from equation 20. The shale Poisson’s ratio- ignores inelastic deformation. This important conclusion,
based model, in which the Poisson’s ratio decreases from first reached by Eaton and Eaton, immediately suggests
0.48 to 0.32 with depth, while porosity decreases from 35 to that reasonable accuracy in fracture gradient prediction
5% (Vernik and Kachanov [2010], their equations 15–17), from sonic velocity data can be achieved only relying on an
is plotted as a dashed line with large dots. This is the theo- empirical adjustment to the uniaxial compaction model.
retical uniaxial strain model. For comparison, equations 24 Another constraint (albeit a much looser one) on the
and 25 are used as plug-ins to equation 20 to realize Eaton minimum horizontal stress comes from the frictional-
and Eaton’s and McLean’s respective empirical approaches. equilibrium condition on faults that have a critical orien-
tation with respect to the stress tensor. The lower bound
on this stress component can be expressed by the follow-
ing relation (Zoback and Healy, 1984):

Sv + Pp ( χ − 1)
Shmin = , (26)
χ

where χ is controlled by the coefficient of friction ϕ for
optimally oriented faults:
2
χ = (ϕ + 1)1/ 2 + ϕ  . (27)

(McLean)
(Eaton and Eaton)

Figure 11.  Alternative methods of fracture-gradient


computation in shales. The assumptions were (1) a normal PP
gradient (green line, at 10.3 kPa/m), (2) very shallow water
depth, and (3) an estimated shale Poisson’s ratio, ranging
from 0.48 (at approximately 0.3 km) to 0.32 (at 4.6 km). That Figure 12.  Mohr’s circles for the two vastly different stress
ratio was estimated with the Vernik-Kachanov model (Vernik states in the sandstone reservoirs shown in Figure 10. Larger
and Kachanov [2010], their equations 15–17) for a range of circles for each depth are based on frictional equilibrium in
shale porosities. Obviously, the uniaxial strain model based optimally oriented faults, whereas smaller circles represent
on the true Poisson’s ratio (dashed line with large dots) is not more realistic stress states that are consistent with available
adequate. The solid black line is the overburden-gradient data from LOTs in the well and also with empirical
model. equation 25.

02-Vernik_Ch02.indd 41 12-08-2016 20:03:16


42 Seismic Petrophysics in Quantitative Interpretation

The friction coefficient normally ranges from 0.4 to 0.7, can be made for a depth interval, our knowledge of the
depending on the fault gauge mineralogy. Figure 12 is a stress tensor in that interval is almost always limited to
Mohr’s circle illustration of the two markedly different just the two components σv and σh.
stress states sampled in the sandstone reservoirs from the However, because stress coupling in shales is greater
West Africa well shown in Figure 10. The shallower res- than it is in coexisting sands (due to the shales’ greater
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ervoir, which is intersected at 3150 m TVD and lies above ductility and consequently larger pseudo-Poisson’s ratio),
the dramatic pore-pressure ramp, is characterized not the effective-stress anisotropy in them is not expected to
only by lower total vertical stress but also by much lower exceed 10 to 15% at the depths of interest in hydrocarbon
pore pressure, compared with those parameters in the exploration. In contrast, the effective-stress anisotropy of
deeper reservoir at 3750 m. 25 to 30% in highly overpressured sandstone reservoirs,
The frictional constraints provide a lower bound on such as the ones described above, can be quite common
the least horizontal stress at both depths, whereas equa- despite the much-less-pronounced total-stress anisotropy
tions 24 and 25, with further differentiation between presented in Figure 11.
sandstones and shales, quite accurately predict the LOT This large relative difference between σv and σh has
results in this well (Figure 10). An even more dramatic also been detected in crystalline rocks of the upper crust
difference is revealed between the two reservoir sand- (Zoback et al., 1993). Nonetheless, the absolute differ-
stones in terms of the effective stresses: σv = 36 MPa and ence in the effective stresses in both cases of siliciclastic
σh = 27 MPa in one, and σv = 14 MPa and σh = 10 MPa in sequences exemplified above does not exceed a few
the other. megapascals. Recognition of the relatively weak stress
A normal-faulting stress regime is the most likely anisotropy at great depths has significant implications in
scenario in the majority of younger sedimentary basins quantitative seismic analysis, because stress-induced ve-
worldwide, meaning that the magnitude of the maximum locity anisotropy is expected not to be a factor in shales
horizontal stress is bounded by the overburden and the and perhaps to be a much smaller factor than previously
least horizontal stress. Any accurate estimation of SHmax in suggested in sands and sandstones (Scott et al., 1993).
any stress regime (normal, strike-slip, or thrust) is hardly Therefore, it may be sufficient to consider just the vertical
possible without a combination of drilling-induced frac- effective stress in rock-physics characterization of the
tures and borehole-breakout analyses using sophisticated subsurface, in the 3- to 7-km-depth range below mud
image logs and triaxial strength tests (Vernik et al., 1992). line — especially in relaxed basins with minor tectonic
Thus, even if a reasonable estimate of the pore pressure loading.

02-Vernik_Ch02.indd 42 12-08-2016 20:03:16


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 3: Seismic Rock Properties and Rock Physics

Seismic rock properties, such as P- and S-wave ve- Theoretical models in rock physics
locities (VP and VS), bulk density, acoustic and shear im-
pedances (AI and SI), and their relationships with other Fluid-saturation effect
rock parameters and effective-stress states, are the pri-
mary focus of this book. Even though rock physics is a All rocks below the water table are fluid-saturated (it
relatively young science (approximately 50 years old), it is to be noted that even dry gas constitutes a pore-filling
has quickly matured and gained popularity among geo- fluid), and therefore such rocks have elastic properties
physicists, petrophysicists, and geologists concerned with that are governed by poroelastic responses. Those re-
quantitative analysis of surface and borehole seismic data sponses cover a wide spectrum of states, from totally
in hydrocarbon exploration and production. ­relaxed to completely unrelaxed (“frozen”), with respect
Undoubtedly, the rock-physics discipline can be con- to pore-pressure equilibration within the representative-­
sidered to be the “engine” of seismic petrophysics. The volume-element-(RVE) subjected to loading. Arguably
main objectives of this chapter are (1) to provide an over- the least controversial and most widely used poroelastic
view of theoretical and empirical relations between seis- theory was published by Biot (1956) and Gassmann
mic properties of rocks and their intrinsic (composition, (1951) and is a continuum-mechanics approximation of
microstructure) and extrinsic (stress state, temperature) poroelastic properties of isotropic rocks. It results in
parameters, and (2) to draw attention to many controver- Gassmann’s relations:
sial issues that persist in this area. Having encountered −1
these issues when using comprehensive seismic, log, and  φ β − φ
M = Md + β 2  + ,
core data analyses over the years, I have developed my  Kf K m  (1)
own preferences, which, by definition, are subjective. G = Gd ,
Therefore, I often resort to the detailed comparative de-
scription of alternative models to clarify my point of view. where M and G are the undrained fluid-saturated P-wave
It is assumed that the reader is at least superficially aware modulus and shear modulus, respectively, Md is the
of these topics. Special attention is given to proper clas- drained-frame P-wave modulus, β is the Biot’s coefficient
sification of the rock-physics models in order to avoid given by β = 1 − K d /K m, φ is the (total) porosity, and Kf
mislabeling some of them as theoretical, when they are no and Km are the fluid and matrix bulk moduli, respectively.
more than heuristic or even empirical. Note that β is defined as the ratio of pore-volume change
The focal point of any rock-physics analysis is in de- to bulk-volume change, at constant pore pressure (Biot,
veloping and understanding the relationships — often 1956). The second term of equation 1 can be referred to as
complex — among the elastic moduli, porosity, and li- the poroelastic modulus (i.e., the pore-space incompress-
thology. These relationships are easily transformed into ibility) that defines the effect of the relaxed pore fluid on
velocity-porosity, VP-VS, and AI-SI space, and are fre- low-frequency rock incompressibility. The relaxation im-
quently used in seismic-rock-property crossplots con- plies pore-pressure equilibration throughout the pore net-
structed in the form of templates. Such templates, when work within the RVE, during seismic period or inverse
adequately designed and thoughtfully applied, constitute frequency.
the basis of quantitative seismic analysis, including AVO These formulas are the most basic ones of Gassmann’s
and deterministic prestack inversion. theory, which relates low-frequency (relaxed), undrained

43

03-Vernik_Ch03.indd 43 12-08-2016 20:02:01


44 Seismic Petrophysics in Quantitative Interpretation

P-wave and shear moduli of the rock with interconnected equation will likely be affected, rather than just the second
porosity. The convenience of this specific form of Gass­ term, is not always appreciated in rock-physics literature.
mann’s equations is in its explicit statement that the men- Despite multiple assumptions and limitations (e.g.,
tioned moduli depend on the sum of (1) the dry (or Mavko et al., 2009), Gassmann’s equations work well in
drained) frame modulus and (2) the poroelastic response the vast majority of conventional siliciclastic reservoir
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

of the pore space, given by the second term in the equa- rocks (sands and sandstones), with two possible excep-
tion. The poroelastic response to the shear deformation in tions: (1) very-low-porosity and microdarcy-permeability
an isotropic rock is zero, so G = Gd without any limitation rocks, and (2) laminated or thin-bed (<10 cm) sand/shale
on the pore geometry. lithofacies, such as medial- to distal-levee deepwater tur-
The power of these most fundamental relations in bidites, in which lithologic and fluid-distribution hetero-
rock physics lies in their ability not only to predict the geneity is easily established. One way to perform the fluid
moduli and, hence, the velocities of saturated rocks, but substitution in case 2 is to “decompose” the interval into
also to invert for the dry-frame linear elastic properties two phases (sand and shale), apply Gassmann’s equations
from in situ velocity measurements. The ultimate goal is to the reservoir phase, and, finally, upscale by using
to replace Kf in the first of equations 1, to match the de- Backus (1962) averaging (equations 34 of Chapter 1).
sired fluid saturation on the basis of the following har- The often-cited failure of Gassmann’s predictions of
monic-averaging equation (isostress condition): fluid-saturation effects across the hydrocarbon-water
contacts in sand reservoirs is generally a gross miscon-
1 − Sw 
−1 ception. In reality, such situations are simply caused by
S
Kf =  w + , (2) slightly enhanced diagenetic sand-grain bonding in the
 Kw K hc 
water leg, compared with retardation of that process in the
oil and gas zones of the reservoir. That diagenetic effect is
where Sw is the water (brine) saturation and Kw and Khc are obviously outside the scope of Gassmann’s theory and
the fluid moduli of water and hydrocarbon, respectively, can be easily established by the increase in shear modulus
given by Batzle and Wang (1992) or, alternatively, by across the oil-water or gas-water contact. If this turns out
more recent and accurate proprietary software codes to be the case, the Gassmann’s equation for bulk modulus
available in the industry. The inverted Gassmann’s equa- will have to be adjusted for the simultaneous increase in
tion for drained (or dry) bulk modulus of the rock can be frame moduli when one is transitioning from hydrocar-
expressed as bon- to water-saturated reservoir rocks in any fluid-sub-
stitution routine.
K m [ Kφ ( K m − K f ) − K f ( K m − K )] Even though Gassmann’s theory is free of assump-
Kd = . (3)
K mφ ( K m − K f ) − K f ( K m − K ) tions about the pore geometry, that geometry does impact

fluid sensitivity, as follows from the second term of the
Even though the inversion is straightforward, it can (and bulk-modulus equation: the poroelastic modulus of the
often does) result in an erroneous or even negative value pore space. As mentioned above, this term (1) defines the
for Kd if the input parameters, such as porosity, water sat- stiffening effect of the saturating fluid at low frequency
uration, and matrix and fluid moduli, are not sufficiently and (2) is not independent of the geometry of the pore
accurate. The errors in Kd calculations may also be caused space because it also contains the frame modulus through
by differences in bed resolution and depth of investigation Biot’s coefficient. In fact, both bulk and shear dry (or
among the different logging tools. The process of Kd in- drained) frame moduli depend strongly on the pore
version using equation 3, followed by forward modeling shapes, and that relationship is discussed later in this
with equations 1 and 2, comprises the fluid-substitution chapter.
routine widely used in seismic interpretation. The transition from low-frequency to high-frequency
Arithmetic averaging of fluid moduli, instead of har- elastic-wave propagation in saturated rocks results in ve-
monic averaging as in equation 2, is sometimes suggested locity dispersion, with the critical (transition) frequency
in the case of inhomogeneous saturation (mostly layered depending on the mechanism of viscoelastic equilibra-
saturation or infrequently, “patchy” saturation) that is not tion — the so-called Biot flow or local “squirt” flow be-
accounted for by Gassmann’s theory (Mavko et al., 2009). tween adjacent inclusions with vastly different compli-
However, this is an unlikely scenario because in most ances, such as isometric pores and microcracks (Mavko
cases saturation inhomogeneity is prescribed by lithologic and Nur, 1978). Guéguen et al. (2009) suggest that the
and/or grain-size inhomogeneity that results in capillary- critical frequency fc for the latter effect in rocks contain-
pressure inhomogeneity. The fact that in the case of litho- ing hypothetical open spheroidal inclusions can be given
logic heterogeneity both terms of the first Gassmann’s by the semiempirical formula:

03-Vernik_Ch03.indd 44 12-08-2016 20:02:02


Chapter 3: Seismic Rock Properties and Rock Physics 45

α 3E In these equations the subscripts d, m, and φ refer to dry,


fc = , (4) matrix, and pore space, respectively. The pore-space com-
20η
pressibility cφ can be expressed as a product of matrix com-
pressibility and pore-shape factor p (Kachanov et al., 1994):
where α is the pore aspect ratio, E is the Young’s modulus
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

of the rock (in gigapascals or GPa), and η is the fluid vis- cφ = pcm , (7)
cosity (in pascal seconds or Pa ⋅ s).
Consider a water-saturated sandstone with 20% po-
rosity, an E value of 30 GPa, and the microcrack aspect so that after substituting this formula into equation 6, we
ratio in the range from 0.01 to 0.001 (typically inverted, obtain:
based on crack-closure stress from the laboratory experi-
K d = K m (1 + pφ )−1 . (8)
ments). Because the dynamic viscosity of water is 10−3
Pa ⋅ s, the respective range of critical frequency from
The shape factor varies from low values for spheres to
equation 4 is from 1.5 MHz to 1.5 kHz. Replacing water
very high values in lower-aspect-ratio ellipsoids, but it de-
with high-API oil under high temperatures can reduce
generates in cracks. This follows from the introduction of
fluid viscosity to 0.5 × 10−3 Pa ⋅ s and shift the critical-
the pore-compliance-contribution tensors Hijkl: Per refer-
frequency range to values between 3 MHz and 3 kHz.
ence volume V, the extra strain due to the presence of a
Noting that modern sonic logs operate in the 2- to 8-kHz
single pore is
frequency band, we can see that at seismic to sonic fre-
quencies, velocity dispersion may be minor in most sand- ∆ε ij = Hijkl σ kl , (9)
stones and negligible in shales.

and hence, for multiple pores,


Drained rock-frame moduli
Theoretical models establishing compositional and
∆ε ij = ∑H (m )
ijkl σ kl
pores
= ∆Sijkl σ kl , (10)
microstructural dependencies of the first terms in the m
Gassmann relations — the drained rock-frame moduli —
where the lower limit m refers to a single pore and the
most often start with summation of the compliance contri- (m )
individual H ijkl tensors should, in principle, reflect inter-
butions from the matrix, cracks, and pores to the effective
actions between pores (because they affect compliance
compliance tensor of the material (e.g., Bristow, 1960;
contributions of individual pores). However, this factor is
Zimmerman, 1991): (m )
very difficult to account for, so the individual H ijkl values
m cr acks por es are taken as the ones for isolated pores.
S ijkl = S ijkl + ∆S ijkl + ∆S ijkl . (5)
Equations 9 and 10 express the fact that Δεij are linear
functions of applied stresses; the challenge is to construct
The summation applies to compliances, not to stiffnesses, H-tensors in terms of pore geometries. For any elastically
which is intuitive from the physical point of view because isotropic pore (this covers the spherical shape as well as
cracks and pores are naturally regarded as sources of extra other pore shapes, including “irregular” ones, that pro-
strains and hence their compliance contributions are duce an elastically isotropic response), Hijkl has the fol-
summed. lowing general form:
The assumption that the orientation distribution of
matrix minerals, cracks, and pores in sandstones is ran- Vp  δ ik δ jl + δ ilδ jk 
dom implies isotropic elastic properties. This restriction Hijkl = L + Nδ ijδ kl  , (11)
V  2 
can be removed and results can be extended to arbitrary
orientation distributions by using the micromechanics
framework (Kachanov, 1992; Kachanov et al., 1994). where V and Vp are the bulk and pore volumes,
Then the noninteraction approximation (NIA) relates stat- ­respectively,  and L and N are the factors related to pore
ic drained compressibility of an isotropic composite with geometry — for instance, for the spherical pore,
cavities cd to its solid-matrix and pore-space compress- L = 15(1 − v m ) /[2G m (7 − 5v m )] and N = (1 − v m )(1 + 5v m ) /
ibilities (Zimmerman, 1986): cd = cm + φ cφ , or, in terms [2K m (1 − 2v m )(7 − 5v m )] , where Gm and Km are the shear
of bulk moduli, and bulk moduli of the solid matrix, respectively, and νm
is its Poisson’s ratio. For ellipsoidal shapes, H-tensors
1 1 φ can be obtained from Eshelby’s theory (Kachanov et al.,
= + . (6)
Kd Km Kφ 1994). For more realistic, nonellipsoidal shapes in the

03-Vernik_Ch03.indd 45 12-08-2016 20:02:03


46 Seismic Petrophysics in Quantitative Interpretation

two-dimensional case, H-tensors can be derived analyti- long as the ratios are smaller than approximately 0.10. For
cally for practically any shape by using the conformal example, increasing the pore aspect ratio α from 0.05 to
mapping technique (Zimmerman, 1986; Kachanov et al., 0.10 while keeping the crack diameter (but not the crack
1994). However, in the three-dimensional case progress volume) constant would double the porosity but would
has been limited (Sevostianov et al., 2008). leave the elastic properties almost unchanged. This is il-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Modeling of the compliance contribution of random- lustrated in Figure 1, which shows contributions of a sphe-
cr acks
ly oriented cracks ∆S ijkl in the framework of NIA — roidal pore to the overall elastic compliances as functions
that is, when each crack is assumed to be embedded in a of the aspect ratio: Below a ratio of 0.10, the normalized
remotely applied stress σ ij∞ unperturbed by neighbors — compliance curves are practically horizontal.
yields the following term (Bristow, 1960): The root of the misconception is that, in expressing
the effective elastic properties as a function of crack aspect
64(1 − ν m2 ) ratio α, it is the crack porosity φcrack that is kept constant.
cracks
∆S ijkl = η (δ ik δ jl + δ ilδ jk ) , (12)
9(1 − ν m )E m However, this implies that if the number of cracks per unit
volume is also kept constant as α decreases, the cracks
where Em is the Young’s modulus of the matrix and η is must increase in diameter — that is, they propagate!
the crack-density parameter introduced for penny-shaped Alternatively, the number of cracks must dramatically in-
(circular) cracks by Bristow (1960) and, independently, crease to unreasonable levels to maintain a constant crack
by Walsh (1965): porosity when their aspect ratio is substantially reduced.
The proper way to examine aspect-ratio dependence is to
1

η =
V ∑a ( k )3
.

(13) keep crack diameters fixed — and not their volume — as
the aspect ratio is changed (as in Figure 1). This leads us
to expressing the effective moduli of rocks with narrow
Here, a(k) is the radius of the kth crack and V is the repre- cracklike voids in terms of crack density that is indepen-
sentative volume. This fundamental parameter is general- dent of α and is defined by equation 13.
ized to elliptical cracks by Budiansky and O’Connell Obviously, if all cracks have identical aspect ratios,
(1976). Importantly, the aspect ratios of cracks (crack then crack density η can be related to crack porosity, but
openings) do not enter equation 13. This reflects the fact this assumption appears artificial and is difficult to verify.
that crack compliances are almost insensitive to aspect Moreover, even in cases when the α values are identical,
ratios, provided, as will be demonstrated in the next sec- characterization by crack porosity is a poor choice be-
tion, the aspect ratios are smaller than approximately cause of the very low sensitivity of crack compliances to
0.10. In other words, the aspect ratios, and hence the this parameter. Thus, crack porosity is irrelevant for rocks
“crack porosity,” are irrelevant in the context of the with low-aspect-ratio voids and cracks, as far as their
drained effective elastic properties.
The issue of the importance of aspect ratios in dry
and saturated rock modeling has remained controversial 4
in geophysical literature until recently. The topic clearly
warrants a separate treatment before our discussion on the 3
Normalized compliances

drained rock frame moduli can be continued.


2

Cracks in dry and fluid-saturated rocks:


The effect of aspect ratios 1

Vernik and Kachanov (2012) address the problem of 0


aspect ratios in detail. A frequently advanced viewpoint is
that aspect ratios of narrow, cracklike pores — and hence,
the crack porosity φcrack — are important parameters for 0.001 0.01 0.1 1.0
overall stiffnesses (e.g., Sayers, 2008; Mavko et al., 2009; Aspect ratio α
Smith et al., 2009; Ruiz and Cheng, 2010). This often Figure 1.  Normalized spheroidal pore compliances (red
leads to two conclusions of a practical nature: (1) crack is normal, blue is shear) as functions of the aspect ratio,
porosity is an important parameter for rocks with low-as- provided the pore diameter (and not volume) is kept constant.
pect-ratio pores, and (2) the said ratios need to be known. The curves become almost horizontal at aspect ratios below
However, as Vernik and Kachanov (2012) suggest, this 0.10, thereby indicating pore-compliance insensitivity to α in
viewpoint is incorrect: Aspect ratios have almost no effect this range. Adapted from Figure 2 of Sevostianov and
on dry or fully drained saturated crack compliances, as Kachanov (2011). Used by permission.

03-Vernik_Ch03.indd 46 12-08-2016 20:02:04


Chapter 3: Seismic Rock Properties and Rock Physics 47

static and inverted (from low-frequency-velocity data)


dynamic dry frame moduli are concerned. The relevant
parameter is crack density as given by formula 13 (or by
its generalization to a crack-density tensor in cases of
nonrandom orientations).
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

As Vernik and Kachanov (2012) also pointed out, the


insensitivity of the dry moduli to crack aspect ratios is
actually beneficial because these ratios cannot be mea-
sured reliably in situ. Moreover, they cannot even be de-
fined: Cracklike voids usually have irregular shapes, and
it is not clear what we should call their aspect ratios. The
insensitivity just referred to eliminates this difficulty.
For fluid-saturated cracks, α may be an important pa-
rameter because it determines the stiffening effect of fluid
on the normal compliance of a crack. Under the assump-
tion that fluid mass in each crack remains constant dur-
ing an elastic-wave period (the undrained unrelaxed re-
Figure 2.  Normalized saturated bulk modulus from
sponse) — which typically corresponds to core measure-
Gassmann’s equation, versus the crack aspect ratio. The case
ments at an ultrasonic frequency of 1 MHz — this
of a background porosity of 8% is represented by solid lines,
stiffening effect is controlled by the dimensionless parame- and the case of zero background porosity is shown as dashed
ter ω = K f /αK b (O’Connell and Budiansky, 1974), where lines. Crack density η is kept constant at 0.2 for both cases.
Kb is the bulk modulus of the background rock (Kb = Km Blue is water, green is oil, and red is gas. Adapted from Figure
for a zero-porosity, crack-free matrix) and Kf is the fluid 2 of Vernik and Kachanov (2012). Used by permission.
modulus. Greater values of Kf correspond to a stronger
stiffening effect of the fluid. This relationship indicates
that the stiffening effect of the fluid at high frequencies is To summarize, for dry rocks, as well as for fluid-sat-
stronger for very narrow cracks and may become weak for urated rocks under fully drained conditions (static or
cracks with larger aspect ratios. very-low-frequency dynamic loading), the aspect ratios α
However, at lower frequencies of 10 to 104 Hz (un- do not matter as long as α is less than 0.10. However, in
drained, relaxed response for the majority of conventional the case of an undrained unrelaxed response to high-fre-
reservoir rocks), which are typical in seismic and sonic quency dynamic loading, cracks with a small aspect ratio
applications (e.g., Guéguen et al., 2011), the parameter ω may, depending on the fluid modulus, experience a strong
gradually loses significance. In such cases, Gassmann’s stiffening effect from the fluid. That stiffening is con-
equation directly expresses the saturated moduli in terms trolled by the cracks’ aspect ratios, and in such a case the
of φpore, dry frame moduli, fluid compressibility, and total exact α values need to be known.
porosity φ, which comprises both the crack and pore con- However, in typical seismic and sonic applications
tributions (φ = φpore + φcrack ). Figure 2 illustrates the effect we encounter the undrained relaxed response, in which
of aspect ratios on the fluid-saturated bulk modulus for the crack aspect ratio may or may not be an important
two different scenarios: (1) φpore = 0, and (2) φpore = 0.08. parameter. The importance will depend on the fraction of
In both cases, φcrack = (4 / 3)παη . It is seen that the exact total porosity that is due to cracks. If the fraction attribut-
values of crack aspect ratios inverted from this formula in able to cracks is nonnegligible, the aspect ratios are im-
combination with Gassmann’s equation, although being portant because of the contribution from crack porosity to
important in the absence of pores, become increasingly the total porosity φ in Gassmann’s equation. More typi-
irrelevant as the background porosity increases from zero cally, however, if the crack porosity is a negligible frac-
toward values of 5 to 10%. Such porosity values are typi- tion of total porosity φ, once again the aspect ratios are
cal of tight, low-porosity sandstones. relatively unimportant.
As will be discussed in the next section, the dry frame
moduli entering Gassmann’s formula can be micro­ Drained rock-frame moduli: A mixture
mechanically evaluated in terms of crack density, total of cracks and pores
porosity, and pore-shape factors. Unless a special case
with relatively high-aspect-ratio (“inflated”) cracks is hy- Pore shapes may contribute strongly to the effective
pothesized, the more typical cracks and grain-boundary elastic compliances. In the 3D case, the ellipsoidal shape
microcracks in situ account for a negligible fraction of is the only one for which analytical results are available,
total porosity, and their aspect ratios barely matter. and analyses of pores are often based on those results.

03-Vernik_Ch03.indd 47 12-08-2016 20:02:05


48 Seismic Petrophysics in Quantitative Interpretation

However, pores in rocks usually have highly irregular, −1


 3(1 − ν m ) 16(1 − ν m2 ) 
nonellipsoidal shapes, and that may make the ellipsoid- K d = K m 1 + φ+ η ,
based results inapplicable or misleading. In particular, as  2(1 − 2ν m ) 9(1 − 2ν m ) 
demonstrated by Sevostianov et al. (2008), the concavity/ −1
 15(1 − ν m ) 32(1 − ν m )(5 − ν m ) 
convexity factor plays a strong role: Concave shapes pro- G d = G m 1 + φ+ η .
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

duce much larger compliance contributions than do con-  7 − 5ν m 45(2 − ν m ) 


vex ones of the same volume.
(15)
Moreover, in light of the previous section, represent-
ing the entire pore space of sands and sandstones by el- It should be emphasized that these expressions imply
lipsoidal inclusions of varying aspect ratios is incorrect that compliances (the inverses of Kd and Gd), and not
because at a fixed porosity the reduction in aspect ratio stiffnesses, are linear functions of φ and η — that is,
causes the inclusions’ diameters or numbers per unit vol- equations 15 are fully consistent with equation 5.
ume to increase. At aspect ratios smaller than 0.01 these The numerical values of the shape factors for spheres
effects result in dramatic growth of the crack-density — the coefficients at φ in equations 15 — are 1.81 and
term to values that are incompatible with the strong co- 2.04, respectively, for arenites with a matrix Poisson’s
hesion attributable to the consolidated rock matrix (e.g., ratio of 0.146 (computed from Table 2 of Chapter 1). The
Sevostianov and Kachanov, 2011). numerical values of the coefficients at the crack-density
The effect of shape on dry frame compressibility can term for arenites are A(νm) = 2.46 for Kd and B(νm) = 1.59
be described by the factor p defined in equation 7. In gen- for Gd. Whereas the contrast between p and q in clean ar-
eral, shape factors are different for bulk and shear moduli, enites (vcl < 0.02) is greater and cannot be ignored, for
so p ≠ q. For example, for spherical pores in the matrix arenites (0.02 < vcl < 0.12) one can assume the difference
with Poisson’s ratio νm (e.g., Berryman, 1980): between them to be negligible in practical applications;
that will be discussed later in more detail. Incidentally,
3(1 − ν m ) the pore-shape factor of circles for the bulk modulus in
p = ,
2(1 − 2ν m ) 2D is independent of material properties and equals 2.0.
(14)
15(1 − ν m ) Equations 15 can be extended to the ellipsoidal pores
q= .
7 − 5ν m and planar cracks of elliptical in-plane shapes. The shape

factors for ellipsoids are controlled by their aspect ratio
Similar expressions for spheroids are also available in the and attain the minima for a sphere because that is the stiff-
literature (e.g., Berryman, 1980). They reduce to equa- est shape of a given volume (Berryman, 1980; Zimmerman,
tions 14 when α = 1 (Figure 3). 1991). Schematic illustration of more realistic pore shapes
For a mixture of spherical pores and circular cracks observed in porous materials is given in Figure 4. To reit-
considered in the NIA, the equations above yield the follow- erate the general rule: The greater the concavity, the larg-
ing expressions for the dry effective bulk and shear moduli er the p-factor value. Also, as mentioned above, pore-
(Walsh, 1965; Benveniste, 1987; Kachanov et al., 1994): shape factors governing shear compliance are not neces-
sarily similar to p-factors in 2D computations; in fact, the
q-factor is more than twice the value of p for hypotrochoi-
dal pore shapes illustrated in Figure 4.
The problem with the NIA model generalized for
mixtures of pores and cracks is that it violates Hashin-
Shtrikman’s upper bound when applied to a material
with spherical pores and zero crack density. Therefore,
its application in the form given by equations 15 is not
advisable for rocks with porosity greater than a few per-
cents. However, most importantly, it does represent a
building block for some of the effective-rock-property

p = 2.0 p = 3.2 p = 6.0 p>6 p = π/4α

Figure 4.  Pore-shape factors p for the bulk modulus,


Figure 3.  Ellipsoidal pore-shape factors versus aspect ratio computed in 2D for different shapes. Data are from Mavko
(Berryman, 1980), for arenites in 3D. and Nur (1978) and Kachanov et al. (1994).

03-Vernik_Ch03.indd 48 12-08-2016 20:02:05


Chapter 3: Seismic Rock Properties and Rock Physics 49

models that, as demonstrated below, are consistent with a)


the theoretical bounds.

Effects of pore/crack interactions


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

The form of equations 15, stating that effective com-


pliances of dry solid with spherical pores and cracks (the
inverses of Kd and Gd) are linear functions of crack density
and porosity, assumes that there are no interactions among
cracks and pores — thus giving rise to the term NIA. In the
case of cracks, computations by Grechka and Kachanov
(2006) indicate that the NIA constitutes a satisfactory ap-
proximation for low to moderate crack densities (values
smaller than 0.14). In the simulations of Saenger et al.
(2004) and Kushch et al. (2009), on the other hand, crack
interactions were found to have a noticeable softening ef-
fect that caused the effective constants to lie between the
predictions of the NIA and the differential scheme and to b)
be substantially closer to the latter.
Even then, however, the interaction effect is relatively
mild. In fact, the uncertainties related to describing cracks
of irregular shapes in terms of the crack-density parame-
ter η are likely to be a stronger factor than the interaction
effect is. These considerations justify application of the
NIA to very-low-porosity rocks.
The presence of pores enhances the interaction ef-
fects because it raises the average stress in the matrix and
thus increases compliance contributions of both pores and
cracks. Mori and Tanaka’s scheme is the simplest one that
accounts for this factor, and it applies to an arbitrary mix-
ture of pores and cracks (Mori and Tanaka, 1973). In this
scheme, inclusions are placed into the average (over the
solid phase) stress σ ij . Because this average stress is
S
related to the remotely applied stress σ ij∞ by the simple
formula σ ij = (1 − φ )−1 σ ij∞ , results for the effective Figure 5. Water-saturated VP-versus-porosity space for
S
constants are obtained from the ones in the NIA by divid- arenites with hypothetical ellipsoidal pores and cracks,
ing φ and η over (1 − φ): showing Hashin-Shtrikman (HS) bounds and Mori-Tanaka/
Gassmann model lines, with (a) a zero crack density and pore
−1
 pφpore η  aspect ratios varying from 0.01 to 1.0, and (b) a constant
K d = K m 1 + + A (ν m ) ,
1 − φ 
aspect ratio of 0.15 and a crack density that is allowed to vary
 1−φ
(16) widely. Note that a crack density of 5.0 is only theoretically
−1 possible for consolidated rocks. Matrix properties of arenites
 qφpore η 
G d = G m 1 + + B (ν m ) .
φ 
are used in all computations.
 1 − φ 1 −

with the Hashin-Shtrikman (HS) bounds (Hashin and
For thin microcracks as well as low-aspect-ratio mac- Shtrikman, 1963); both models are applied to water-saturat-
roscopic fractures that withstand the effective-stress load- ed arenites. The respective Hashin-Shtrikman upper and
ing in situ, the crack porosity can be neglected, so lower bounds for two-phase composites can be expressed as
φpore ≈ φ and the equations simplify further. Note that the
product pφpore approaches the crack density if the pores + Mm (Km − Kf ) φ
are shrunk to cracks, and is, therefore, absorbed by the η K HS = Km − ,
M m − ( K m − K f )(1 − φ )
term in this limit (Kachanov et al., 1994). (17)
Figure 5 illustrates the Mori-Tanaka model for a mix- +  5Mm φ 
GHS = Gm 1 −  ,
ture of ellipsoidal pores and circular cracks, in comparison  5 M m − 2( K m + 2Gm )(1 − φ ) 

03-Vernik_Ch03.indd 49 12-08-2016 20:02:07


50 Seismic Petrophysics in Quantitative Interpretation

and where l m = G m (9K m + 8G m ) 6( K m + 2G m ) , the P-


 ( K − K f ) (1 − φ )  wave modulus of the solid matrix Mm = Km + (4/3)Gm, and

K HS = K f 1 + m , p and q are the respective shape factors expressed by
 K f + ( K m − K f )φ  (18) Berryman (1980) for spheroids in terms of their aspect

Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

GHS = 0. ratio. The model can be run using any isotropic mixture of

spherical pores and microcrack concentrations (e.g.,
In computations, the water modulus Kf is equal to 2.5 GPa, Smith et al., 2009), in which case the pφ and qφ terms are:
the water density ρf is equal to 1.02 g/cm3, and the matrix
properties of arenites are specified in Table 2 of Chapter 1. pφ = ppore (φ − φcrack ) + pcrack φcrack ,
In the final step, VP is computed as VPb = ( Mb /ρb)1/ 2 , (20)
qφ = qpore (φ − φcrack ) + qcrack φcrack .
where the subscript b denotes brine saturation.
Figure 5a is constructed by keeping the crack-density
term η equal to zero and varying the aspect ratio α of the It should be emphasized that using the Kuster-Toksöz
ellipsoidal pores from 0.01 to 1.0. Aspect ratios are con- model or its empirical derivatives (e.g., Xu and White,
verted to shape factors as shown in Figure 3; however, 1995) without porosity partitioning, as expressed in equa-
aspect ratios below 0.01 are considered to be more typical tions 20, is not quite relevant for most sedimentary rocks
of cracks and are not taken into account. In contrast, and can lead to incorrect conclusions about their micro-
Figure 5b is designed to show the effect of crack density structure. Moreover, when using low-ellipsoidal-inclu-
at the fixed aspect ratio of the ellipsoidal pores (α = 0.15). sion aspect ratios that are more typical of cracks (α < 0.01)
The crack density is allowed to vary over an extremely to represent the total porosity (i.e., φcrack = φ ), one should
wide range (0.1 to 5.0), even though that is not physically be aware that even at low to moderate porosities, such a
meaningful for consolidated rocks. Note, for instance, representation leads to an unrealistically high crack den-
that the crack density of 5.0 corresponds to a crack poros- sity that, as previously mentioned, can be envisioned only
ity of 2% if the constant aspect ratio of 0.001 is assumed; for quite disintegrated rocks.
crack density and crack porosity of such a magnitude can On the other hand, one should also realize that typical
be justified only for totally disintegrated rocks. pores, at least in siliciclastic rocks, are not spheres but in-
Both realizations produce a family of curves that lie stead are more compliant geometries. Even the sintered
within the elastic bounds. Moreover, the upper bound coin- glass-bead samples selected by Berge et al. (1995) to repre-
cides with the Mori-Tanaka model for the case of spherical sent artificial sandstones clearly display elastic-moduli de-
pores (α = 1) without cracks, which was first proved by pendence on porosity that would be typical of a more-com-
Benveniste (1987). Although application of the model to the plex pore-space geometry. In fact, spherical pores can only
simple and quantifiable spherical pore geometry is straight- be assumed to exist in their low- to medium-porosity sam-
forward, it may have only theoretical significance because, ples. At porosities greater than 20 to 30%, a gradual migra-
as will be discussed later, real sandstones (as well as the tion toward softer, hypotrochoidal pore cross sections must
majority of other consolidated sedimentary rocks) actually be envisioned (see the thin sections in Berge et al., 1995).
contain more complex pore shapes. Referring to Figure 5b, Moreover, some low crack densities cannot be ruled out in
it is very important to be aware that it is impossible to ex- their experiments, because microcracks are indeed ob-
tract matrix moduli or velocities when porosity approaches served in their thin-section photomicrographs. Respectively,
zero in rocks with open cracks (see also Smith et al., 2009). Berge et al.’s conclusion that the Mori-Tanaka and Kuster-
Levin et al. (2004) have demonstrated that the Mori- Toksöz schemes both fail in moduli prediction or micro-
Tanaka scheme (elaborated by Benveniste, 1987) and the structure inversion is questionable. Clearly, by using a
Kuster and Toksöz (1974) scheme both belong to the combination of pores and microcracks, as implied in the
same effective-field group of theories. The Kuster-Toksöz Mori-Tanaka model of equations 16, a reasonable match to
equations for dry effective moduli of rocks with ellipsoi- the experimental data could be achieved.
dal inclusions of varying aspect ratios can be simplified Figure 6 compares the dry frame P-wave modulus
considerably and can be given as Md from a Kuster-Toksöz model, which was applied to a
hypothetical mixture of spheres and low-aspect-ratio in-
 4  clusions (α = 0.001) in an arenite matrix, with the Md
K m  M m − G m pφ  modulus from a corresponding Mori-Tanaka model. The
 3 
Kd = , correspondence between the two models is established
M m + K m pφ
(19) by means of the crack-density parameter η varying from
zero to 0.3 in the Mori-Tanaka scheme and a respec-
G m G m + l m (1 − qφ ) 
Gd = G = , tive crack-porosity computation for the Kuster-Toksöz
l m + G m (1 + qφ) scheme using φcrack = (4 / 3)παη . Pore-shape factors for

03-Vernik_Ch03.indd 50 12-08-2016 20:02:08


Chapter 3: Seismic Rock Properties and Rock Physics 51

As Mavko et al. (2009) noted, for a fixed-inclusion


geometry and porosity, the EM schemes yield elastic
moduli that are softer than those from the Kuster-Toksöz
theory, let alone the Mori-Tanaka model’s predictions in
light of the results shown in Figure 6. Kachanov et al.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

(1994) provide a simple 2D analysis of these various ap-


proximate schemes for cracked solid materials and con-
clude that, for a random distribution of crack locations,
the Mori-Tanaka scheme and the DEM yield physically
sound results. In contrast, the self-consistent scheme of
Budiansky and O’Connell (1976) grossly underpredicts
elastic stiffnesses and has a “cutoff” point of approxi-
mately η = 9/16 for a solid with randomly oriented cracks.
When the crack density reaches the cutoff value, the pre-
dicted moduli reduce to zero, which is obviously
incorrect.
To summarize, in practical rock-physics applications
to quasi-isotropic rocks, such as sandstones, it could suf-
Figure 6.  Comparison of the dry frame P-wave modulus of fice to model their drained effective moduli by using the
arenites with hypothetical spherical pores and cracks, as a functional form of equations 16. Whereas noticeable dif-
function of porosity and crack density, produced with the ferences exist between the Mori-Tanaka, Kuster-Toksöz,
Mori-Tanaka (MT) and Kuster-Toksöz (KT) theories. The and DEM models, those differences appear to be second-
upper two overlapping curves are the MT scheme realization ary compared with the strong effect of pore-shape factors.
for spherical pores and zero crack density and the KT model Consequently, estimating pore-shape factors for reservoir
for spherical pores; these curves also coincide with the rocks is clearly one of the most vexing challenges in rock
Hashin-Shtrikman upper bound. For the lower six curves, the physics and in geophysics in general.
crack aspect ratio is 0.001 and the crack porosities of 0.042,
0.084, and 0.125% are derived from crack densities of 0.1,
0.2, and 0.3, respectively. Contact models
The grain-contact-based models are, in the strict sense,
spherical pores in an arenite matrix are ppore = 1.81 and only appropriate for static and dynamic moduli estimates in
qpore = 2.04, and for cracks the hypothetical factors are unconsolidated sands. They are almost exclusively based on
pcrack = 586 and qcrack = 381 (Figure 3); matrix properties the Hertz-Mindlin solution (Mindlin, 1949) for the elastic
are taken from Table 2 of Chapter 1. stiffnesses of two elastic spheres in contact. Generalization
For spherical pores without cracks, both models coin- of this solution in computing the effective moduli of sands
cide with the Hashin-Shtrikman upper bound (Figure 6; requires knowledge not only of the individual contact areas
see also Benveniste, 1987). However, increasing the crack but also of the coordination number (the number of contacts
porosity φcrack produces a gradually increasing separation per grain). It is important to note that these models are based
between these two effective-field (EF) models. The sense on spherical grains, whereas most grains in unconsolidated
of this separation is to indicate that the Kuster-Toksöz sediments may deviate from the ideal spherical shape.
model produces dramatic underprediction of elastic stiff- Dvorkin and Nur (1996) formulate the Hertz-Mindlin
ness as compared to the Mori-Tanaka model at medium to effective moduli as:
high total porosity and when crack density is greater than
0.1–0.2. The reason for the discrepancy is not currently 1
understood, so additional research, including numerical  n 2 (1 − φc )2 G m2  3
K HM = σ ,
modeling, will be needed to shed more light on it. 18π (1 − ν m ) 
2 2

Another group of approximate schemes includes effec- 1


(21)
tive-medium (EM) models, which place an ellipsoidal inclu- 5 − 4ν m  3n (1 − φc ) G m2
2 2 3
sion into an “effective matrix” of reduced stiffness. This can G HM =  σ ,
5(2 − ν m )  2π (1 − ν m )
2 2

be done in one step (using the self-consistent scheme of
Budiansky and O’Connell, 1976), or in infinitesimal incre-
ments, with the reference matrix being recalculated at each where KHM and GHM are, respectively, the effective bulk
step (using the differential scheme, or DEM, developed by and shear moduli, φc is the critical (i.e., roughly deposi-
Vavakin and Salganic, 1975; Norris, 1985, and others). tional) porosity, σ is the isotropic effective stress, and n is

03-Vernik_Ch03.indd 51 12-08-2016 20:02:09


52 Seismic Petrophysics in Quantitative Interpretation

the coordination number, which is largely unknown and well as many other contact models) provides less-than-
therefore presents the greatest uncertainty of the model. optimal prediction of the laboratory results, compared
Moreover, careless application of this model to sands in with results from the exponential model (Figure 7).
which varying porosities (at different burial or current It is also noteworthy that the contact models are even
depths and different levels of diagenetic grain bonding) less accurate in predicting the stress-dependent velocity
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

replace their depositional porosity is rather misleading. variation measured on core samples of more-consolidated
For the same reason, it is totally inappropriate to apply sands that have ever so slightly flattened initial grain con-
such models to even slightly consolidated rocks. tacts caused by early diagenetic processes. Such a contact
However, a primary pitfall of this model is its failure to configuration clearly disqualifies those sands from appli-
accurately predict both moduli simultaneously and hence cation of the Hertz-Mindlin theory. More detailed discus-
VP/VS. This observation is often attributed to the fact that sion of the stress-dependent elastic properties is deferred
different grain pairs in reality may have different stresses to the following section.
(both normal and shear) acting on them. According to
Bachrach and Avseth (2008), that difference in stresses
among grain pairs results in partial slip in some contacts Drained frame moduli and
but not in others, thereby requiring introduction of an ad-
ditional fitting parameter to describe what fraction of the effective stress
grain pairs is frictionless. The effective-stress dependence of drained-rock elas-
Moreover, the power-law stress dependence given by tic constants lies in the foundation of experimental rock
the Hertz-Mindlin theory is not consistent with laboratory physics. Indeed, the pioneering work of King (1966), Nur
measurements of VP and VS as a function of confining pres- and Simmons (1969), and others started with laboratory
sure on loose sand packs, let alone on more-consolidated experiments under elevated isotropic and anisotropic stress
sands and sandstones. In fact, the actual measured stress conditions and sought to explain rock behavior under
dependence of the elastic moduli typically is exponential stress, for seismic interpretation. It is important not to con-
(Eberhart-Phillips, 1989; Vernik and Hamman, 2009). fuse the stress effects on rocks with the diagenetic effects,
To illustrate this point, Figure 7 shows the experi- including the simultaneous actions of pressure, tempera-
mental data acquired on dry Ottawa sand by Domenico ture, and subsurface fluid chemistry that are responsible for
(1977), superposed by the best fitting realization of the transforming the initial sediment into more-consolidated
Hertz-Mindlin model. The critical porosity of this clean- sands, sandstones, and shales. Obviously, these processes
arenite sand is taken to be 38%, which was the actual will receive special attention later in the book.
measured porosity at relatively low stress, and the grain In light of the discussion presented in Chapter 2 on the
coordination number used is chosen to be as low as n = 5 effective-stress tensor operating in the majority of sedimen-
for this realization. Even so, the Hertz-Mindlin theory (as tary basins, the vertical effective stress is used throughout
the rest of this book as a proxy for the mean effective stress.
Moreover, because we are interested in the behavior of the
dry rock frame moduli as a function of stress, the pore-pres-
sure term is set to zero so that the effective stress is equal to
the total stress. The term “dry” refers to humidity conditions
maintained during the laboratory experiments. If both Md(σ)
and Gd(σ) functions are known, prediction of the fluid-satu-
rated-rock behavior in the majority of cases is straightfor-
ward with the use of Gassmann’s relations.
Because intra- and intergranular microcrack closures
have been widely accepted as the main elastic mechanism
responsible for the variation in dry rock frame moduli
with stress in consolidated rocks (King, 1966; Nur and
Simmons, 1969; Mavko and Jizba, 1991), it is important
to consider the so-called crack-closure stress σclose that
has been derived for an elliptical inclusion in a solid ma-
trix in 2D (Nordgren, 1972; Mavko et al., 2009):
Figure 7.  Velocities versus effective-stress measurements in
dry Ottawa sand (Domenico, 1977), fitted by using the α oEm
exponential function (Vernik and Hamman, 2009) and the σ close = , (22)
2(1 − ν m2 )
Hertz-Mindlin model. Fitting parameters are given in the text.

03-Vernik_Ch03.indd 52 12-08-2016 20:02:09


Chapter 3: Seismic Rock Properties and Rock Physics 53

where αo is the initial crack aspect ratio and Em and νm are sand, deepwater Gulf of Mexico; the Troll sand, North
the Young’s modulus and Poisson’s ratio of the matrix, re- Sea; the Berea Sandstone, USA; and sandstone from the
spectively. The localized effective stress between the grains Travis Peak Formation, USA). To avoid exceeding the
of a porous material is amplified by a factor of (1 − φ )−1 in maximum burial stress, which can be accompanied by in-
consolidated sandstones or by a factor of (1 − φ /φc )−1 in elastic deformation in the form of grain crushing and/or
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

loose sands, with respect to the far-field stress. Consequently, rotation, the pressure range was limited with respect to
it follows that for porous sandstones the microcrack-closure the consolidation/cementation level of each rock sam-
stress is proportional to the initial aspect ratios of these mi- ple. Hence, the ultimate pressure considered varied from
crocracks and the solidity term, which may not be neglected 40 MPa in unconsolidated Ottawa sand to 100 MPa in
as it is in equation 22. Obviously, closure of microcracks low-porosity, well-cemented Travis Peak sandstone.
naturally reduces the crack density in equation 16 and, Under stress, velocities measured on low-porosity
hence, increases the elastic moduli. sandstones are often reported to behave similarly to some
In consolidated sandstones, low-aspect-ratio cracks crystalline rocks such as granites — the initially high gra-
lead to a reduction in the dry elastic moduli but have prac- dients of increase in velocity with loading gradually flat-
tically no impact on porosity. In principle these micro- ten and can be approximated by a quasi-linear increase, in
cracks can occur anywhere in the rock frame, but they are the stress range from 0.15 to 1.0 GPa. That quasi-linear
more likely to be associated with flat grain contacts. In behavior can be attributed to the elastic stiffening of rock-
unconsolidated sands lacking any diagenetic cement the forming minerals beyond the microcrack-closure stress,
true point contacts exist between the grains, and the very but it is largely attributable to continued closure of the
concept of closure stress becomes inapplicable. It is re- subordinate number of ellipsoidal inclusions with a great-
placed by the loading-induced elastic-contact stiffening, er aspect ratio (α > 0.01) than commonly is attributed to
which eventually leads to the mechanical grain crushing. microcracks.
The role of “soft porosity” in such sediments is performed To capture these observations, Vernik and Hamman
by the irregular grain contacts that are prone to slippage (2009) have rearranged the exponential-rise function used
with possible grain rotation. As a result, stress-induced by Eberhard-Phillips (1989) and later rederived by
porosity reduction in unconsolidated sands can be much Shapiro (2003), as follows:
more noticeable (as great as 2 porosity units or 2%).
It is also important to realize that stress-relief-induced VPd = VPo + bP [σ − c exp(− dσ )],
(23)
and cooling-induced microcracks can be introduced into VSd = VSo + bS [σ − c exp(− dσ )],
more or less consolidated rocks during core recovery or,
to a lesser extent, in situ during uplift of geologic forma- where b, c, and d are the fitting parameters and sigma is
tions. These effects can dramatically alter the dry VP and isotropic effective stress. These fitting parameters can be
VS values measured in cores recovered from very deep identified with soft-inclusion aspect-ratio distribution,
wells (e.g., Vernik et al., 1994). crack density, and the inverse of the dominant aspect ratio,
As we saw in equation 7 of Chapter 1, an empirical respectively (Shapiro, 2003). VPo and VSo are microcrack-
model accounting for the exponential stress dependence free velocity intercepts obtained by extrapolating the quasi-
of sandstones was introduced by Eberhart-Phillips (1989). linear trends at high stresses back to the unstressed condi-
This model was later discussed by Shapiro (2003), who tions. Whereas this task is quite straightforward for con-
demonstrated that the model’s parameters actually may solidated sandstones, it is much less so for soft, high-porosity
have some physical meaning related to “soft inclusions” sands, which is a definite weakness of the model.
and their aspect ratios. Interestingly, there is a strong correlation between the
Vernik and Hamman (2009) have compiled several fitting parameters in equations 23 and mirocrack-free ve-
published data collections (Domenico, 1977; Han et al., locities (Vernik and Hamman, 2009), which allows fairly
1986), along with in-house data, on dry VP and VS varia- accurate predictions of the velocity-versus-stress behavior
tions with confining pressure, with the assumption that in arenites and clean arenites (i.e., sands with less than 12%
stress-anisotropy effects are negligible in deep reservoirs clay), especially in their consolidated varieties. Because the
located in extensional sedimentary basins (Chapter 2). To unstressed, microcrack-free velocities are initially un-
minimize the effect of core damage, they made a deliberate known, several iterations of equations 23 are required to
attempt to avoid data on low-porosity sandstone cores re- produce the velocity-stress curves consistent with the initial
covered from deep boreholes. Instead, the data set includes VPd and VSd values at known stress levels. These initial dry
relatively unweathered sandstone samples from quarries. velocities can be obtained from the Gassmann inversion of
This small but diverse data set spans a wide porosity VP and VS logs, if the water-saturation curve is available.
(and consolidation/cementation) range and includes one The question remains, though, whether a similar or
clean arenite (Ottawa sand) and four arenites (the Troika more accurate result could be achieved when one is using

03-Vernik_Ch03.indd 53 12-08-2016 20:02:10


54 Seismic Petrophysics in Quantitative Interpretation

theoretical equations instead of purely empirical formulas a)


such as that of Eberhard-Phillips (equation 7 of Chapter
1) and equations 23. To answer this question, let us resort
to the experimental data set of Vernik and Hamman (2009)
and equations 16 — the simplest outcome of the effec-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tive-field-theory application to the noninteraction approx-


imation. A heuristic exponential model can be introduced
to relate the crack density to stress (Vernik and Kachanov,
2010):

η = η0 exp(− dσ ), (24)

where η0 is the crack density at zero stress. Note that both


η0 and the exponent d in this relation must be derived
empirically from the data set.
Equations 16, in combination with equation 24, are b)
applied, using the matrix parameters of arenites, to all the
samples except Ottawa sand. For Ottawa sand, which is
almost 100% quartz, the parameters for clean arenite are
used (see Table 2 of Chapter 1). Figure 8a and 8b shows
dry VP and VS, respectively, for five different sands and
sandstones, superposed by the model computations opti-
mized to match the data. Somewhat surprisingly, the d-
exponent for these quite diverse sands/sandstones can be
kept constant at d = 0.07, so the only parameters that have
to be optimized are the shape factors and the zero-stress
crack density. Moreover, the latter appears to be the same
for both velocities (as expected), which additionally il-
lustrates the internal consistency of the model.
These fitting parameters are plotted against the VPd at
20 MPa as a reference velocity in Figures 9 and 10. Again, Figure 8. (a) VP and (b) VS values in dry sands and
very strong correlations are found, and the best-fitting sandstones, as a function of isotropic (confining) stress, and
functions are the power-law formulas included in the fig- superposed by the modified EF-model predictions (equations
ures. The absolute difference between p and q among the 16 and 24). Note that each data point is represented by a
five samples varies from 0.5 to 2.0, and the relative differ- filled circle corresponding to the accuracy of the P-wave
ence averages 7% from the mean values. Obviously, other velocity measurements at stresses greater than 5 MPa.
reference effective-stress levels could also be selected, Adapted from Figure 2 of Vernik and Hamman (2009). Used
with an expectation of somewhat altered fitting parame- by permission.
ters. The bottom line is that those parameters should only
be used as general guides in finding the best data-fitting The only difficulty with application of this semitheo-
velocity-versus-stress relationship by using equations 16 retical model can be observed for the velocities in Ottawa
and 24. sand at the initial and maximum stress levels of 5 MPa
For practical applications, it is recommended to de- and 40 MPa, respectively (Figure 8). However, it should
rive dry velocity curves from sonic logs in homogeneous be noted that this sand is totally unconsolidated and may
sand intervals and use them in combination with the ini- suffer from some initial grain crushing and rotation at the
tial effective-stress value to calibrate the model’s equa- maximum stress of 40 MPa. Additionally, the accuracy of
tions; that will be discussed later in this chapter in the the velocity measurements themselves at low stress nor-
section on 4D seismic feasibility modeling. In the process mally is compromised by ultrasonic wave scattering/at-
of calibration, preference should be given to the model tenuation that results in poor signal quality.
shown in Figure 10 for zero-stress crack density, and it- As was mentioned before, equations 16 are designed
erations should be performed for the pore-shape factors for solids with pores and cracks and in the strict sense can
only. This ensures that the functional form of the velocity only be used for consolidated rocks. The very concept of
dependence on stress is retained and the actual position of crack density in poorly consolidated sands is question-
the curve can be adjusted as necessary. able, so the model’s application quickly turns into an

03-Vernik_Ch03.indd 54 12-08-2016 20:02:11


Chapter 3: Seismic Rock Properties and Rock Physics 55

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b)
Figure 9.  Bulk moduli pore-shape factors p and shear
moduli shape factors q, obtained from fitting the modified EF
model to the core data shown in Figure 8, and plotted against
VP at a 20-MPa confining pressure.

Figure 11.  Velocity ratio in (a) dry and (b) low-frequency,


water-saturated sands and sandstones, as a function of
effective stress. The modified EF-model lines are superposed.
Gassmann’s equation is used for water-saturated predictions
in (b). Data are from Vernik and Hamman (2009).

plug-ins, can be quite serviceable in seismic petrophysics


applications and seismic interpretation in general.
Figure 10.  The crack-density parameter at zero stress
(obtained from fitting the modified EF model to the core data Values of VP /VS as a function of effective stress are
shown in Figure 8) versus VPd at 20 MPa. shown in Figure 11a for dry sands and sandstones, and in
Figure 11b for their low-frequency, water-saturated coun-
terparts. These plots use the same experimental data set and
empirical exercise when crack densities are allowed to the same model’s predictions. Gassmann’s equation is ap-
grow above 1.0 and ultimately to reach a totally unrealis- plied to “saturate” the rocks with 100-kppm-salinity brine
tic value of 15.0 in unconsolidated sands. (ρf = 1.05 g/cm3, Kf = 3.0 GPa). A nearly flat response, with
We will encounter these kinds of problems later, in a relatively minor stress-induced increase in VP /VS, from
the discussion on diagenetic trends in velocity-(and/or 1.5 to 1.55, is seen in the dry sands/sandstones. Two phe-
impedance)-versus-porosity and VP-VS spaces. Even the nomena account for this behavior: (1) the pore-shape fac-
most adequate theoretical models have limited applicabil- tors p and q are reasonably close to each other in arenites,
ity, and/or their parameters will need to be constrained by and (2) the crack-density term is the same for both the bulk
using some type of empirical relationship. Nonetheless, and shear moduli derived from fitting the model to the data.
as is illustrated in this section, this type of hybrid model, The saturated velocity ratio at low stress is quite high
using the solid theoretical functional form with empirical in unconsolidated sands but gradually decreases with

03-Vernik_Ch03.indd 55 12-08-2016 20:02:12


56 Seismic Petrophysics in Quantitative Interpretation

loading. Again, the magnitude of the effective-stress-in- Because the majority of siliciclastic reservoir rocks belong
duced reduction in VP / VS values for wet sands is con- to the petroelastic group of arenites, we should focus ini-
trolled by sand consolidation. tially on the semitheoretical modeling of this group of
Figure 11b once again helps us to better understand sands/sandstones. That undertaking will be supplemented
the mismatch between the data and the model predictions by more empirically inclined modeling of conventional
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

in Ottawa sand (on both ends of the applied-stress range). shale, which is defined as a microlaminated siliciclastic
As previously mentioned, that mismatch is likely to be rock with greater than 30 to 35% clay minerals in their the
related to the lower traveltime-picking accuracy at 5 MPa solid phase (Chapter 1). Theoretical approaches are hardly
and the onset of inelastic deformation in the manner of sustainable in conventional (inorganic) shales, for reasons
grain crushing at 40 MPa. Therefore, this could be a curi- that are presented later.
ous example of the model being better than the data it is
trying to emulate. Pore shapes and porosity thresholds in sands
Stress-dependence models for dry (or drained) elastic
stiffnesses and velocities of wave propagation can be Typical thin-section micrographs of consolidated are-
combined with Gassmann’s equation and the petrophysi- nites are shown in Figure 12. The total clay content in this
cal model of any clastic reservoir in question, to construct most ubiquitous petrophysical group of sandstones typi-
a software code designed to investigate the possible cally varies from 2 to 12% of the solid-rock matrix charac-
stress-variation effects in situ. Availability of this kind of terized by grain-supported texture (Chapter 1). The matrix
code is imperative in exploration and, especially, in 4D properties within this group are relatively unaffected by
feasibility studies and interpretations. variations in clay content (Vernik, 1998; Vernik and
Kachanov, 2012) and can be assumed to be constant. This
substantially simplifies further petroelastic modeling.
Rock-physics modeling in The initial intergranular porosity of these sandstones
sand/shale sequences is estimated by point counting to be in the range from 33
to 40%, which is by definition their critical-porosity range
Even though extensive literature exists on rock-phys- (Nur et al., 1991; Nur, 1992). The sample displayed in
ics modeling (e.g., Wyllie et al., 1956; Raymer et al., Figure 12b has a lower initial porosity (see also Figure 3
1980; Han et al., 1986; Krief et al., 1990; Greenberg and of Chapter 1) and is characterized by a more poorly sorted
Castagna, 1992; Berge et al., 1995; Dvorkin and Nur, grain texture. However, dramatic porosity variations
1996; Keys and Xu, 2002; Pride et al., 2004; Avseth et al., among the samples shown in Figure 12 — from 7.3 to
2005; Vernik and Kachanov, 2010), this relatively imma- 25.2% — are primarily attributed to the degree of me-
ture science remains quite controversial and subject to chanical compaction followed by quartz cementation. The
continuous research in both academia and the oil and gas cement volume varies widely, from 3 to 23%.
industry. Mechanical compaction should be treated as a diage-
In fact, no model can be viewed as being well estab- netic process involving compaction, grain rotation, and
lished and universally accepted, even though substantial partial crushing, as well as initial, minute pressure solu-
progress has been made in identifying the variables that tion at the grain contacts — the sites where stress is con-
control the dry frame properties of sands and sandstones. centrated and activation energy for solution/precipitation
Those variables include mineral composition, grain sort- reactions is elevated. This range of processes cannot be
ing, mechanical compaction, cementation, porosity, pore easily reproduced in a lab experiment using grain packs.
geometry, crack density, and effective stress. It is particu- Therefore, measurements conducted on rocks that have
larly difficult to model the evolution of microstructure — undergone a certain amount of diagenesis both in situ and
from unconsolidated sediment to tight, low-porosity in the laboratory are very useful.
sandstone — caused by geologic processes that can be It is interesting that the combined intergranular vol-
collectively referred to as diagenesis. Indeed, as Nur et al. ume (IGV) of porosity and cement for the sandstones in
(1991), Avseth et al. (2005), and Vernik and Kachanov Figure 12 lies in a relatively tight range, from 25 to 30%.
(2010, 2012) have pointed out, a single, continuum-me- This is consistent with the worldwide observations of
chanics-based theory cannot describe the entire range of Paxton et al. (2002), who conclude that the porosity value
sand and sandstone diagenesis. of 26% may be close to a lower limit resulting from me-
The primary goal here is to relate the effective elastic chanical compaction of moderately well-sorted sand-
properties of siliciclastic rocks to their composition and stones. That value is, incidentally, the IGV expected to
microstructure, which is a result of complex sedimentary occur in a rhombohedral pack of well-sorted grains.
and diagenetic processes that these rocks have been sub- The 25 to 30% IGV range, extended slightly down-
jected to during their their formation and burial history. ward to accommodate poorly sorted sands, is indicative of

03-Vernik_Ch03.indd 56 12-08-2016 20:02:12


Chapter 3: Seismic Rock Properties and Rock Physics 57

concept of the critical-porosity threshold was introduced


by Vernik and Kachanov (2010); the consolidation poros-
ity marks the transition band where the micromechanics
of effective elasticity, in principle, should change.
As commonly is observed in thin-section analysis,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

pore shapes in sandstones evolve from softer, concave ge-


ometries at higher porosities (Figure 12a) to stiffer, trian-
gular and rectangular shapes in lower-porosity rocks with
substantial amounts of quartz cement (Figure 12c). In up-
coming text this qualitative information is quantified in
terms of the pore-shape factor as a function of porosity.
Casting the pore-shape factor in terms of a more easily
obtained parameter, such as porosity, results in a signifi-
cant improvement in the theoretical model’s predictions
for the database of dry core measurements. A key out-
come of these observations is that porosity modeling in
sandstones cannot be adequately accomplished without
considering the changes in pore geometries that accom-
pany cementation-induced porosity reduction.
It should be mentioned that obtaining undisturbed
thin-section photomicrographs of poorly consolidated
sands recovered from boreholes is virtually impossible.
That leads to difficulties with visualizing pore, crack, and
grain-contact geometries and designing robust microme-
chanics-based models for situations in which φ is greater
than φcon. Nonetheless, one very important petrographic
observation can be made: Porosity of poorly consolidated
grain-supported sands (both clean arenites and arenites) is
largely determined by their grain sorting and packing co-
ordination; the latter is controlled by mechanical compac-
tion in the near-surface environment.
It is noteworthy that the grain-sorting and packing-
coordination parameters also determine the consolidation
porosity, which can be as low as 22% in poorly sorted or
tightly packed sands and as high as 30% in moderately to
well-sorted or loosely packed varieties. On the other
hand, in some circumstances the consolidation porosity
could also be controlled by the geothermal gradient — the
higher the temperature and the greater the pore-saturating
Figure 12.  Thin-section photomicrographs of the most
brine activity, the greater is the chance for the onset of
ubiquitous petrophysical group of consolidated sandstones
chemical diagenesis early in the burial history of the sedi-
worldwide (arenites), showing typical pore-shape variation
ment; that is, at higher porosity. Extreme situations occur
with porosity reduction, from cuspoidal to triangular/
rectangular geometries. (a) 25.2% porosity, 3% quartz experimentally when one is fusing glass beads (Berge
cement; (b) 17.4% porosity, 8% quartz cement; and (c) 7.3% et al., 1995) and in nature with high-porosity limestones
porosity, 23% quartz cement. Qo are quartz overgrowths, Ip (ooid grainstones). Both cases can be characterized by a
are intergranular pores, and Sp are secondary pores. Adapted consolidation porosity φcon exceeding 30%.
from Figure 1 of Vernik and Kachanov (2010). Used by Obviously, the critical porosity of grain-supported
permission. sands can also vary widely — commonly, from 30 to 45%
— and those values are capped by the porosity of the hex-
their consolidation porosity φcon, which is approximately agonal and cubic packing arrangements. This critical-po-
22 to 30% and is the porosity of grain-supported sands, rosity range in young sediments under negligible effective
such as clean arenites and arenites, attained at the end of stress and low temperature is consistent with the results of
the mechanical diagenesis phase and prior to the onset of Beard and Weyl (1973) for artificially mixed sand packs
noticeable quartz cementation. This addition to the with variable sorting. Also, as is mentioned in Chapter 1,

03-Vernik_Ch03.indd 57 12-08-2016 20:02:13


58 Seismic Petrophysics in Quantitative Interpretation

detrital clay particles should not be included when one is


estimating sorting parameters in rock-physics applica-
tions. Otherwise, confusion can be introduced due to the
competing effects of the mineral composition and textural
characteristics of clastic sediments.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Consolidated sandstones
At porosities below the consolidation porosity φcon
(i.e., below 22 to 30%, depending on grain sorting and
stress state), a sandstone can be described microstructur-
ally as a continuous material with pores and cracks in be-
tween cemented grains. The fundamental effective-field
formalism for computing frame moduli of consolidated
arenites and clean arenites is given by equations 16 — the
noninteraction approximation incorporating the Mori-
Tanaka scheme.
Figure 13.  Dry frame moduli Gd and Md = Kd + (4/3)Gd for
The model given by equations 16 can be explicitly eval-
arenites, at a high confining stress, superposed by the Mori-
uated for ellipsoidal shapes, as shown in Figure 5. However, Tanaka model lines for the shape factors p = q = 3.6 and 6.6.
pore shapes typically do not even approximately resemble This plot is used to empirically derive the shape factors’
ellipsoids (Figure 12). Importantly, a large proportion of dependence on porosity. Adapted from Figure 2 of Vernik
pores are concave. This factor has a strong effect on pore- (1998). Used by permission.
compliance contributions and, consequently, cannot be ig-
nored (see the analysis of Sevostianov et al., 2008, for 3D
geometries). The challenge emerges of estimating the shape though a minor tendency can be observed for p to be
factors of such pores from available empirical data. slightly smaller than q. Note that p and q are substan-
As we already know, realistic pore-shape factors should tially greater than two, which confirms that concave
be larger than those for a sphere because a sphere is the pores are much more compliant than are spherical
stiffest shape of a given volume or porosity (Figures 4 and ones of the same volume fraction.
5). Indeed, typical pore shapes in sandstones tend to be • The crack density η (σ ) is a function of the effective
more concave than spherical (Figure 4). Vernik and stress σ; η decreases with increasing stress because
Kachanov (2010) have demonstrated that gradually increas- there is progressive closure of compliant microcracks.
ing the shape factors of pores in sandstones helps predict The form of this function is quite sensitive to crack
their effective moduli from the Mori-Tanaka scheme. geometries. A crack of the simplest idealized geome-
Indeed, the micromechanics-based equations 16 need try — a circular or elliptical in-plane crack with an
to be supplemented by empirical relationships to account ellipsoidal cross-section — closes at once as the criti-
for the following two factors that cannot be derived cal stress (which depends on the aspect ratio) is
analytically: reached. However, even this simplest case requires
information on the distribution of aspect ratios. In re-
• The shapes of pores observed in arenites depend ality, crack geometries are much more complex, and
on quartz-cementation-induced porosity reduction, as that factor may strongly affect the closure conditions.
Figure 12 illustrates (the evolution of the shapes as a In particular, planar cracks that have “irregular” in-
function of diagenesis in the framework of noninter- plane shapes, as well as details of near-tip microgeom-
action approximation has been discussed by Vernik, etries, may cause microcracks to close gradually rath-
1998; see also Figure 13). This dependence can be ap- er than stepwise (Mavko and Nur, 1978). Nonplanarity
proximated by the following empirical expression for of cracks and their intersections further complicates
porosities smaller than the consolidation porosity φcon: the matter. These factors necessitate relations of an
empirical nature for η (σ ) that provide a good fit to
p ≅ q = 3.6 + bφ , (25) the data. Vernik and Kachanov (2010) used the em-
pirical form given by equation 24 for the stress de-
where the slope b can vary from 8 to 12. It should be pendence of crack density that closely approximates
emphasized that for arenites p and q are very close to the stress-sensitivity model of Vernik and Hamman
each other for all practical purposes — that is, pore (2009) in the stress range from 5 to 50 MPa. This ex-
shapes affect Kd and Gd in a similar manner, even pression is of foremost interest in quantitative

03-Vernik_Ch03.indd 58 12-08-2016 20:02:14


Chapter 3: Seismic Rock Properties and Rock Physics 59

interpretation, and it is repeated here to emphasize its


importance:

η (σ ) = η0 exp(− dσ ), (26)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

where η0 = 0.3 + 1.6 φ and d = 0.07 are the crack


density at zero stress and the compaction coefficient
(in MPa−1), respectively, and the stress σ is in MPa.

Combining equations 16 with 25 and 26 results in the


following expressions, which in this book are referred to
as the Consolidated Sandstone Model:
−1
 pφ η exp( −dσ ) 
K dry = K m 1 + + A (ν m ) 0  ,
 1 − φ 1−φ 
−1
(27)
 qφ η0 exp( −dσ ) 
G dry = G m 1 + + B (ν m )  ,
 1−φ 1−φ 
where p = 3.6 + b φ , η0 = 0.3 + 1.6 φ , and d = 0.07 , with
Figure 14.  Dry core P-wave modulus versus porosity
the shape factors p = q when φ > 0.03; these terms must
for three levels of effective stress (10, 20, and 40 to
be calculated from porosity by using empirically derived 70 MPa), measured on 50 arenites from the core database
b values in the range from 8 to 12. The zero-stress crack and superposed by respective realizations of the
density and d value are also empirically selected to fit our Consolidated Sandstone Model parameterized for pore-shape
database, for which we only have to account for micro- factors and crack densities, as shown in the plot. Adapted
cracks. It should be mentioned, however, that for in situ from Figure 3 of Vernik and Kachanov (2010). Used by
logging-data analysis the empirical plug-ins may require permission.
minor adjustment.
The matrix coefficients A(νm) and B(νm) in equations
27 are given by the fractions at the crack density terms in
equations 15 and equal 2.46 and 1.59, respectively, for aren-
ites with the matrix properties νm = 0.146, Gm = 33.0 GPa,
and Km = 35.6 GPa (computed from Table 2 of Chapter 1).
Finally, as was discussed in Chapter 1, there is practically no
need to account for minor clay content variations in this
group of sandstones. That substantially simplifies the model
realization and makes it applicable worldwide, although the
presence of exotic minerals (e.g., zeolites) in arenites still
must be taken into account.
Figure 14 illustrates the performance of the Consol­
idated Sandstone Model against a subset of 50 arenites
from our database of core measurements in the range of
effective-stress values from 10 to 70 MPa. The root-
mean-square (rms) error of the P-wave modulus Md pre-
diction by the model is approximately 3%, which is a con-
vincing manifestation of its predictive power from the use
of porosity and stress simultaneously.
Figure 15 plots Md and Gd of these dry arenites mea-
sured at 20 MPa (2900 psi) confining stress together with Figure 15.  Dry moduli Md and Gd versus porosity,
the Consolidated Sandstone Model lines for 10, 20, and measured at 20 MPa in arenites from the core database
50 MPa, using the same input parameters as described in (a subset of Figure 14), superposed here by the Consolidated
the text below expressions 27. We can see that both moduli Sandstone Model lines for 10, 20, and 50 MPa and using
are predicted with a high degree of confidence, thereby con- the input parameters for equations 27 as described in the
firming that the minor theoretical difference between p and text. Adapted from Figure 4 of Vernik and Kachanov (2010).
q can be comfortably ignored in practical applications. Used by permission.

03-Vernik_Ch03.indd 59 12-08-2016 20:02:15


60 Seismic Petrophysics in Quantitative Interpretation

It is instructive at this point to bring up a controver- The upper and lower modified HS curves are often
sial issue in rock physics — heuristic modification of the referred to as “stiff” and “soft” sand models, respectively
theoretical elasticity bounds of Hashin and Shtrikman (Mavko et al., 2009). It is also claimed that the upper
(1963). The modification was designed (Dvorkin and Nur, modified HS model provides a trend that is useful in diag-
1996) to fit the experimental data on velocity and porosity nosing sandstone diagenesis in moduli-versus-porosity
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

better by honoring the critical-porosity concept intro- space (Avseth et al., 2010). With this in mind, let us ex-
duced by Nur (1992) — that is, by replacing the porosity amine Figure 16, which plots the same subset of consoli-
and fluid-modulus terms in the Hashin and Shtrikman dated arenites that were measured at 20 MPa and included
equations with φc /φc , where dry moduli Kc and Gc refer to
the dry moduli at critical porosity. The equations for dry a)
effective moduli originating from the upper HS bound are
(Gal et al., 1998):

−1
+  φ /φc 1 − φ /φc  4
K HSM =  +  − G ,
K
 c + ( 4 / 3)Gm K m + ( 4 / 3)Gm  3 m
−1
+  φ /φc 1 − φ /φc 
GHSM =  +  − zm , (28)
 Gc + zm Gm + zm 

Gm  9 K m + 8 Gm 
where zm = .
6  K m + 2 Gm 
Similarly, the equations for dry effective moduli orig-
inating from the lower HS bound are (Dvorkin and Nur,
1996):

−1
−  φ /φc 1 − φ /φc  4
K HSM =  +  − G, b)
K
 c + ( 4 / 3)Gc K m + ( 4 / 3)Gc  3 c
−1
−  φ /φc 1 − φ /φc 
GHSM =  +  − zc ,
 Gc + zc Gm + zc 
(29)
Gc  9 Kc + 8Gc 
where zc = .
6  Kc + 2 Gc 
It was previously mentioned that the critical porosity
of sand can vary from 30 to 45%, depending on its sort-
ing, grain shapes, and packing; beyond those porosities
the sand/fluid mixture turns into a suspension. Some vari-
ations of these models include an end member, which is
selected at a porosity lower than the critical porosity (e.g.,
Avseth et al., 2005).
As Vernik and Kachanov (2012) pointed out, the
“modified HS” models simply change the argument of the
curves from φ to the ratio φ /φc , in order to enforce the Figure 16.  (a) Dry P-wave and shear moduli versus porosity
cutoff at a selected high-porosity end member. It should in the same set of consolidated arenites measured at 20 MPa
be emphasized that these replacements dramatically dis- and shown in Figure 15, here superposed by the modified HS
tort the shape of the HS curves so that they no longer can bounds optimized to constrain the data; (b) the same data set
be considered elastic bounds and should be treated as overlain by the modified HS upper bound (“stiff sand”
semiempirical model lines. Moreover, the end-member model) optimized to match the medium to high-porosity
moduli in these models are either arbitrarily selected or sandstones. The end-member porosity in both plots is 40%.
computed using the Hertz-Mindlin contact theory, there- Adapted from Figure 4 of Vernik and Kachanov (2012). Used
by inheriting all its aforementioned deficiencies. by permission.

03-Vernik_Ch03.indd 60 12-08-2016 20:02:16


Chapter 3: Seismic Rock Properties and Rock Physics 61

in Figure 15. This time, however, the modified HS models Sandstone Model incorporates the crack-density factor and
are superposed instead. The end-member porosity is se- renders it to be stress-dependent simply to reflect the pro-
lected at 40%, whereas the low-end moduli are empiri- cess of crack closure with increasing stress. The resulting
cally optimized as shown in the plots. Some instructive flexibility of the model given by equation 27, which can be
observations can be made: adjusted to any stress loading that operates in sedimentary
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

basins worldwide, is what makes it particularly attractive.


• First, Figure 16a confirms that the HS-bound modifi- Two additional remarks should be made with respect
cation with optimized end-member moduli can be to crack effects:
somewhat useful for constraining the data in the en-
tire porosity range that is typical of consolidated • Once the aspect ratio of a crack falls below approxi-
sandstones, even though these tighter bounds are not mately 0.10, the crack’s elastic response to small
strictly adhered to even by this limited subset of data. strains does not depend on its aspect ratio; thus, crack
• Second, as follows from Figure 16b, no moduli opti- porosity is not a relevant parameter.
mization or critical-porosity selection is capable of • Linear elastic crack compliances relevant for dy­
producing an adequately good fit to the data. The namic wave propagation, unlike nonlinear effects in
model consistently overestimates the moduli when stress-strain tests, are insensitive to the exact mi-
porosity is less than 20%. crogeometry of the near-tip geometry of the crack.
• Third, comparing Figure 15 and Figure 16b, the good-
ness of fit over the entire porosity range provided by Unconsolidated and poorly consolidated sands
the Consolidated Sandstone Model given by equations
27 is visibly superior in its predictive power to the At higher porosities, specifically from the consolida-
modified HS-upper-bound model. The main reason tion porosity φcon to the critical porosity φc, the amount of
for the mismatch lies in the fact that the latter model intergranular cement in grain-supported sands is lacking or
does not take into account microcracks, which are still very insignificant. Therefore, the rock or sediment can be
partly open at the moderate stress level of 20 MPa. treated as a granular material rather than as a cohesive solid
These microcracks have a negligible contribution to containing pores and cracks. Of course, this does not apply
the total porosity, but they significantly reduce effec- to fused glass-bead packs or carbonate reservoir rocks,
tive moduli even in low-porosity sandstones (Figures which typically remain cohesive even at higher porosities.
14 and 15; also see Smith et al., 2009). That cohesiveness is the result of chemical grain bonding
and of lower thermal-stress effects on the carbonate-miner-
The Consolidated Sandstone Model matches labora- al-grain contacts, compared with the effects on quartz.
tory measurements accurately in the entire range of effec- Therefore, the micromechanics of effective elasticity
tive stresses that are of interest, primarily because it in poorly consolidated arenites and clean arenites is differ-
­accounts for the main factors controlling the elasticity of ent from that in consolidated sandstones — it is controlled
sandstones undergoing chemical diagenesis — their by the statistics and evolution of grain contacts, as a result
porosity, prevalent pore shapes, and grain-boundary
­ of mechanical adjustment, grain rotation, and initial, min-
microcracks. ute pressure solution between quartz and/or feldspar
The process of chemical diagenesis largely manifests grains. These processes can be collectively referred to as
itself in gradually reducing porosity from the consolida- mechanical diagenesis and are reflected in changes in the
tion porosity φcon to very low values. By far the most im- slope of wave-velocity-versus-porosity curves at the con-
portant mechanism operating during chemical diagenesis solidation porosity (Vernik, 1998) — a phenomenon that
is quartz cementation, which can cause pore geometries will be extensively reviewed later in this chapter.
to change from highly concave or hypotrochoidal at po- The effective elastic properties of such materials have
rosities from 26 to 14%, to triangular or rectangular at been frequently modeled in the framework of the Hertz-
porosities below 14%. The porosity-controlled shape fac- Mindlin theory (e.g., Avseth et al., 2005; Mavko et al.,
tor plugged into the Mori-Tanaka effective-field model 2009). However, to describe laboratory and sonic-logging
(equations 27) tends to account for this effect. data satisfactorily, this approach requires the introduction
Another factor impacting elastic moduli of consolidat- of parameters such as the extent of sliding between grains
ed sandstones is the microcrack density, which is a­ ffected and the assumption of the markedly varying radii ratio of
by the current stress state and the stress and thermal histo- the grains in contact (Bachrach and Avseth, 2008). Such
ries of the rock. For example, some of the cracks originate parameters are essentially unknowns, so they play the role
during stress release and cooling during core recovery or of fitting parameters rather than micromechanical ones.
during geologic processes such as uplift and erosion Note also that to obtain dry compressional velocities
(Vernik et al., 1994; Holt et al., 2005). The Consolidated that exceed 2.5 km/s in poorly consolidated arenites,

03-Vernik_Ch03.indd 61 12-08-2016 20:02:16


62 Seismic Petrophysics in Quantitative Interpretation

which are not 100% quartz sands, the assumed grain-co- depth intervals. The match can be optimized by using the
ordination number must be unrealistically high. Finally, coefficients p and q either as fitting parameters or as those
as is illustrated in Figure 7, the stress dependence pre- based on an empirical formula (presented later in this sec-
dicted by contact theories is not supported by multiple tion) relating them to the effective stress and, hence, to
laboratory measurements. Therefore, in order to describe the level of mechanical compaction.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

the early, mechanical-diagenesis trend, a semiempirical Therefore, despite their formal theoretical basis,
approach can be suggested that is consistent with the reli- equations 31, when applied to poorly consolidated sands,
able core and log data. should be viewed as a heuristic curve-fitting procedure
The following simple relationships that connect two rather than as a micromechanics-based approach. One can
end data points, φcon and φc, provide a reasonable fit to the speculate, however, that at the very low strain amplitudes
effective moduli Gd and Md as functions of porosity that are typical of seismic waves, even poorly consoli-
(Vernik and Kachanov, 2010): dated sands remain virtually cohesive and thereby may
justify application of the Mori-Tanaka scheme.
φ − φcon 
n Note also that equation 31 in combination with the

M d = M con 1 − , Gassmann equation (simplified by equating the Biot coeffi-
 φc − φcon 
(30) cient to unity) and the familiar expression for bulk density as
m
 φ − φcon  a function of porosity (Chapter 1, equation 33) can be used
Gd = Gcon 1 − ,
 φc − φcon  to derive the density versus velocity template for wet, poorly
consolidated sands. Such a template should have several
model lines with varying pore shape factor, which accounts
where, in order to provide a continuous transition to the
for variations in compaction and/or effective stress.
consolidated regime, the values of Mcon and Gcon are taken
Realizations of this model for the shape factor ranging from
as predicted by equations 27 at φ = φcon , and n and m are
15.0–25.0 are shown in Figure 24b of Chapter 1.
the empirical exponents. It can be found that the selection
The virtual cohesiveness of poorly consolidated (but
of n = 2.00 and m = 2.05 provides a reasonable fit to (1)
not entirely unconsolidated) sands is consistent with the no-
the naturally occurring trends of mechanical compaction-
tion that they have been subjected to the initial pressure so-
dominated sand diagenesis, and (2) a slight increase in
lution and that that exposure has resulted in laterally small,
dry VP/VS values with increasing porosity, for poorly con-
but nonetheless “non-Hertzian” contacts between the grains
solidated samples. It should be emphasized that applica-
that can be easily broken during coring and core recovery.
tion of this model should be limited to the φcon -φc range.
Therefore, equations 31 can trace grain-sorting variations
Here, it is referred to as Soft Sand Model 1.
where only minimal variations in pore-shape factors are ex-
Equations of similar structural form, but for the entire
pected as a result of a relatively constant stress state and
porosity range from zero to φc, were introduced by Nur et al.
mechanical diagenesis within limited stratigraphic inter-
(1991). Note, however, that use of equations 30 in forward
vals. Equations 31 are referred to as Soft Sand Model 2.
modeling requires knowledge of two values, φc and φcon, that
To check the performance of this two-prong modeling
are hard to predict accurately using log data alone. Additional
approach, the core-based data set has been augmented by
geologic information on sorting and/or geothermal gradi-
another one representing a mixture of log-derived interval
ent is necessary to constrain these parameters.
averages with core and/or sidewall-core porosity measure-
To aid us in modeling the effects of porosity variation
ments. This methodology was successfully tested by Issler
on elastic moduli and velocities in poorly consolidated
(1992) for conventional shales. The log data from deepwa-
sands, the following Mori-Tanaka-scheme-inspired rela-
ter Gulf of Mexico and West Africa have been processed to
tions can be proposed (Vernik and Kachanov, 2010):
extract information on 50 relatively homogeneous, 5- to
−1
15-m sand intervals covering a porosity range from 9 to
 φ  38% and clay content vcl values between 2 and 10%. Both
Md = Mm 1 + p ,
 1 − φ  monopole compressional-velocity and dipole shear-veloc-
−1
(31) ity data were available for each interval, as well as core-
 φ  calibrated log-based porosity, saturation, and clay-content
Gd = Gm 1 + q .
 1 − φ  data; the dry frame moduli and velocities for each data

point were derived using arenite matrix properties (Table 2
Although the Mori-Tanaka scheme does not apply theo- of Chapter 1) and Gassmann inversion.
retically to granular materials, equations 31 (derived from Wireline pressure measurements in reservoir-quality
the more general equations 27 by omitting the crack-den- rocks have been supplemented with pore-pressure pro-
sity term) can quite accurately fit log data that are limited files computed by following the workflow described in
to specified stratigraphic units encountered in narrow Chapter 2 and allowing the vertical-effective-stress

03-Vernik_Ch03.indd 62 12-08-2016 20:02:16


Chapter 3: Seismic Rock Properties and Rock Physics 63
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 18.  Comparison of dry P-wave modulus M d -φ


Figure 17.  Dry compressional-wave velocity versus porosity models; the curves were obtained from the modified lower
for a data set of 50 core-calibrated log intervals of consolidated HS bound (showing a range of endpoint moduli from 16.0 to
and poorly consolidated arenites with a reasonably well- 4.2 GPa at a critical porosity of 40%); and Soft Sand Model 2
defined consolidation porosity of approximately 23 to 25%, results (showing a matching range of pore-shape factors from
from 12 deep wells drilled in deepwater West Africa and the 6.0 to 24.0). The rms difference between these two models in
Gulf of Mexico, and superposed by Soft Sand Model 2 lines this range is less than 2%. Adapted from Figure 5 of Vernik
with pore-shape p factors varying from 5 to 45. Adapted from and Kachanov (2012).
Figure 2 of Vernik and Kachanov (2010).
Figure 18 compares the dry P-wave modulus Md
evaluation from the log velocity and density data in asso- curves obtained from the modified lower HS bound for a
ciated shales. Again, due to limited stress anisotropy in wide range of end-point moduli Mc (from 16.0 to 4.2 GPa)
the provinces mentioned, the vertical effective stress is at a critical porosity φc of 40%, and the single value of
used as a proxy for the mean effective stress. Mm = Km + (4/3)Gm at zero porosity (Table 2 of Chapter
Figure 17 plots dry P-wave velocity versus porosity 1). As follows from the plot, these curves can be quite ac-
for the log data set. Note that this data set is composed of curately matched by Soft Sand Model 2 with varying
consolidated and poorly consolidated arenites with a rea- shape factors p. Again, an assumption of shape-factor
sonably well-defined consolidation porosity of approxi- equality is made, so that p = q. The p-factor is allowed to
mately 23 to 25%. These values are 5 to 7 porosity units vary to accommodate the variations in effective stress.
below the characteristic value of the core data set, which Note that in order to avoid lithology mixing effects in
on the high-porosity end is dominated by samples from seismic petrophysics applications, this model should be
the North Sea province (see Figures 13 and 14). applied to relatively homogeneous sand intervals that are
Interestingly, a significant number of well-consolidat- free of laminar shale and/or silt with an elevated clay con-
ed arenites from the log data set shown in Figure 17 (e.g., tent — both of which can be mistaken for variations in the
data points with a porosity of less than 20%) can be de- sand-sorting parameter.
scribed by our Soft Sand Model 2 (equations 31), with a The entire core/log data set of poorly consolidated
pore-shape factor of 7.0 ± 1.0. Three well consolidated arenites at low to intermediate stress levels of 10 to
sandstones with medium porosities (from 15 to 18%) and 26 MPa is shown in Figure 19. The rock-physics template
a pore-shape factor greater than 10.0 come from the dis- including Soft Sand Models 1 and 2 is overlain, thereby
tinctly overpressured intervals where the vertical effective allowing the user to (1) diagnose the main controls on the
stress is below 13 MPa. Another striking feature gleaned elastic properties of the sands and (2) visualize distinct
from this crossplot is that even though the data cloud cuts vectors of variations in their elastic properties. Both soft-
across the model lines, some data clusters are aligned sand models are included in this template to enable visu-
along those lines. For example, the softest cluster with VPd alization of the two distinct vectors of variations in prop-
at or below 1.7 km/s, which came from some very shallow erties. Soft Sand Model 1 is realized using three end-
wells with unconsolidated sands, is almost perfectly member pairs of φcon and φc: 22 to 34%, 26 to 38%, and 30
matched by the model lines with a shape factor p of 45. to 42%. Four curves of Soft Sand Model 2, representing

03-Vernik_Ch03.indd 63 12-08-2016 20:02:18


64 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 19.  The rock-physics template mapping the modulus-


versus-porosity space for poorly consolidated arenites, with Figure 20.  Crossplot of pore-shape factors p and q, inverted
curves from two different soft sand models parameterized as by using Soft Sand Model 2, from compressional and dipole
shown. Core data at 10 and 20 MPa and log/core data at an shear velocities for the data set of 50 processed log intervals
effective stress ranging from 12 to 25 MPa are used. of arenites. Adapted from Figure 7 of Vernik and Kachanov
(2010). Used by permission.
pore-shape factors p of 10, 15, 25, and 45, cross the grain
of model 1 at oblique angles. reduced-major-axes (RMA) slope of 0.98. Therefore, at
Figure 19 illustrates how this rock-physics template least in poorly consolidated sands, for all practical pur-
maps the modulus-versus-porosity space, and it allows the poses we can assume that p ≈ q.
interpreter to understand the reason for the large amount of Moreover, as follows from Figure 21, the pore-shape
natural scatter in the data, with no clear trends. The model 1 factors can be inversely related to the current vertical ef-
lines are steeper and correspond to sand mechanical-consol- fective stress by using p = Aσ − B , where σ is given in
idation trends, whereas the model 2 lines have gentler slopes MPa and the parameters (A and B) are affected by the
that increase with the decreasing pore-shape factors. level of elastic unloading that is due to overpressure or
These gentler model lines strike in the direction of the uplift/erosion events late in the diagenetic history. This
grain-sorting effect, with each of them corresponding to the finding allows us to use equations 31 heuristically in for-
same level of mechanical compaction. As previously men- ward modeling to account for variations in porosity, grain
tioned, an excellent alignment of clusters of poorly consoli- sorting, and effective stress simultaneously, which is quite
dated arenite data points along the constant p lines is quite useful in petroelastic model building.
obvious (especially for p = 45), which is an explicit mani- To reiterate, the Mori-Tanaka scheme, like any other
festation of the grain-sorting effect. Consequently, the com- EF or EM theory, is strictly designed to handle solids with
bination of equations 30 and 31 — that is, of Soft Sand pores. The fact that it can have such a predictive power
Models 1 and 2 — completely accounts for the data scatter in poorly consolidated clastics (Soft Sand Model 2)
in sonic-velocity-versus-porosity plots at relatively high po- ­indicates — albeit tentatively at this point — that the rela-
rosities. This allows us to model both local grain-sorting- tively clean, grain-supported sand reservoirs typically en-
related porosity variations within specified stratigraphic in- countered in hydrocarbon exploration are not necessarily
tervals and global trends of sand mechanical compaction. characterized by mere point, or so-called Hertzian, con-
Because the statistics for porosity and dry elastic tacts between the grains. This is especially true in more
moduli were available for the represented log intervals, advanced stages of mechanical diagenesis when pressure-
equations 31 can be used to invert for the pore-shape fac- solution mechanisms start to act with or without initial
tors p and q and to crossplot the results in Figure 20 for quartz cementation.
the entire core/log data set. Even though the linear depen-
dence of these parameters on porosity differs slightly be- Sandstone diagenesis models
tween p and q among consolidated sandstones (with shape
factors smaller than 12), the entire porosity range displays In frontier exploration, it is often important to predict
a very tight correlation between p and q, with a the porosity of sands and sandstones by using very

03-Vernik_Ch03.indd 64 12-08-2016 20:02:19


Chapter 3: Seismic Rock Properties and Rock Physics 65
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 22.  Dry compressional velocity versus porosity for


Figure 21.  Pore-shape factors versus vertical effective stress the core data set at 40 to 70 MPa, superposed by the Sand
estimated for each of the 50 processed log intervals of Diagenesis Model parameterized with σ = 50 MPa, φc = 0.40,
arenites. The exponential models shown are likely to reflect and φcon = 0.30. Adapted from Figure 5 of Vernik and
different amounts of stress unloading that was due to Kachanov (2010). Used by permission.
overpressure or uplift/erosion history (the dotted curve
matches medium-porosity sandstones with hard respective core and log data sets of arenites discussed pre-
overpressure). Adapted from Figure 8 of Vernik and viously. Depending on context, we will refer to those
Kachanov (2010). Used by permission. merged lines as Sandstone Diagenesis or Sand Diagenesis
Models. Notice the distinct “knee” occurrence around
limited resources (e.g., seismic velocities). It is even more φcon = 0.30 for the core data set in Figure 22, which is
critical to be able to predict shear-wave velocity accurate- dominated by the moderately to well-sorted sandstones
ly from scarce information on VP and within certain lithol- from the North Sea at the higher-porosity end. A similar
ogy constraints. location of the knee along the porosity axis is also implied
Building sand/sandstone- and shale-diagenesis trend by the data presented by Berge et al. (1995) on synthetic
lines in rock-physics crossplots is quite helpful in these (also well-sorted) sandstones and fused glass beads.
situations. In the process, we must be able to properly iden- To match the core-calibrated log data set dominated
tify the two main stages of diagenetic alteration of the orig- by deepwater turbidites from the Gulf of Mexico and
inally deposited clastic sediment: mechanical diagenesis West Africa (a poorly consolidated subset already dis-
and chemical diagenesis. The former alters physical prop- played in Figure 17), an effective stress of 20 MPa had to
erties of unconsolidated and poorly consolidated sands in be used in equations 27, and a lower consolidation poros-
which φ exceeds φcon, whereas the l­ atter deals with consoli- ity of 23% in equations 30 had to be chosen (Figure 23).
dated sandstones in which φ is smaller than φcon. Note that whereas only a high stress range (of 40 to
Because we have already established the building 70 MPa) is selected for the ultrasonic core data in Figure
blocks for these respective stages, all it takes now is to 22, the sonic-log intervals included in Figure 23 are sub-
merge them together. Vernik and Kachanov (2010) sug- jected to a much lower average in situ effective stress. In
gest that φcon be the merge point, with the realization that combination with the smaller φcon values in predominantly
this threshold porosity can vary depending on the grain deepwater turbidites with poor sorting characteristics,
sorting, grain shapes, and packing geometry of the rock in this results in the less-distinct knee and more ambiguous
question. The equations to be merged are 27 (the interpretation of φcon in the log-data-driven realization of
Consolidated Sandstone Model) and 30 (the Soft Sand the Sand Diagenesis Model.
Model 1) in that specific order, because the consolidation The difference in the consolidation porosity between
moduli entering equations 30 are derived from equations the North Sea and deepwater GOM data sets could also be
27 at the consolidation porosity. a result of the difference in geothermal gradients, which
Figures 22 and 23 show two realizations of the result- are significantly greater in the North Sea province. Under
ing sandstone-diagenesis-trend lines, superposed on the those circumstances, the onset of chemical diagenesis can

03-Vernik_Ch03.indd 65 12-08-2016 20:02:20


66 Seismic Petrophysics in Quantitative Interpretation

shales, the matrix of which is usually always fully water-


saturated as a result of extremely high capillary pressure.
For that reason, conventional shales typically do not
require any fluid-substitution modeling. This simplifies
our task of developing consistent velocity-versus-porosi-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ty relationships for shales, at least for the bedding-normal


direction of wave propagation, because their matrix prop-
erties are generally anisotropic. Note that conventional
shales may be leaky, most notably when they overlie over-
pressured reservoirs. Such shales may have fractured
zones with some degree of gas saturation, which can have
an effect on their elastic properties measured in situ (e.g.,
Chopra and Castagna, 2014).
Although a micromechanics-based approach can, in
principle, be applied to shale, it does not seem promising,
for several reasons: (1) shale is notoriously anisotropic,
which greatly complicates the models, (2) model formu-
lation in terms of dry elastic moduli will necessitate
Gassmann fluid substitution in these commonly “non-
Figure 23.  Dry compressional velocity versus porosity for
the log data set shown in Figure 17, but superposed by the
Gassmann” rocks, (3) some clay minerals in shales are
Sand Diagenesis Model parameterized with σ = 20 MPa, subjected to significant rock-water interaction, which
φc = 0.40, and φcon = 0.23. Adapted from Figure 6 of Vernik cannot be modeled by using purely mechanical concepts,
and Kachanov (2010). Used by permission. and (4) ellipsoidal inclusion-based models can be gener-
ally misleading on shale microstructure.
As will be shown below, ultrasonic core and wireline
occur earlier and at shallower depths of burial and higher sonic measurements in bedding-normal directions (or near-
porosities. This interesting observation may warrant more ly normal directions) suggest that wave velocities in con-
research in the interdisciplinary field that combines pe- ventional shales are more stable, are monotonic (at least
trology, rock physics, and basin modeling. compared with arenites), and are slightly nonlinear with
As was mentioned before, a few log-data points in respect to porosity (for porosities less than 35 to 40%).
Figure 23 that originate from the overpressured intervals Therefore, the focus should be on developing a universal
with φ values between 15 and 20% lie well below the con- empirical model for fully water-saturated shale, subject to
solidated-model-line predictions for the average stress of two requirements: (1) the model should contain only ob-
20 MPa. These intervals have been subjected to greatly servable and quantifiable parameters, such as porosity and
reduced effective-stress levels of 10 to 14 MPa (e.g., clay content, and (2) it should be immune from geographic
Figure 10 of Chapter 2). It is obvious that the Sand peculiarities, so that it can be applicable worldwide.
Diagenesis Model is not unique and that all the variables The following model for the anisotropic elastic constant
in its constituent building blocks must be properly adjust- governing compressional-wave velocity in the bedding-nor-
ed to obtain accurate predictions, even for application to a mal direction was proposed by Vernik and Kachanov (2010):
single lithology such as the group of arenites. Nonetheless,
the “hockey stick” appearance of these trends holds in the C33 = C33 m (1 − φ )k * , (32)
majority of settings, except in the relatively low-consoli-
dation-porosity case — that is, when the deflection ap- where the matrix stiffness C33m in this direction is com-
pears to be more subtle and the transition from sands to puted by using either lower bound Hashin-Shtrikman or
sandstones in VP -φ space is more gradual. Reuss averaging over clay and nonclay volu­metric compo-
nents, as shown in equation 12 of Chapter 2, and the ex-
Conventional shales ponent k* can be termed the “pseudo shape factor”, which
is empirically related to clay content:
In the context of hydrocarbon exploration, shales have
historically been considered to be nonreservoir lithologies
k * = 5.3 − 1.3ν cl . (33)
and seals. Recently, however, organic-rich shales have be-
come the topic of intense interest because of their ubiqui- Finally, the P-wave velocity normal to bedding is
tous nature and associated hydrocarbons (both oil and computed by using equation 11 of Chapter 2. As was also
gas). In this section, we limit ourselves to conventional discussed in Chapter 2, the solid-grain P-wave modulus

03-Vernik_Ch03.indd 66 12-08-2016 20:02:21


Chapter 3: Seismic Rock Properties and Rock Physics 67

of the nonclay phase can be approximated as that of coincides with the interval velocity from the vertical seis-
quartz, whereas C33cl can be taken from Table 2 of Chapter mic profiling (VSP). This is illustrated in Figure 24,
1. These relationships ensure that the bedding-normal VP which compares log and VSP data from part of a well in
predictions converge on the velocity of mud at a porosity West Africa. Apparently the critical frequency for shales
of approximately 40% and match experimental and log is much greater than that used in sonic logging, so little
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

data, as is demonstrated in the following section. frequency dispersion is expected in these lithologies;
To account for the gradual smectite/illite transforma- note, however, that this may not necessarily be the case in
tion, it is assumed that the shale-matrix density also grad- associated sands.
ually varies with compaction and the smectite-to-illite Figure 25 is a VP- and VS-versus-porosity plot show-
transition (see equation 8 of Chapter 2) and averages ing the shale-model lines on the background of the subset
2.73 ± 0.03g/cm3 at porosities below 25%, for all shales, of shale cores measured at 20 to 40 MPa in the direction
independently of clay content. This grain-density model normal to bedding and under fully water-saturated condi-
can be applied to shales with a bulk density in excess of tions. The data set is subdivided into two subsets: (1)
1.9 g/cm3. shales with vcl values mostly below 45% (red), or rela-
Note that the shale model given by the above formu- tively low-clay-content shales, and (2) shales with vcl val-
las is similar in structure to that derived from the DEM ues mostly greater than 45% (blue). An excellent match to
scheme by Keys and Xu (2002), with two key distinc- the experimental data is seen for both VP and VS. The
tions. (1) It does not relate the (pseudo) pore-shape factor model lines approach suspension P-wave velocity values
to the (ellipsoidal) inclusion aspect ratio α, which, as we at a porosity approaching 40%, and at that porosity the
already know, can have an uncertain bearing on the linear S-wave velocity drops sharply to zero. From this display
elastic constants. Instead, it is derived from clay content, it follows that conventional shales may, quite surprising-
which can be determined from a core-calibrated petro- ly, have a critical porosity value similar to that of well-
physical analysis of log data. (2) The elastic constant C33m sorted sands.
of the shale matrix is also dependent on the clay content. The bedding-normal VP -φ crossplots for the data set
The most direct and accurate way to compute VS is to from 22 wells worldwide (but mostly from Gulf of Mexico
use a VP-VS transform for shales (e.g., Vernik et al., 2002): and West Africa wells), filtered using vcl values greater
than 30%, is shown in Figure 26. The most probable shale
1
data contain 40% to 70% clay, which is consistent with

(
VS = aVP4 + bVP2 − c ) 2
,

(34)
most typical values reported for shales worldwide. This
well data set is basically the same as the one plotted on
where VP (in km/s) is the compressional-wave velocity the bulk-density-versus-P-wave-velocity graph in Fig­ure
given by equation 11 of Chapter 2, and a = 2.84e–3, 25 of Chapter 1, which includes 25 wells. The three wells
b = 0.287, and c = 0.79. This relationship describes an ac- excluded from Figure 26 are suspected to contain extend-
celerated, nonlinear drop in VS as the sediment porosity ed log intervals with shales undergoing mar­ked thermal-
approaches 40%. It can also be adapted for sands/sand- expansion-related pore-pressure generation, which alters
stones, although with different coefficients, as is de- the VP -φ relationship.
scribed later in this chapter. As we shall see below, this Data from a subset of 12 wells, on which pore-pres-
model is consistent with dipole shear data from both soft sure and effective-stress analyses have been performed by
and tight shales in most Tertiary basins. The coefficients using the methodology discussed in Chapter 2, are shown
in the VP-VS transform are calibrated using shales with vcl in the velocity-versus-porosity space of Figure 27, color-
values from 40 to 70% and a predominantly illite/smec- coded for the vertical effective stress. An important obser-
tite/chlorite clay mineralogy. Note that the extremely vation here is that VP monotonically increases with the
clay-rich shales with abundant kaolinite clay (e.g., from effective-stress-induced porosity reduction.
the equatorial zone offshore West Africa) may have a This monotonic and predictable velocity variation
slightly different value for a (as high as 0.006), and, re- with porosity (at least, in the 6 to 30% range sampled) and
spectively, smaller VP/VS values. with burial stress makes it possible to estimate shale pore
The equations presented in this section, along with pressures with sufficient accuracy by combining the
formulas 11 and 12 of Chapter 2, comprise our Shale shale-compaction model φ = φo exp(−σ /Cm ) with equa-
Model. In contrast to the models for sands/sandstones, the tion 32. The resulting shale-velocity model (equation 14
shale model is formulated in terms of water-saturated ve- of Chapter 2) has the advantage of being able to account
locities, so it can be immediately applied to core and log for the typically overlooked clay-content effects in pore-
data in the zero to 40% porosity range. pressure predictions (Vernik, 2011, and Chapter 2). It is
It can be stated that generally the sonic velocity of also instrumental in the construction of velocity profiles
shales measured in wells with a low relative dip angle and earth models in compaction-dominated basins.

03-Vernik_Ch03.indd 67 12-08-2016 20:02:21


68 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 24.  A log interval of several hundred meters from a vertical well in West Africa, comparing sonic VP log data with
interval velocity data from vertical seismic profiling (VSP).

Obviously, the shale model discussed here is data-driv- Katahara (1996). Katahara’s work is based on the assump-
en rather than micromechanics-based. Nonetheless, it turns tion that elastic properties of illite are the same as those of
out to be quite useful in many applications in both geo- muscovite. However, this assumption is not necessarily
physics and rock mechanics — notably in estimation of the justified, because, despite some structural similarities,
bedding-normal anisotropic elastic stiffnesses of clay-ma- muscovite contains less structural water (and/or hydroxyl
trix aggregates, which are extremely hard to measure or groups) than illite does, which leads to a difference in den-
calculate (e.g., Militzer et al., 2011). Assuming that the sities of approximately 0.10 to 0.12 g/cm3. In fact, illite is
predominantly illite/chlorite/kaolinite clay mineralogy sur- often referred to in petrographic literature as hydromusco-
vives the advanced stages of shale diagenesis, the bedding- vite. Thus, a respective reduction in illite’s elastic proper-
normal matrix elastic properties of this clay mixture (with ties is to be expected. It should be also pointed out that a
a strong, but quantitatively unknown alignment of clay par- quantitative assessment of the clay-particle preferred ori-
ticles) can be inverted from the core and log data using our entation (or orientation distribution function, ODF) is re-
shale model (Vernik and Kachanov, 2010): C33cl = 33.4 GPa, quired to invert for C33 and C44 elastic stiffnesses of actual
and C44cl = 8.5 GPa. These values are included in Table 2 of clay minerals from the aforementioned effective clay ag-
Chapter 1, together with the mean shale matrix density of gregates with a strong but still imperfect clay-particle ori-
2.73 g/cm3 attributable to conventional mudrocks. entation typical of very tight shales.
These data-derived properties of clay aggregates To summarize, the heuristic model for conventional
are fairly close to the estimates by other workers who used shales that is formulated in equations 32 through 34 for the
the same extrapolation technique (e.g., Eastwood and bedding-normal direction is fairly descriptive of the experi-
Castagna, 1987) and are much lower than the estimates of mental core and log data acquired in compaction-dominated

03-Vernik_Ch03.indd 68 12-08-2016 20:02:23


Chapter 3: Seismic Rock Properties and Rock Physics 69

5.5
cl
Vsh = 1
5.0 0.2

4.5
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4.0
0.8
3.5

VP (km/s)
3.0

2.5

2.0

1.5

1.0

0.5
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Porosity

Figure 25. Bedding-normal VP and VS versus porosity for a 7 40


data set of shale cores measured at 20 to 40 MPa, overlain by Color: Effective stress (MPa)
the shale model predictions for clay contents from 20 to 80%. Figure 27.  Data for a subset of 12 wells from Figure 26, but
Typical shales with 40 to 70% clay are shown here in blue, color-coded here for the vertical effective stress. Note the
whereas shales with lower clay content are in red. Adapted monotonic variation of the shale porosity and VP values with
from Figure 11 of Vernik and Kachanov (2010). Used by stress. Adapted from Figure 13 of Vernik and Kachanov
permission. (2010). Used by permission.

basins worldwide. Notably, this model predicts a dramatic


5.5 increase in VP /VS values, with the shale porosity approach-
cl Vsh = 1 ing the critical porosity in a manner similar to that observed
5.0 0.2
in poorly consolidated sands. It also describes a normal
4.5 compaction trend for the shale velocities, as well as pore-
4.0 pressure-variation effects on sonic and seismic velocities.
0.8 Importantly, the Shale Model applies to shales in differ-
3.5
ent parts of the world, with some exceptions being for high-
3.0 ly cemented or recrystallized mudstones and for tight shales
VP (km/s)

undergoing unloading. It should be pointed out that the con-


2.5
ventional shale model presented here can be applied in earth
2.0 model building for seismic imaging, AVO modeling/inter-
1.5
pretation, and petroelastic-inversion projects involving both
compressional- and shear-impedance volumes.
1.0
0.5
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
Rock-physics templates in siliciclastic sequences
Porosity Combining into one template all the models that are
1 97 applicable to sandstones and shales with variable porosi-
Color: Frequency ties is likely to be quite complex and unnecessary.
Figure 26. Bedding-normal VP as a function of porosity, for However, depending on the situation, a more straightfor-
shales from 22 wells (from the 25 wells included in Figure 25 ward, fit-to-purpose template can be constructed to ana-
of Chapter 1), superposed by the shale-model predictions for lyze, evaluate, and ensure the quality of log data and log-
a variety of clay contents. Color-coding is of the probability model computations (such as shale content, porosity, sat-
density, showing a trend that can be extrapolated to mud uration, and the like).
velocity (1.5 to 1.7 km/s) at a shale porosity of approximately Figure 28 presents a log plot for an extensive depth in-
40%. Adapted from Figure 12 of Vernik and Kachanov terval from a deep well in the deepwater Gulf of Mexico (at
(2010). Used by permission. a water depth of almost 2 km). Because the well was drilled

03-Vernik_Ch03.indd 69 12-08-2016 20:02:25


70 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 28.  Log plot for the lower half of a deep well drilled in the Gulf of Mexico and penetrating wet Tertiary turbidites. The
well was drilled with oil-based mud, so correction for the bulk density in sands had to be applied (black curve replacing the
original red one). Computed vertical effective stress is shown in the porosity track.

with oil-based mud (OBM), an invasion correction to the Color-coding in Figure 29 shows the vertical effec-
density log was carried out to restore the water-saturated tive stress in shales, which is computed by using the
sandstones to their in situ conditions before porosity com- methodology described in Chapter 2. Due to the relatively
putation. Shale matrix density was computed by using simple geologic structure penetrated by the well, an as-
equation 8 of Chapter 2. Log-model porosity, computed sumption is made that this effective-stress profile also ap-
as discussed in Chapter 1, clearly shows a gradual reduc- plies to the associated sandstones.
tion in sandstones (from 26 to 12%) and in shales (from Note that the shales form a cluster that can be accu-
18 to 8%). rately matched by the Shale Model with vcl = 0.6 ± 0.2.
The VP and VS logs displayed in the log plot of Figure Again, the effective stress increases monotonically with
28 are the result of careful reprocessing, including disper- depth, causing shale compaction that is reflected in the
sion correction. These velocities can be analyzed using the porosity reduction. This behavior is similar to the behav-
rock-physics templates that are based on the sand/sand- ior already established for our worldwide subset of 12
stone and shale models presented above. The VP -φ plot in wells displayed in Figure 27.
Figure 29 displays just the massive end-member litholo- Sands and sandstones from the upper portion of the
gies from beds with thickness greater than 1 m, which is the displayed interval of the well seem to indicate φcon values of
average log resolution. Two realizations of the Sandstone 22 to 26% and gradually migrate with depth from the 10-
Diagenesis Model, each with the same critical and consoli- MPa model to the 50-MPa model. Note, however, that the
dation porosities but different effective stresses, are added data never quite reach the 50-MPa line, because the actual
to the template. Three shale models, representing clay con- stress within the displayed zone only ranges from 16 to
tents of 40, 60, and 80%, are also included. 38 MPa. The knee at the consolidation porosity seems to be

03-Vernik_Ch03.indd 70 12-08-2016 20:02:26


Chapter 3: Seismic Rock Properties and Rock Physics 71

5.5 4.5
50 MPa
50 MPa
10 MPa 4.0
4.5 20 MPa
Sandstones
10 MPa
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Oil sands
3.5
3.5 6

VP (km/s)
VP (km/s)

3.0
9
Shales
2.5
2.5 13

Shales
1.5 2.0 25

0.5 1.5
0.00 0.10 0.20 0.30 0.40 0.15 0.20 0.25 0.30 0.35
Porosity Porosity

10 40 0 1
Color: Effective stress (MPa) Color: Vsh

Figure 29.  Compressional velocity versus porosity for the Figure 30. A VP-versus-porosity template constructed using
well interval shown in Figure 28. Only homogeneous the two soft-sand models and the Shale Model, showing oil-
sandstones and shales with a bed thickness greater than the saturated sands and shales from three wells in the North Sea.
log resolution of 1 m are included. Two Sand Diagenesis Again, mixed lithologies and thin beds are excluded. See the
Model realizations (at 10 and 50 MPa effective stress) for text for parameterization of all models. Soft Sand Model 1 is
water-saturated sandstones are shown. Critical porosity is 40% shown in green (three lines), and Soft Sand Model 2 (four
and consolidation porosity is 26% for the 50-MPa model, and lines) is in brown. The expected range for the shale model is
critical porosity is 36% and consolidation porosity is 22% for indicated by the two black lines.
the 10-MPa model. The Shale Model is represented by three
lines, with clay contents of 40, 60, and 80%.
models are realized using the same oil properties. The ef-
fective stress ranges from 10 to 50 MPa and is combined
recognizable, even though very few poorly consolidated with φcon values from 25 to 29% and φc values from 37 to
sand layers were sampled in the displayed interval. 41% for Soft Sand Model 1. A range of pore-shape factors,
The second case history comes from the North Sea from 6.0 to 25.0, is used for Soft Sand Model 2.
and involves a Tertiary siliciclastic sequence that includes As Figure 30 demonstrates, shales plot in the expect-
Heimdal Member sandstones and Hermod Formation ed range given by the Shale Model, whereas sands fill in
sands, both of which are classified as arenites. The same the grid made by the two soft-sand models. Note that Soft
sequence, but in a different location, was previously ana- Sand Model 2 is applied in different dynamic ranges of
lyzed by Avseth et al. (2005). In our case history it was porosity for different pore-shape factors. This is done to
penetrated by three wells in the 1900- to 2500-m depth create an en echelon appearance of the model 2 lines that
range and is characterized by a rapid increase in velocity is consistent with the consolidation-porosity and critical-
with depth. That velocity increase is related to a transition porosity parameters selected for Soft Sand Model 1.
from mechanical diagenesis to chemical diagenesis, with Avseth et al. (2005) presented solid petrographic
minute quartz cementation at depths greater than 2200 m. ­evidence of the transition from mechanical compaction
A dedicated, core-calibrated log model was built to com- to minor quartz cementation in this sand/sandstone se-
pute Vsh, porosity, permeability, and water saturation in the quence, at a comparable depth range elsewhere in the
300- to 500-m-thick intervals of interest. North Sea. Our case history with soft Hermod sands tran-
Figure 30 is a VP -φ template constructed for this North sitioning to Heimdal sandstones is totally consistent with
Sea sequence by using Soft Sand Models 1 and 2, in the dual sand model described above and sheds new light
­addition to our two realizations of the Shale Model, with on the elastic rock properties of these prolific oil and gas
vcl values of 60% and 80%. Because some of the sands/­ reservoirs. Specifically:
sandstones are oil-saturated, whereas others are wet,
Gassmann’s equation is used to “saturate” all of them with • The effect of grain-sorting variations on porosity in
oil of the same composition. Therefore, all sand diagenesis this data set is limited to no more than 3 to 4 porosity

03-Vernik_Ch03.indd 71 12-08-2016 20:02:27


72 Seismic Petrophysics in Quantitative Interpretation

units and is accompanied by a minor velocity change;


hence, we have the en echelon appearance of the sand
clusters representing different levels of mechanical
and initial chemical diagenesis.
• The mechanical-compaction and initial chemical-dia-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

genesis effects (such as pressure solution and initial


quartz cementation) produce similar trends in the di-
agnostic templates, making it difficult to discriminate
between these two different mechanisms. When com-
bined with grain sorting, these effects are largely re-
sponsible for the observed variations in porosity and
elastic properties in poorly consolidated sands.
• This data set strongly confirms the role of the consol-
idation-porosity concept in the rock physics of sands
and in their evolution into sandstones during their
burial history. That evolution produces a characteris-
tic inflection point on their diagenesis trends in the
velocity-porosity space. This inflection-point (knee) Figure 31. Water-saturated VP-VS models for arenites,
is especially distinct in Tertiary sands from the North constructed using the Sand Diagenesis Model parameterized
Sea because of their better sorting and their more dra- with 10, 20, and 50 MPa effective stresses in combination
matic manifestation of the late stages of mechanical with the following critical and consolidation porosity couples,
diagenesis via the onset of quartz cementation, com- respectively: 34 and 22%, 38 and 26%, and 42 and 30%. The
pared with the rock-physics behaviors of sands in location of φcon for the middle trend is shown for reference.
other young sedimentary basins of the world. Brine salinity of 100 kppm is used for all three lines.

the model lines at velocities lower than those correspond-


VP-VS and AI-SI relationships ing to φcon is merely a reflection of the different critical
porosities used in the realizations included in Figure 31.
in siliciclastics Of course, the lower the critical porosity is, the greater is
The VP-VS relationship the P-wave velocity of unconsolidated sand with shear ve-
locity approaching zero.
The Sandstone Diagenesis Model described above as It is instructive to zoom in on the mechanical-diagen-
a combination of the Consolidated Sandstone Model and esis portion of the VP-VS plot with both soft-sand-model
Soft Sand Model 1 can be used immediately to display lines superposed, as is demonstrated in Figure 32. A set of
VP-VS evolutionary trends — from the sand suspension at dashed lines from Soft Sand Model 2 displays short, en
φc to extremely low-porosity sandstones (Figure 31). echelon segments with slowly rotating slopes, which map
The three model realizations shown in Figure 31 are the grain-sorting effects in VP-VS space. These lines are
generated using the following parameters: (1) 10 MPa with generated using pore-shape factors that vary from 600 to
φcon = 0.22 and φc = 0.34 ; (2) 20 MPa with φcon = 0.26 10 and a limited range of 16 porosity units spanning dif-
and φc = 0.38 ; and (3) 50 MPa with φcon = 0.30 and ferent porosity values for each line. When the pore-shape
φc = 0.42 . Water saturation is applied via the bulk-density factors drop below 10.0, notably around the consolida-
equations 33 of Chapter 1 and Gassmann’s equation for tion-porosity point (Figure 29), these segments gradually
the bulk modulus. A brine salinity of 100 kppm is used, align themselves with the Sandstone Diagenesis Model
and the porosity range from 3% to φc is plotted for all three realizations for arenites and also with the Greenberg-
cases. The latter condition minimizes the spurious effect Castagna sandstone line, as will be illustrated later.
of microcracks at lower stresses in very-low-porosity It is clear from Figure 32 that no unique VP-VS line
rocks (<3%), which are largely outside of our scope of can provide an accurate prediction of S-wave velocity in
interest. soft sands, even for a narrow lithology group such as ar-
Note that the consolidation-porosity points (the mid- enites. That is because it is hard to predict grain sorting,
dle one is shown on the plot) of the sand-diagenesis-mod- even qualitatively, let alone to predict φcon and φc during
el lines provide a seamless connection in VP-VS space, in seismic exploration and early reservoir-development stages.
contrast to that in VP -φ space. Further, the effective stress With that in mind, Vernik et al. (2002) introduced and
does not seem to have any impact on the VP-VS relation- Vernik and Kachanov (2010) have refined the semiem-
ship in arenites, and the slightly increasing separation of pirical VP-VS models for sands/sandstones and shales (in

03-Vernik_Ch03.indd 72 12-08-2016 20:02:29


Chapter 3: Seismic Rock Properties and Rock Physics 73
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 33. Water-saturated VP-VS plot for arenites from the


Figure 32. Water-saturated VP-VS plot for soft arenites, with core data set (at 20 MPa) and the log data set (at a 5- to
three realizations of Soft Sand Model 1 (the same as in Figure 38-MPa effective-stress range), superposed by (1) the Sand
31) and six realizations of Soft Sand Model 2. Pore-shape Diagenesis Model realized for 20 MPa, critical and
factors are as shown and porosity ranges are from 29 to 45% for consolidation porosities of 38% and 26%, respectively, and a
the segment with p = 600 and from 19 to 35% for the segment brine salinity of 100 kppm, (2) the Vernik-Kachanov (2010)
with p =10. A brine salinity of 100 kppm is used. Adapted from empirical model for arenites with the same brine salinity, and
Figure 6 of Vernik and Kachanov (2012). Used by permission. (3) the Greenberg-Castagna (1992) model for sandstones.

Table 1.  VP-VS transform coefficients for wet sands and salinity-dependent fluid modulus in Reuss averaging. For
shales. Data are from Vernik and Kachanov (2010). shales, the lack of any noticeable VP/VS variation with clay
content in either the core or log data sets is apparent, which,
Lithology Shale Arenite in combination with the Shale Model, aids us in constrain-
a (2.84 to 6.00*) × 10−3 4 × 10−4 ing the bedding-normal elastic moduli and, finally, the ve-
locities. The salinity effect in shales is neglected, although,
b 0.287 0.45 + 0.93 × 10−4 S
in principle, coefficients b and c can be made salinity-de-
c 0.790 1.41 + 2.86 × 10−3 S pendent for shales in a way similar to that for sands.
* The a coefficient values in excess of 5 × 10–3 are typical of kaolinite- Figure 33 compares wet arenite VP-VS lines generated
rich shales from equatorial West Africa. by using (1) the Sand Diagenesis Model with φc = 0.38 ,
φcon = 0.26 , and σ = 20 MPa (the middle line in Figure
the bedding-normal direction) by properly setting the ma- 31), (2) the semiempirical model of equation 34 with the
trix properties (from Table 2 of Chapter 1) for a zero- coefficients for arenites (Table 1) and a water salinity S of
porosity end point and computing the other end point at 100 kppm, and (3) the Greenberg-Castagna sandstone
critical porosity for the water-saturated velocities, using model. Following the method of Greenberg and Castagna
Reuss averaging. The generalized model for both sand/ (1992), the final VS = f(VP,Sw) computation for arenites
sandstones and shales can be expressed by equation 34: containing hydrocarbons is obtained by iteration with the
( )
1/ 2
V S = aV P4 + bV P2 − c , in which the coefficients for Gassmann fluid-substitution equation. The latter process
conventional shales and arenites are given in Table 1. is also applied to obtain water-saturated velocities in the
Those coefficients are determined by first fixing the end core and log/core data sets of arenites for fixed brine prop-
points and then fitting the polynomial functions to the ar- erties (a modulus of 3.2 GPa and a density of 1.05 g/cm3).
enite and shale data sets in the squared VP-VS crossplot The respective data points are included in Figure 33.
(hence the exponent of 1/2 in the equation above; see The rms error of using the functional form of equa-
Vernik et al., 2002). One can verify that the pure, zero- tion 34 for arenites, compared with using the Sand
porosity clay aggregate point (Table 2 of Chapter 1) would Diagenesis Model with a critical porosity of >38% for the
plot right on the shale line with a = 0.00284. entire porosity range, is smaller than 1%, which is ade-
In this plot the matrix velocities for arenites are quate in most applications. A similar accuracy is observed
VPm = 5.48 km/s and VSm = 3.53 km/s, and their P-wave ve- when equation 34 for arenites is compared with the sand-
locity at a φc value of 40% is computed by using a diagenesis-model realizations at φc >38% (not shown).

03-Vernik_Ch03.indd 73 12-08-2016 20:02:31


74 Seismic Petrophysics in Quantitative Interpretation

However, the model is less accurate when φc is less than gas-saturated curves for arenites with an irreducible water
38%, simply because it is not designed for such a case. saturation of 30% and oil and gas moduli of 1.2 GPa and
Consequently, the VS underpredictions by the semiempiri- 0.25 GPa, respectively. These curves are derived from
cal model may be nonnegligible only in very soft, high- fluid substitutions, starting with dry velocities of the
porosity sands under very low stress. Sandstone Diagenesis Model.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

The linear-regression model proposed by Greenberg Superposed on these lines are the semiempirical mod-
and Castagna (1992) for wet sands is also accurate to with- els (dotted lines), which for water-saturated arenites are
in 1% of the model given by equation 34 for arenites and given by equation 34 with the coefficients from Table 1
the Sandstone Diagenesis Model, when the wet P-wave and for hydrocarbon-saturated sands by the formula:
velocity is greater than 2.8 km/s (Figure 33). However,
1
when sandstones are transitioning to the domain of poorly
consolidated sands, the linear VS predictors start to depart
(
VS = bVP2 − c ) 2
,

(35)
gradually from the Sandstone Diagenesis Model line, so
that the separation between them culminates in totally un- with the respective coefficients bo = 0.437 and bg = 0.427;
consolidated sands under very low effective stress at shal- co = 0.71 and cg = 0.22 for relatively light oil (Ko =
low depths. Indeed, it is well established that the values of 1.0 GPa) and wet gas (Kg = 0.25 GPa) cases, respectively.
VP /VS in such low-rigidity sediments become infinitely The subscript o represents oil, and subscript g represents
larger than those computed from the linear VP-VS model. gas here and following.
Let us reiterate this important point: Values of VP/VS calcu- Note that due to nonzero microcrack density at the
lated for wet sands from the Greenberg-Castagna equa- medium stress level σ used (20 MPa) and to the minimal
tions are too low when the P-wave velocity is below ap- porosity of 3%, the models in Figure 34 are not extended
proximately 2.8 km/s (Smith, 2011). to the matrix velocities of arenites. It is also noteworthy
Be aware that in Figures 31 through 33 the P-wave that although the empirical oil- and gas-saturated arenite
velocity of the sediment suspension at the critical-porosi- transforms provide excellent fits to the respective realiza-
ty point depends on the brine fluid modulus, which of tions of the Sandstone Diagenesis Model with φc = 0.38 ,
course is a function of fluid pressure, temperature, and they are hard-wired to specific oil and gas fluid properties
brine salinity (Batzle and Wang, 1992). Figure 34 exem- (Ko = 1.0 GPa and Kg = 0.25 GPa) and should only be
plifies the salinity effect on the VP-VS relationship in wet used as generalized guidelines to simplify the construc-
arenites. In addition, this figure displays the oil- and tion of templates for VP-VS or acoustic impedance versus
shear impedance (AI-SI). However, an important conclu-
sion derived from Figure 34 is that relatively simple
semiempirical models can be used successfully, instead of
more theoretical ones.
When we add to the VP-VS template shale lines with
the coefficient a = 2.84e – 3 for typical shales and
a = 5.00e – 3 for clay-rich, kaolinite-dominated shales
(encountered, for instance, in equatorial West Africa), the
result is seen in Figure 35a. The core data set of ultrasonic
measurements at 20 to 40 MPa confining pressure and
water-saturated conditions suggests that the Shale Model
coefficients given in Table 1 provide a good fit to these
data. Probably somewhat counterintuitive is the fact that
clay content in conventional shales does not affect their
VP-VS relationship measurably.
As is the case for arenites, the shale VP-VS model is
not unique and can easily be adjusted by elevating the
coefficient a from a value of 2.84e – 3 to one of 5.00e – 3,
as is shown in Figure 35a. This results in bifurcation with
Figure 34.  VP-VS models for water-, oil-, and gas-saturated an increasing shale compaction and, hence, P-wave veloc-
arenites. The thin lines are the Sand Diagenesis Models and ity, whereas both model lines converge on the shale po-
the dotted lines are the semiempirical models (equations 34 rosity of approximately 40%.
and 35). Note: 100 kppm water salinity models are the same Figure 35b replots the most-typical-shale line and
as those shown in Figure 32. Adapted from Figure 15 of compares it with the mudrock line of Castagna et al.
Vernik and Kachanov (2010). Used by permission. (1985) and the shale line of Greenberg and Castagna

03-Vernik_Ch03.indd 74 12-08-2016 20:02:32


Chapter 3: Seismic Rock Properties and Rock Physics 75

a) 3.0
Vsh = 1

a = 4e – 3
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

2.0

VS (km/s)
1.0

0
1.0 2.0 3.0 4.0 5.0
VP (km/s)

b) 1 89
Color: Frequency

Figure 36. A VP-Vs plot for shale intervals from 12 deep


wells worldwide. The shale transform using a = 4e – 3
(black) is superposed with a Greenberg-Castagna shale line
(dotted green) to illustrate their coincidence at VP > 2.2 km/s.
Two arenite transforms, for gas (red) and oil (green), are
included for reference.

model (using a = 4.0e – 3) for VP values greater than 2.2.


This is illustrated in Figure 36, which plots VP versus VS
logs in shale intervals from 12 deep wells worldwide
(mostly from the Gulf of Mexico, West Africa, and North
Sea). However, because Greenberg and Castagna’s model
is linear, it tends to overpredict VS in softer shales and
muds characterized by an illite/smectite/kaolinite-domi-
nated clay composition.
Finally, Figure 37 demonstrates the accuracy of pre-
Figure 35.  (a) A VP-Vs template with 100-kppm water, oil, dicting VS from the semiempirical model given by equation
and gas lines for arenites and two realizations of the Vernik 34 with inputs from Table 1 for the log data from 12 wells
et al. (2002) shale model superposed on shale core data at
worldwide, which includes massive sands (overwhelming-
20 to 40 MPa. (b) A typical worldwide shale model, with
ly arenites) and shales as well as mixed lithologies such as
a = 2.84e – 3 (the lower line in (a)) compared with a mudrock
wackes and laminated facies. Time-averaging of velocities
line of VS = 0.862VP – 1.172 (Castagna et al., 1985) and a
Greenberg-Castagna shale line with VS = 0.77VP – 0.867
using Vsh is applied to the intervals with sand/shale lamina-
(Greenberg and Castagna, 1992). tions where 0 < Vsh < 1. It is important to note that just si-
liciclastic rock formations were sampled by the logs shown
(1992). It is quite obvious that the mudrock line is ques- on the plot, with the typical sampling rate of 0.5 ft.
tionable with respect to the prediction accuracy of VS. Its Because uncertainties with the dispersion-corrected
significance as the generalized wet-rock trend, including flexural mode of wave propagation (e.g., Cheng, 2015)
water-saturated sands and shales, might be limited to VP make the ground truth for VS logs in soft formations dif-
values greater than 2.8 km/s, but it is quite inaccurate for ficult to determine from dipole sonic logs, the correlation
VP values below 2.5 km/s — that is, in very soft sediments in Figure 37 should be viewed with caution. Note, for in-
under low stress — velocities that are of interest in geo- stance, the slight tendency for the dipole sonic data to
technical applications. overpredict VS in the softer end of the velocity range,
Interestingly, the shale line of Greenberg and where dispersion corrections are more biased because of
Castagna (1992) coincides almost ideally with our shale the lack of a sufficiently low frequency in the power

03-Vernik_Ch03.indd 75 12-08-2016 20:02:33


76 Seismic Petrophysics in Quantitative Interpretation

4.0 can be combined with VS predictors to quickly generate fit-


R = 0.96 for-purpose AI-SI model lines. A desired family of such
lines, overlain on an AI-SI plot as a template, is quite instruc-
3.0 tive and will be used extensively in Chapters 4 and 5. An
example of such a rock-physics template is given in Figure
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

VS prediction (km/s)

38a. It is constructed for arenites at the medium stress level


of 20 MPa and saturated with water (Sw = 1), oil (Sw = 0.3),
2.0 and gas (Sw = 0.3). Water salinity of 100 kppm was chosen
for all three fluid saturation scenarios with an oil modulus of
1.2 GPa and a gas modulus of 0.25 GPa. Also added is the
model for the most typical shale, which, as previously men-
1.0
tioned, will have to be slightly readjusted for shales in equa-
torial West Africa and adjacent provinces.
Another popular crossplot in rock physics plots VP/VS
0 values versus AI, which visually expands the space, espe-
0 1.0 2.0 3.0 4.0 cially at lower impedances, but does so at the cost of intro-
VS log (km/s) ducing significant nonlinearity for all four fluid models
1 142 (Figure 38b). The same argument is often advanced when
Color: Frequency one is using other template options, such as Poisson’s ratio
Figure 37.  Shear-velocity-prediction accuracy for the versus AI and λρ -μρ , where λ is the Lamé constant and
log-data set of 12 deep wells, with extensive dipole sonic μ = G is the shear modulus. One must always realize,
acquisition. A total of approximately 15 km (50,000 ft) of though, that using one template rather than another is
sonic logs is plotted. Equation 34 with inputs from Table 1 merely a personal choice. Most importantly, the informa-
was used with Vsh as a lithology partition term. The tion content of these crossplots and templates remains the
coefficient of correlation R is 0.96. Adapted from Figure 16 same, as does their ability to relate to AVO attributes.
of Vernik and Kachanov (2010). Used by permission. The quasi-linearity of the shale and hydrocarbon-sat-
urated-sand trends in AI-SI space is particularly striking
spectrum of the dipole sonic tool. Also, the model-based when one is limiting the dynamic range of impedances, as
prediction of VS in thin-bedded rocks is influenced by the is shown in Figure 39. Moreover, the gas-saturated and
accuracy of the Vsh log computation. light-oil-saturated sand lines (with slopes of 0.65 and
Nonetheless, the crossplot in Figure 37 illustrates an 0.70 at their midpoints, respectively) are quasi-parallel to
adequate quality for VS prediction and/or quality checking the shale line. This feature markedly simplifies AVO in-
of dipole shear log data, over a much wider dynamic terpretation and simultaneous impedance inversion in
range than is provided by other popular models that are terms of net-to-gross values and pay volumes, as will be
frequently used in quantitate seismic interpretation. demonstrated later in Chapters 4 and 5.
Because the unified porosity-based model of sand-
The AI-SI relationship stone diagenesis is used implicitly in AI-SI template con-
struction, it is straightforward in Figure 39 for us to map
A foremost conclusion of the previous section is that sand/sandstone porosity onto all three model lines that
we can easily rely on the semiempirical nonlinear equa- correspond to different saturating fluids. Such porosity
tions directly relating P-wave velocity to S-wave velocity mapping is included in the figure and is quite instructive
in water-saturated siliciclastics. These models and em- in quantitative seismic interpretation.
pirical observations become very helpful in transitioning More than 2400 m of total sand intervals (logged
toward understanding the relationships between acoustic with dipole-shear sonic tools) in 19 wells worldwide (in
and shear impedances. Indeed, a simple multiplication of the Gulf of Mexico, North Sea, and West Africa) are
the sonic velocities by the bulk density log yields AI and represented in the AI-SI template shown in Figure 40.
SI log curves over the entire logged intervals in boreholes. Only wet sands and sandstones are displayed and super-
These curves can be evaluated in AI-SI crossplot space, posed by the wet (100-kppm brine) and gas-saturated
and with the appropriate rock-physics template (litholo- arenite lines and the shale line. The color-coding in this
gy, porosity, and saturation) they can be used for data- plot is for the total porosity determined from log model-
quality checks and precursory AVO feasibility studies. ing, with red colors corresponding to the high-porosity
The model parameters discussed for the sandstone dia- sands (greater than 25%), green colors representing the
genesis model in VP -φ space, such as the critical and con- medium porosity sands (15 to 25%), and blue colors
solidation porosities, fluid saturation, and effective stress, being the low-porosity sandstones (less than 15%). As

03-Vernik_Ch03.indd 76 12-08-2016 20:02:34


Chapter 3: Seismic Rock Properties and Rock Physics 77

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 39.  The same template as in Figure 38a, but zoomed


b)
in to focus on the dynamic range of interest in quantitative
seismic analysis. Arenite porosity is mapped on all three lines
for different fluids.

7.0
Wet arenites
from 19 wells
6.0

5.0
SI (g/cm3 × km/s)

4.0 Gas sand


line
Shale line
3.0

2.0
Figure 38.  (a) An AI-SI template for typical shales and
Wet sand line
arenites, constructed by using the sand diagenesis and shale (S = 100 kppm)
1.0
VP -φ models and nonlinear VS predictors multiplied by the
4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0
respective ρb -φ models. (b) A VP/VS-versus-AI template
constructed by using the same models (the lines are truncated AI (g/cm3 × km/s)
at φ = 0.32 for shales and φ = 0.35 for sands). 0.05 0.35
Porosity

follows from the color bar, each color segment corre- Figure 40.  AI-SI crossplots for 19 deep wells worldwide
sponds to five porosity units. (in the Gulf of Mexico, North Sea, and West Africa), showing
Overall, the wet-arenite line fits the data reasonably water-saturated sand intervals only. The wet (100-kppm
well, especially in more or less consolidated sandstones brine) and gas-saturated arenite lines, as well as the shale
with porosities lower than 28% (wet AI > 6.9 km/s ×  line, are included for reference.
g/cm3). Some tendency of less-consolidated, lower-im-
pedance sands to deviate from the trend can also be ob- The situation is particularly notable for older-generation
served. However, as was mentioned above, the dipole sonic-tool runs, when the computed shear velocity might
shear velocity in softer sedimentary rocks should be be a much better option for going forward.
taken with a grain of salt, primarily because of the dis- Excellent agreement is found between the porosity
persion-correction bias that complicates the derivation of mapped onto the AI-SI sand lines in Figure 39 and that of
true shear velocity from the flexural-mode traveltimes. the actual data displayed in Figure 40. This agreement is

03-Vernik_Ch03.indd 77 12-08-2016 20:02:36


78 Seismic Petrophysics in Quantitative Interpretation

helpful in porosity prediction from simultaneous prestack K m [ Kφ ( K m − K f ) − K f ( K m − K )]


impedance inversion of seismic data, even though natural Kd = ,
K mφ ( K m − K f ) − K f ( K m − K ) (36)
scatter in the data related to lithologic and fluid-composi-
tion uncertainties is always present. However, there is a Gd = G.
caveat: Because of the interference of sedimentologic and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

diagenetic effects on the impedance-versus-porosity rela- For the next step, any fluid or fluid mixture can be
tionship, the task is likely to be more challenging in poorly used in forward modeling the saturated elastic moduli
consolidated clastics. An expression of this challenge can and, hence, the velocities (equation 1). By using the aver-
be found in Figure 40, which shows less-regular porosity/ aging technique of Backus (1962), this approach can be
color succession and overlapping in the high-porosity extended for inhomogeneously laminated depositional
sands than is seen in the more-consolidated sandstones. ­facies such as levees in deepwater turbidites. A specially
tailored fluid-substitution code can be written that takes
advantage of the laminar Vsh curve and that seamlessly
Fluid substitution and computes “fluid-replaced” density and velocity curves in
the AI-SI template one step (e.g., Skelt, 2004).
An example of this routine having been applied to an
The process of fluid substitutions using Gassmann’s interval containing oil-saturated diagenetically heteroge-
equation is well described in the literature. The principle neous but largely poorly consolidated turbidites is pro-
consists of inverting equation 1 for the dry frame bulk vided in Figure 41. Homogeneous sand beds and bed-sets
modulus, using log data and estimated fluid and lithologic from this interval belong to the channel depositional fa-
properties. A useful form of the inverted Gassmann’s rela- cies, whereas heterogeneous sand/shale thin beds and
tions solved for dry frame moduli and used as the main laminations represent various levee facies of the turbidite
building block of any fluid substitution routine is system.

RHOBb Vpb Vsb AIb SIb


1.65 G/C3 2.65 2 KM/S 4 1 KM/S 2 3 G/C3 × KM/S 9 2 G/C3 × KM/S 5
GRn RHOBo Vpo Vso AIo SIo
DEPTH
30 GAPI 280 (METERS) 1.65 G/C3 2.65 2 KM/S 4 1 KM/S 2 3 G/C3 × KM/S 9 2 G/C3 × KM/S 5
GRn Vsh PHIT Sw RHOBg Vpg Vsg AIg SIg
280 GAPI 30 0 V/V 1 0 V/V 0.4 1 V/V 0 1.65 G/C3 2.65 2 KM/S 4 1 KM/S 2 3 G/C3 × KM/S 9 2 G/C3 × KM/S 5
4140.0

4150

4196.0

Figure 41.  Fluid-substitution results for oil-saturated Miocene sands from the Gulf of Mexico. Gas (red curves) and wet (blue
curves) cases are shown for bulk density, compressional and shear velocities, and impedances. Note the less significant fluid
effects in thinly laminated levee deposits identified by brown and green colors in the first track, compared with the channel
facies (yellow dotted pattern). The oil-saturated VS curve is computed from the nonlinear VP-VS model for arenites.

03-Vernik_Ch03.indd 78 12-08-2016 20:02:38


Chapter 3: Seismic Rock Properties and Rock Physics 79

X-ray diffraction (XRD) analyses of the very-fine- 5.5


grained to medium-grained sands from cores show that 5.0
clay volumes vary from 3 to 11%, which clearly classifies
these sands as arenites. The very-fine-grained sands and 4.5
coarse silts typically come from thin beds and laminated
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4.0
levee deposits, with XRD-based clay content in the same Sands
range, and with estimated shale volumes on the log scale 3.5 Vsh = 0

VP (km/s)
(Vsh) varying from 20 to 80%. p = 13
3.0
The following fluid properties are used to compute Wet
fluid-mixture properties from equation 2: 2.5 Shale
Oil
cl = 0.6 Gas
• 70 kppm brine: Kw = 3.1 GPa, ρw = 1.04 g/cm 3 2.0
• 30°API oil with a gas-oil ratio (GOR) of 600 ft3 per
1.5
standard barrel: Ko = 1.2 GPa, ρo = 0.76 g/cm 3
• Gas: Kg = 0.1 GPa, ρg = 0.24 g/cm 3 1.0

The density and velocity variations as a function of dif- 0.5


ferent fluid saturations, including gas, oil, and wet cases, 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40
are clearly a strong function of Vsh as continuous net/ Porosity
gross curves. As is to be expected from the Gassmann Figure 42. A VP-versus-porosity plot, showing just the
equation, the fluid effect on SI in soft sediments is a channel arenite sands (Vsh = 0) from Figure 41. Note that all
small fraction of that on AI. Indeed, SI can often be used three fluid-saturated cases quite accurately match the Soft
as a lithology indicator, whereas AI is affected by lithol- Sand Model 2 lines with a constant pore-shape factor p of 13
ogy and fluid. and the mean matrix properties of arenites. The shale-model
Well-designed log displays showing the original and realization for vcl = 0.6 is included for reference.
fluid-substituted curves can be informative and useful for
understanding spatial variations of seismic rock properties
as a function of fluid type. The log layouts should always Figure 41 is approximately 40 m thick), the sand porosity
be complemented by crossplots of various parameters varies quite substantially and is mostly a function of the
(VP-VS, AI-SI, etc.), overlain by the appropriate rock-phys- grain sorting. However, this variation results in relatively
ics templates discussed in previous sections. As an exam- minor changes in elastic stiffness and/or velocity, which
ple, a VP-versus-porosity crossplot is shown in Figure 42. are accurately captured by the model predictions (Figure
Figures such as this can be designed to focus on specified 42). Note that the slopes of the VP -φ functions are gentler
intervals, facies, or porosity and saturation ranges. than the median shale trend line, which is included for
Because the reservoir section in question here is of comparison.
relatively high porosity but has a low to intermediate con- The last but obviously not the least important tool to
solidation level, the best approach is to model and simulta- use in the seismic rock-property analysis is an impedance
neously quality-check the fluid-substitution results with crossplot populated with the appropriate templates. If di-
Soft Sand Model 2 (equation 31) for drained elastic modu- pole sonic data are lacking or provide a questionable VS
li, in combination with Gassmann’s relationships. As is log, taking advantage of properly calibrated VP-VS models
shown in Figure 42, all three fluid models share the same might be the best option. In fact, as is shown in Figure 43,
matrix properties prescribed for arenites (Table 2 of Chapter the predicted VS and hence the SI could even be the pre-
1), as well as the same pore-shape factor p equals 13.0. ferred approach if any doubts about the quality or consis-
The wet-, oil-, and gas-model lines in Figure 42 are tency of the dipole sonic data persist.
truncated to emphasize their applicability to soft sands The resulting tight, elongated data clusters illustrated
only, but they share the same matrix velocity with aren- in the AI-SI plot of Figure 43 are totally predictable from
ites. It is evident from Figure 42 that these lines are well the model point of view. These clusters are better sepa-
matched by the data. The oil line occupies a median posi- rated on the AI-axis projection, but they overlap in SI and
tion between the wet and gas lines; however, this is mere- can be further collapsed into small ellipses that are coax-
ly a consequence of the relatively light oil in the case con- ial with the trend lines and represent long-wavelength
sidered. Lighter oil with a higher GOR will push the equivalents of the acoustic and shear impedances.
model line toward the gas velocity, whereas heavier oils The relative positions of such ellipses corresponding to
with lower GORs will push it toward the wet case. the likely seismic-property variations of the upper and lower
As is typical of turbidite deposits belonging to the media across an interface control its reflectivity and AVO
same stratigraphic complex (the gross interval shown in signatures. This will be discussed extensively in Chapter 4.

03-Vernik_Ch03.indd 79 12-08-2016 20:02:39


80 Seismic Petrophysics in Quantitative Interpretation

6.0 the changes in P- and S-wave velocities due to changes in


effective stress can be quite accurately estimated for both
Sand
soft and consolidated sands, using the approach presented
5.0 in the section in this chapter on “Drained frame moduli
and effective stress.” The starting point comprises the fol-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Shale
lowing fundamental equations describing the Consolidated
SI (g/cm3 × km/s)

4.0 Sandstone Model (equation 27), which are reiterated here:


−1
 pφ η exp( −dσ ) 
K d = K m 1 + + A (ν m ) o  ,
3.0
 1 − φ 1−φ 
Gas −1
(37)
Oil  qφ η exp( −dσ ) 
G d = G m 1 + + B (ν m ) o  .
2.0 Wet
 1−φ 1−φ 
As we discussed previously, this model combines the Mori-
1.0 Tanaka scheme with the exponential dependence of micro-
4.0 5.0 6.0 7.0 8.0 9.0 10.0 crack density on effective stress. Also as we discussed be-
AI (g/cm3 × km/s) fore, for the group of arenite sands with νm = 0.146, we
Figure 43. An AI-SI template, showing the same subset of have A(νm) = 2.46 and B(νm) = 1.59. Very strong correla-
channel sands presented in Figure 42. The almost perfect tions between the shape factors and the initial crack density
match of the fluid-substitution template to the data is entirely (Figures 9 and 10), along with a dry compressional velocity
due to the estimation of the S-wave log from the VP-VS model. that can be obtained from fluid substitution at the initial
stress condition, substantially simplify the task of fitting
equations 37 to the reservoir sand in question.
4D modeling of sands and sandstones Moreover, the process can be automated within the
rock-physics software that is available for predicting the
The variation of reservoir properties over production likely changes in dry P- and S-wave velocities as a func-
time frames is a complex superposition of pore-pressure tion of the effective-stress changes caused by pore-pres-
and saturation changes. That complexity is related not sure depletion. The computed dry-frame properties can
only to the need to accurately predict new seismic rock then be applied in a fluid-substitution routine that uses
properties of the reservoir lithologies, but also to the spa- Gassmann’s equation to forward model the synthetic seis-
tial distribution and the reservoir-engineering properties mic response of the reservoir in different scenarios of in
(gravity segregation, connectivity, capillary pressure, and situ stress and fluid-saturation conditions.
relative permeability) controlling that distribution. As an example, let us apply the model to the same oil
For these reasons, time-lapse (4D) seismic projects to reservoir displayed in Figure 41, but at an offset location.
monitor reservoir changes are always subjected to the The gross thickness of the reservoir at this new location
nonuniqueness of interpretation, even if the quality of the is 27 m (90 ft), and the net-to-gross ratio, which is the
4D signal is good and the magnitude of the stress/satura- accumulated (1 – Vsh) curve divided by the gross thick-
tion change is strong. Whereas saturation variations can ness, is 0.82. Again, the channel sands and coarse-grained
be adequately handled by Gassmann’s relations, stress silts from the laminated levee facies are characterized by
changes due to pore-pressure depletion within the reser- vcl values from 3 to 11% — that is, they qualify as aren-
voir are less well understood, and multiple conceptual ites and can be adequately modeled by using mean solid
and empirical models exist to account for this effect on matrix elastic properties for this group (from the list in
seismic rock properties. Table 2 of Chapter 1). Initial water saturation in the res-
Four-dimensional seismic-acquisition, processing, ervoir is sand-lithofacies-dependent; it varies from 15 to
and analysis projects have traditionally focused on rela- 45% and averages 30%. The initial in situ vertical effec-
tively soft reservoir rocks because of the perception that tive stress in the interval of interest is approximately
for production time scales, changes in fluids and satura- 20 MPa. No velocity data on core samples from the res-
tions will be the dominant factors in seismic-amplitude ervoir are available, so their stress dependence must be
variations. However, the 4D techniques still have great estimated.
potential to be helpful with more-consolidated sandstones Because stress-related changes in crack density have
— especially those under high initial reservoir pore pres- a negligible effect on total porosity, it can be assumed that
sure and experiencing marked rates of pressure depletion the bulk density will not be affected by stress variations.
with production. In the absence of core measurements, Reservoir properties of sands as well as fluid properties

03-Vernik_Ch03.indd 80 12-08-2016 20:02:40


Chapter 3: Seismic Rock Properties and Rock Physics 81

are given in the previous section. By using equations 36, lifespan of a reservoir. More commonly, the residual oil
we compute dry frame moduli and velocities at initial saturation in fine-grained arenites can be on the order of
conditions first. This is followed by application of equa- 20 to 30%, so the maximum water saturation attainable by
tions 37 and use of the fitting-parameter correlation fairly homogeneous sand bodies is 70 to 80%. Another
shown in Figures 9 and 10. Minor adjustments with itera- consideration during 4D feasibility studies is the possibil-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tions may be necessary to find the shape factors that fit ity of gas exsolution below the reservoir’s bubble point.
the sonic-log data inverted for dry velocities and to keep In the 4D modeling to follow, we shall be concerned
the zero-stress crack-density parameter’s dependence on with four plausible scenarios:
VPd unchanged.
In the next step, we apply fluid substitution to satu- 1. σ2 = 20 MPa (remains constant), and Sw changes from
rate the dry sand with various fluid mixtures and compute 0.25 to 0.70
their velocity dependences on stress in the 5- to 50-MPa 2. σ2 = 40 MPa due to a 20-MPa Pp depletion, and
range (Figure 44). Note that at lower stresses and more- Sw = 0.25 (remains constant)
compressible saturating fluids, VP varies more significant- 3. σ2 = 40 MPa due to a 20-MPa Pp depletion, and Sw
ly with stress. changes from 0.25 to 0.70
It is also instructive to plot the calculated effect of 4. σ2 = 40 MPa due to a 20-MPa Pp depletion, and
stress on impedances by using our AI-SI template adjusted Sw = 0.7, Sg = 0.1
for an initial effective stress of σ1 = 20 MPa and the initial
fluid properties and saturations. The result is demonstrated Note that in the first case, the saturation changes are due
in Figure 45. Note in particular that the stress effect in only to aquifer encroachment. In the second case, which
AI-SI space is fairly consistent with the compaction-in- is less likely to occur in nature (a simple depletion drive),
duced porosity variations, as is highlighted by the tem- the fluid saturation is kept constant but the change in
plate model lines. This observation sheds more light on the stress is quite dramatic. The remaining two cases repre-
fact that log data in VP-VS and AI-SI crossplots typically sent both fluid and stress changes.
are tightly aligned and hence are readily predictable. That These four scenarios are illustrated in Figure 46,
is, VS and SI under a new stress state can be estimated quite which shows zoomed-in versions of Figure 44. The focus
reliably from the respective changes in VP and AI. here is on the initial position of the reservoir sand in VP -σ
Of course, use of the fully gas-saturated and wet-sand space (the green point at 20 MPa) and on possible time-
impedances calculated as a function of stress, as displayed lapse changes, which are indicated by arrows.
in AI-SI template, is not feasible in 4D modeling because The dashed lines in these figures are the result of
these conditions are not reached during the production stress-effect modeling for the residual oil case (blue) and
the residual oil case with exsolved gas (red). As expected,

Figure 44.  Variation in VP and VS with effective stress in


high-porosity, initially oil-saturated sand (φ = 0.30, Figure 45. An AI-SI template, showing the stress points from
Sw = 0.25), as well as in the gas-saturated and wet sands Figure 44. Note that the stress points are slightly misaligned
obtained from the dry-moduli model given by equation 37 in with the model lines for the gas-saturated arenite but are
combination with Gassmann’s equation. fairly consistent for the oil and wet cases.

03-Vernik_Ch03.indd 81 12-08-2016 20:02:41


82 Seismic Petrophysics in Quantitative Interpretation

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b) Figure 47. An AI-SI template, zoomed-in to focus on relative


changes in impedances as a result of four time-lapse
scenarios. The dots correspond to the average sand porosity
of 30% and can be viewed as upscaled impedance loci for the
reservoir interval considered.

case of pressure depletion, water-saturation increase,


and gas exsolution produces a vector dominated by the
variation in SI. These relative changes in the upscaled
seismic rock properties will lead to predictably different
reflectivity and AVO behaviors, which will be examined
next.
Figure 48 shows initial and perturbed sets of density
and velocity logs from the well in question. As is de-
scribed above, each set of model logs is computed with
Gassmann’s inversion for dry frame moduli, stress-effect
Figure 46.  Four scenarios of (a) P-wave and (b) S-wave modeling, and the fluid substitutions. To complete the
velocity changes caused by the 4D effects described in the workflow, the final set of seismic-rock-property logs at
text. Solid lines are those from Figure 44, and dashed lines the selected pressure and saturation conditions can be
show the stress effect on velocity for Sw = 0.7, Sg = 0 (blue) used to predict the prestack seismic response for a time-
and for Sw = 0.7, Sg = 0.1 (red). lapse survey.
The reservoir sands turn out to be quite heteroge-
neous, not only in terms of reservoir properties but also
the fluid effect on VP can be comparable to the stress ef- in terms of their dry frame elastic moduli. This may be
fect; however, the dependence of VS on stress is much due to minute differences in the grain-contact flatness
greater than its dependence on fluid saturation. It is note- between medium-grained and very-fine-grained sands,
worthy that an introduction of 10% gas saturation into the albeit no detectable quartz cement was observed in ei-
pressure-depleted reservoir results in almost no change in ther textural variety. It is noteworthy that basal facies of
P-wave velocity, because of the opposing effects of fluid turbidite cycles, separated by thin shale beds, are typi-
and stress changes in this case. cally coarser-grained and are characterized by slightly
Figure 47 is an expanded view of the AI-SI crossplot higher velocities.
and template, focusing on the seismic-rock-property Another complicating factor is that the overburden
variations corresponding to the four scenarios depicted shales are softer than the underlying ones, which are like-
in Figure 46. Note that the case of constant pressure with ly to be marly shales or marls typically found in deepwa-
a water-saturation increase produces a vector corre- ter settings. However, even with the background of these
sponding to the preferential increase in AI, whereas the complexities, one can clearly recognize the dramatic

03-Vernik_Ch03.indd 82 12-08-2016 20:02:43


Chapter 3: Seismic Rock Properties and Rock Physics 83
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 48.  Initial and time-lapse-modeled logs of seismic rock properties (density, VP, VS, AI, SI). Light green is the original logs (Sw
is 0.15 to 0.45 in oil-saturated sands and σ is 20 MPa); blue represents logs with original stress but an increased Sw of 0.7, that is the
residual oil saturation of 30%; light blue shows logs with residual oil saturation and a stress increase from 20 to 40 MPa; red depicts
logs with residual oil saturation (Sw of 0.7), exsolved gas (Sg of 0.1), and an increase in stress from 20 to 40 MPa.

effect of the 20-MPa-pressure depletion causing a com- the impedances to collapse to better identifiable, slightly
mensurate 20-MPa increase in the effective stress elongated clusters in the AI-SI crossplots (Figure 50).
(σ 2 = 40 MPa) . The overburden and underlying shales were up-
The log interval shown in Figure 48 is plotted in the scaled separately for Figure 50, and that resulted in their
AI-SI space of Figure 49. Again, because the initial VS log split into two distinct clusters that are still fairly consis-
is computed, the data points align strongly along the tent with the shale AI-SI model. Especially noteworthy
model lines; that alignment is related to both the initial in this figure is the fact that the data clusters correspond-
impedance heterogeneity and the stress-induced elastic ing to the upscaled gross reservoir interval moved nearer
stiffening. Note that the stress effect changes both AI and each other while they also gravitated closer to the shale
SI, in keeping with Figure 46, but practically no change is trend — behavior that is due to the presence of the intra-
observed in the alignment of log data points with the oil reservoir shale beds and laminations. Such upscaled po-
model line unless the initial saturation is also perturbed. sitions of the shales and the reservoir facies at the well
The range of plausible saturation scenarios consid- location will predictably control the AVO characteristics
ered in this section moves those data points away from the of both synthetic and seismic gathers of the time-lapse
oil model line for pure, shale-free arenite sands (100% surveys.
net/gross). Upscaling the gross reservoir interval log data To demonstrate such 4D effects on synthetic gath-
from Figure 48, by using a simple box-car filter, causes ers, a bandpass wavelet with a 5/10/40/50-Hz frequency

03-Vernik_Ch03.indd 83 12-08-2016 20:02:44


84 Seismic Petrophysics in Quantitative Interpretation

5.0 band is created and convolved with the reflection-­


Sand
coefficient series by using the Aki-Richards AVO equa-
tions (Figure 51).
As expected, all four scenarios used in synthetic AVO
forward modeling display class III AVO behavior.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4.0 However, there also are observable differences among the


SI (g/cm3 × km/s)

four cases. For example, the 4D AVO response corre-


sponding to a water-saturation increase at constant effec-
Gas Shale tive stress (case 1, Figure 51a) does not change the AVO
gradient by much in the offset range of interest (0 to 40°
3.0 Oil incident angles), and the 4D signal is mainly due to the
overall amplitude reduction. Although the time-lapse
Wet AVO differences and apparently also the full-stack ampli-
tude differences are detectable in this noise-free synthet-
ic, they are quite small and can be overwhelmed by seis-
2.0
mic noise and/or processing effects.
5.0 6.0 7.0 8.0 9.0
Figure 51b illustrates the AVO perturbation resulting
AI (g/cm3 × km/s)
from the effective stress only, without any saturation
1 6
In Situ SW = 0.7 40 MPa 40 MPa 40 MPa changes (case 2). It is quite obvious that, whereas the AVO
Sw = 0.7 Sw = 0.7 class remains the same, the AVO gradient undergoes a
Sg = 0.1 measurable change primarily caused by the reduction in
near-incidence reflectivity giving rise to a more signifi-
Figure 49. An AI-SI template, showing original (green) and
four time-lapse (4D) log-data cases. Colors mostly
cant 4D difference. The change in gradient is likely to be
correspond to those in Figure 48. Shales are kept unchanged easily detectable on CDP gathers, even in relatively ad-
for the 4D cases. No heterolithics with 0 < Vsh < 1 are verse conditions.
included; that is, only pure sands and shales are shown. Figure 51c compares the original synthetic gather
with the one corresponding to case 3, which has the most
dramatic changes in reservoir stress and saturation — a
5.0 20-MPa pressure depletion along with an oil-saturation re-
duction to a residual value of 30% (Sw is 70%). Those
stress and saturation changes result in an AI and SI con-
trast reduction with the encasing shales and, hence, in a
substantial amplitude decrease across the gather, unlike in
4.0
the previous case. Note that the time-lapse synthetic clear-
SI (g/cm3 × km/s)

ly displays the asymmetric top/base reservoir responses,


Base which are primarily related to the separation between the
Gas shale
upper and lower reservoir-encasing shales in AI-SI space.
Top Again, the dramatic changes in both stress and saturation
3.0 Oil shale
imposed in this case produce a 4D signal that is quite vis-
ible and should be easily discernable in field studies.
Wet Finally, Figure 51d shows the potential 4D response
for the case of a stress increase to 40 MPa, accompanied
2.0
by oil depletion to an So value of 20% and gas exsolution
5.0 6.0 7.0 8.0 9.0 due to a hypothetical bubble-point crossing (case 4). The
AI (g/cm3 × km/s)
gas saturation modeled is only 10%, which appears to be
1 6 enough to effectively counterbalance the stress impact
In Situ SW = 0.7 40 MPa 40 MPa 40 MPa and produce a very minor 4D AVO difference that is com-
Sw = 0.7 Sw = 0.7 parable to the first case considered (Figure 51a).
Sg = 0.1
The synthetic forward models of AVO for the most
Figure 50.  An AI-SI template, showing upscaled original realistic “life of reservoir” scenarios discussed above
(green) and four time-lapse (4D) log-data cases. Colors suggest that 4D feasibility results can be compromised
correspond to those in Figure 49. Again, only pure sands and in certain cases by poor-quality seismic data, low signal-
shales are shown, so there are no heterolithic data points to-noise ratios, and difficulties in maintaining constant
connecting them. acquisition geometry and processing steps because of

03-Vernik_Ch03.indd 84 12-08-2016 20:02:46


Chapter 3: Seismic Rock Properties and Rock Physics 85

Original Sw = 0.7 4D
a)
4400
4410
4420
4430
4440
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4450
4460
4470
4480
4490
4500
4510
4520
4530
4540

Original σ = 40 MPa 4D
b) 4400
4410
4420
4430
4440
4450
4460
4470
4480
4490
4500
4510
4520
4530
4540

c) Original Sw = 0.7, σ = 40 MPa 4D


4400
4410
4420
4430
4440
4450
4460
4470
4480
4490
4500
4510
4520
4530
4540

Sw = 0.7, Sg = 0.1,
d) Original σ = 40 MPa 4D
4400
4410
4420
4430
4440
4450
4460
4470
4480
4490
4500
4510
4520
4530
4540

Figure 51.  Synthetic forward modeling of four 4D cases. (a and b) Saturation and stress effects, respectively; (c and d) cases in
which both the saturation and the stress in the reservoir changed (see text for details). The original gather is repeated in all
panels. The incident angles range from 0 to 40°. Adapted from Figure 5 of Vernik and Hamman (2009). Used by permission.

03-Vernik_Ch03.indd 85 12-08-2016 20:02:49


86 Seismic Petrophysics in Quantitative Interpretation

the relatively weak 4D signal in both the AVO and full- It is also quite obvious that our ability to accurately
stack amplitude senses. On the other hand, cases involv- interpret changes in reservoir conditions and to distin-
ing significant stress changes in oil reservoirs, with or guish stress effects from fluid-saturation effects will be
without water encroachment but without gas exsolution, greatly improved when we are using prestack in addition
are capable of producing strong 4D AVO and full-stack to full-stack quantitative analyses. In this respect, rock-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

amplitude signals. The time-lapse signals related to the physics-based 4D AVO analysis and prestack inversion
time shifts may also be significant but are beyond the still hold significant potential for improved reservoir
scope of this book. management.

03-Vernik_Ch03.indd 86 12-08-2016 20:02:49


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 4: AVO Analysis: Rock-physics Basis

Techniques for analyzing amplitude variation with aniso­tropy effects on seismic reflectivity, and, finally, con-
offset (AVO) and/or with angle of incidence (AVA) link sider what may be the most promising prestack seismic
seismic interpretation and rock physics and have made attribute to be used in exploration — the “fluid factor.”
steady progress during the 30 years since Ostrander first
introduced them (Ostrander, 1984). Advances in seismic
acquisition technology and seismic processing, including Linearized AVO equations
amplitude-preserving conditioning methods that are based and their features
on noise attenuation and residual moveout analysis, make
it possible to deliver high-quality, “AVO-ready” offset and Modern velocity analysis, multiple-attenuation tech-
angle gathers from 2D and 3D surface seismic surveys. niques, spectral balancing, and other amplitude-preserving
In addition, equally impressive progress has been methods of gather conditioning for AVO studies often
made in the fields of sonic- and density-log acquisition yield high-quality data in favorable geologic settings, such
and processing and in rock physics. As demonstrated in as those free of salt diapirs, magmatic intrusions, or struc-
Chapter 3, the science of rock physics uses laboratory tural complexity with intense faulting and steeply dipping
measurements of seismic rock properties in combination interfaces. Persistent difficulties with AVO interpretation
with information on rock composition, pore geometry, in its early years were related to low fold of seismic data,
fluid saturation, and effective stress to provide accurate NMO stretch, contaminating multiples, and anisotropy, all
estimates of P- and S-wave velocities and bulk density of of which now can be routinely addres­sed by prestack-da-
geological formations in situ. Therefore, seismic rock ta-conditioning software placed in knowledgeable hands.
properties not only define AVO behavior but also link it The initial analysis of high-quality, conditioned gath-
with reservoir parameters in the framework of quantita- ers can be performed by using one of the available AVO
tive seismic interpretation applied in exploration and res- equations. Some software packages include the exact
ervoir-characterization settings. Because these relation- AVO equation, taken from Zoeppritz (1919); however, it
ships are the main subject of seismic petrophysics and is most common to operate with various simplifying ap-
rock physics, AVO analysis has grown to be more robust. proximations of that equation. The approximations, which
Even so, it is still an ambiguous seismic-exploration are linearized versions of the Zoeppritz equations, hold
method that is prone to pitfalls stemming from issues with well for relatively small elastic-property contrasts across
seismic image quality, signal-to-noise ratio, resolution, an interface between two geologic layers, and for rela-
and nonuniqueness of hydrocarbon saturation effects on tively small angles of incidence for plane P-waves im-
the seismic properties of reservoir rocks. pinging on that interface. Most importantly, the approxi-
Nonetheless, comprehensive AVO (and, of course, mate AVO equations linearized with respect to seismic
AVA) analysis and interpretation markedly improve seis- rock properties are more intuitive than the original one in
mic-based reservoir-quality evaluation and, occasionally, the way that they relate the P-wave reflection coefficient
fluid predictions, thereby significantly reducing explora- to the angle of incidence by using the densities and
tion drilling risk in seismic-amplitude-­supported plays and ­velocities of the layers in contact. Those approximations
prospects. As such, the AVO-related attributes are among can be readily used to explore AVO behavior of a single
the most powerful direct hydrocarbon indicators (DHI). interface between two geological formations with infinite
In this chapter we revisit popular AVO equations, de- thicknesses in, so-called, half-space models. This kind of
velop a rock-physics-based AVO classification scheme, modeling does not require any specialized software, but,
demonstrate a straightforward link between acoustic- of course, it can be rather imprecise.
impedance and shear-impedance templates and the syn- The exact AVO equation or its approximations are
thetic gathers generated on their basis, account for polar also used in synthetic forward modeling with either
87

04-Vernik_Ch04.indd 87 12-08-2016 20:01:22


88 Seismic Petrophysics in Quantitative Interpretation

conceptual (blocky) geologic models or with well logs where the first term contains the normal-incidence P-wave
sampling the seismic rock properties of the actual geo- reflectivity RP = A = ΔAI/(2AI) and the second term is the
logic formations with finite thicknesses in the subsurface. so-called Poisson’s ratio reflectivity, which is a function
The generation of synthetic seismic gathers, and compar- of the contrast in the Poisson’s ratio. Even though equa-
ative analysis of those gathers with conditioned seismic tion 3 is parameterized differently, it is numerically simi-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

CDP gathers, are essential in any quantitative amplitude lar to equation 2 for small contrasts.
interpretation. Fatti et al. (1994) reformulated the Aki-Richards
Castagna et al. (1998) used the work of Aki and equations in terms of acoustic impedance, shear imped-
Richards (1980) to formulate the following popular ap- ance, and density:
proximation for the P-wave reflection coefficient of an
1 ∆A I  1 ∆A I ∆SI  2
interface: R (θ ) ≈ + cos −2 θ − 4K sin θ
2 AI  2 AI SI 
R (θ ) ≈ A + B sin 2 (θ ) + C sin 2 (θ ) tan 2 (θ ) , 1 ∆ρ
− (tan 2 θ − 4K sin 2 θ ) , (4)
1  ∆V P ∆ρ  2 ρ
A = + ,
2  V P ρ 
where K = (VS/VP)2. This form turns out to be relevant not
1 ∆V P
2
 V   ∆V S ∆ρ  (1) only in prestack simultaneous impedance inversion but
B = − 2 S   2 + ,
2 VP  V P   VS ρ  also in AVO analysis, as will be demonstrated below.
As with the Aki-Richards (or the Shuey) equations, the
1 ∆V P
C = . Fatti isotropic approximation can be reduced to two terms
2 VP by dropping the last term, which in this case is the density
In these equations, VP, VS, and ρ define the average P- and term and which by itself only affects the AVO curvature at
S-wave velocities and the bulk density for each of the two far to very far angles of incidence. Combining this two-
layers that form the interface. Seismic rock-property con- term Fatti equation with equation 2, the S-wave reflection
trasts across the interface between the two layers are de- coefficient RS = ΔSI/(2SI) can be expressed as a function of
fined as ∆V P = V P 2 − V P1, ∆V S = V S2 − V S1, and the AVO gradient at near angles — that is, when cos −2 → 1:
∆ρ = ρ2 − ρ1 , respectively. Chopra and Castagna (2014), RP − B
who provide a comprehensive review of AVO equations, RS ≈ . (5)
8K
refer to equations 1 as the Aki-Richards approximation.
The three-term Shuey equation is almost identical to However, there is a principal difference between the lin-
equations 1, but its simplified version has only two terms ear Aki-Richards two-term formula and the two-term Fatti
(Shuey, 1985) and is given as equation, which is not linear even at relatively near angles.
That peculiarity will be examined later in this c­ hapter. We
R (θ ) ≈ A + B sin 2 (θ ) , (2) also must be aware that largely because of the noise in seis-
mic prestack data the inversion for acoustic and shear im-
where the coefficients A and B define the intercept and pedance reflectivity may be in error, which can be adequate-
gradient, respectively, of the AVO curve in a plot of the ly mitigated only using well control.
reflection coefficient versus the squared sine of the The two-term-equation forms are quite useful in im-
incident angle θ. Equation 2 is typically claimed to be pedance-based AVO predictions, notably in situations when
accurate for angles of incidence smaller than 25° to 30°, far-offset amplitudes are compromised by a poor signal-to-
if the velocity and density contrasts across an interface noise ratio, persistent multiples, NMO stretch effects, and
are small. The equation is normally referred to as the time misalignment. Mallick (2007) arrives at the same con-
Shuey approximation or the two-term Aki-Richards clusion in his wave-equation modeling study, pointing out
approximation. that interference effects caused by coupling of the P- and
As was emphasized in the original publication by SV-waves in a finely layered medium are difficult to remove
Ostrander (1984), the AVO gradient for a shale-to-gas- with standard gather-conditioning techniques.
sand interface is a strong function of the contrast in the Additionally, the two-term equations can be employed
velocity ratio (or in the Poisson’s ratio) between the two in a simultaneous impedance inversion of seismic data, es-
media. This notion was expanded by Verm and Hilterman pecially if the quality of the prestack data is compromised
(1995) in their version of the approximate AVO equation, beyond 30°. The formulation of the P-wave reflection coef-
and they cast it in terms of near-incidence reflectivity and ficients in terms of the two impedances (e.g., the two-term
the “Poisson’s reflectivity”: version of equation 4) is quite convenient — the inversion
∆ν engine has to deal with two independent variables (AI and
R (θ ) ≈ R P cos2 θ + sin 2 θ , (3) SI) instead of three (VP, VS, and ρ), thereby leading to more
(1 − ν )2 stable optimization solutions. Solving for density would be

04-Vernik_Ch04.indd 88 12-08-2016 20:01:22


Chapter 4: AVO Analysis: Rock-physics Basis 89

crucial in many reservoir-characterization and exploration rise to substantial differences between the static and dyna­
projects, but it is virtually impossible in the most typical mic Poisson’s ratios of the same rock (Thomsen, 1990).
situations when the very-far-offset amplitudes are either On the other hand, the AVO technique has been prov-
unavailable or are noisy and corrupted beyond the limits of en most powerful in this very group of relatively soft
reliable amplitude recovery. sands, so an alternative parameterization is desirable. For
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

instance, crossplotting the Lamé constants multiplied by


bulk density (λρ -µρ ) has been pursued occasionally in
AVO classification in AI-SI space quantitative seismic interpretation, including in AVO
studies (e.g., Goodway et al., 1997). Such an approach
AVO classification is critically important for diagnos- may be justified if these parameters are derived from ac-
tic purposes. It is traditionally performed by using inter- curate log-based acoustic and shear impedances, rather
cept-versus-gradient or A-B crossplotting (Castagna et al., than from less accurate inversion data.
1998). Using this approach, a linear background-reflec- It is quite obvious that the inversion errors are dou-
tivity trend can be established in a limited two-way time bled in the Lamé space compared with the AI-SI space
window and then calibrated with the empirical mudrock (Chapter 3). Consequently, the introduction of these pa-
line predictions of the dominant VP/VS values in this win- rameters does not add any new information and comes at
dow (Castagna et al., 1985). the expense of strong nonlinearity of major lithologic and
Three caveats are associated with this method, how- diagenetic trends in the λρ -µρ crossplot, even for well-
ever. First, it may be accurate only insofar as the VP-VS re- consolidated rocks. The mentioned nonlinearity is also a
lationship is linear, which, as we have established in complicating feature of the rock physics templates rely-
Chapter 3, is not the general case. Second, AVO gradient B, ing on velocity or Poisson’s ratio.
as an attribute of the specified interface, typically is derived It is straightforward to examine the AVO characteris-
from the two-term Aki-Richards (or Shuey) approximation tics created by each sand/shale pair (dipole) in the for-
of the exact Zoeppritz equation, which can be inaccurate at ward-modeling sense, by selecting a shale point and the
relatively large property contrasts and/or at relatively large corresponding (saturated with different fluids) sand points
offsets. Finally, an ellipse encompassing the dominant A-B prescribed by the theoretically validated rock-physics
pairs in the crossplot corresponds to a relatively wide range models in an AI-SI template. To introduce a measure of
of velocity ratios, introducing a bias in the proper defini- uncertainty, those mean impedance points can be replaced
tion of the background reflectivity trend (Ross, 2016). with small ellipses defined by probability density func-
Indeed, depending on the width of the temporal window tions (PDF) in impedance space. This approach can re-
plotted or inherent variability of mudrock properties, such place the process of using conventional intercept-versus-
as VP/VS, inside of this window, the slope of the major axis gradient plots (e.g., Castagna et al., 1998) to interpret the
of the ellipse corresponding to the background trend can A and B attributes in terms of seismic rock properties and
and typically does undergo a noticeable rotation. their internal relationships.
Alternatively, acoustic-impedance-versus-shear-­ Figure 1 demonstrates this alternative rendering of the
impedance (AI-SI) space lends itself conveniently to clas- AVO classification scheme — an a­ lternative that is based
sifying AVO responses from the rock-physics perspective. solely on the AI-SI template. By replacing the A-B scheme
That is the case primarily because the application of rock- with the AI-SI approach, we apparently make the AVO in-
property trends, model lines, and lithology/saturation tem- tercept/gradient interpretation more implicit and we focus,
plates provides important constraints for both incident and instead, on the rock property relationships and controls.
reflective media, which jointly affect seismic reflectivity. This scheme can be equally instrumental in AVO feasibility
Sand and shale diagenesis models, which have been de- studies that use offset well data and in simultaneous imped-
rived in AI-SI space by Vernik and Kachanov (2010) and ance inversion of prestack seismic data, thereby making the
are presented in Chapter 3, will be given a preference in the AI-SI crossplot an integral tool in any amplitude-based ex-
text to follow. However, any alternative models and elastic ploration and reservoir-characterization process.
parameters could prove useful in template construction. It should be emphasized that this AI-SI template-
It is often suggested that the Poisson’s ratio should be based classification of AVO responses is shale-centered;
the property of choice in AVO analysis (Chopra and that is, the most probable P- and S-wave impedances of
Castagna, 2014), and this notion has some merits for gas- the shale formation’s interval of interest must always be
sand identification in a seismic-bright-spot environment. specified in order to draw the classification border lines
However, the Poisson’s ratio is not a well accepted parameter as shown in Figure 1.
from the dynamic-wave-propagation point of view, espe- Note that, as is shown in Figure 1, Class II AVO occu-
cially in poorly consolidated lithologies with relatively soft pies a relatively narrow sector defined by the two symmet-
frame moduli, when viscoelastic, nonlinearly elastic, and ric rays originating from the chosen shale point and in-
inelastic effects can be pronounced. Those effects may give clined at oblique angles to the vertical. Obviously, the

04-Vernik_Ch04.indd 89 12-08-2016 20:01:23


90 Seismic Petrophysics in Quantitative Interpretation

5
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

I
II
AVO class
SI (km/s × g/cm3) 4
III

3 Shale

b = 1/4

2 IV

0
0 2 4 6 8 10 12
AI (km/s × g/cm3)
Figure 1.  An AVO classification scheme based on acoustic-impedance-versus-shear-impedance space. Class boundaries radiate
from the selected shale reference point lying on the Vernik-Kachanov shale trend (Vernik and Kachanov, 2010). Arenite sand-
diagenesis lines for brine (blue), oil (green), and gas (red) delineate AVO class-specific sand segments bounded by class
boundary intersections.

vertical line in AI-SI crossplot signifies that DAI = 0. The 1 ∆A I  1 ∆A I ∆SI  2


R (θ ) ≈ + cos −2 θ − 4K sin θ . (6)
Class II sector encompasses possible sand (stone)/shale di-
2 A I  2 A I SI 
poles satisfying the condition that ΔAI is a small value, and
it can be repositioned as necessary according to the back- For small angles and VP/VS = 2, it follows from equation 6
ground shale impedance. The boundary lines of the Class II that the AVO gradient B is zero when the relative contrast
sector are somewhat loosely defined using the zero-offset in SI is half of that in AI:
reflection coefficient or ­ intercept A for gas sands ∆SI 1 ∆A I
∆ A I (2A I ) = ±0.02, which is consistent with Rutherford = . (7)
SI 2 AI
and Williams (1989) and Castagna et al. (1998).
The right-side half of the sector occupied by the Class According to our assumption, AI/SI = VP/VS = 2, so we
II AVO responses in AI-SI crossplot is often referred to as easily derive the approximate condition for B = 0 at small
Class IIp (e.g., Chopra and Castagna, 2014). This AVO angles as ΔSI/ΔAI ≈ 1/4. In the general case with an arbi-
subclass is characterized by inversion of the reflection-­ trary VP/VS value, the following condition for the zero
coefficient sign from negative to positive for the top AVO gradient applies:
sand(stone) interface, thereby crossing the zero line at ∆SI ∆A I
relatively small to medium incident angles. This phenom- 8K = . (8)
SI AI
enon in general is typical of the AVO behavior transition-
ing from Class II to Class I and is frequently referred to as Hence, except in poorly consolidated sequences (with
polarity reversal, which is confusing ­because the term mean VP/VS > 2.1) and low-porosity sand/shale sequences
“polarity” is traditionally reserved for describing wavelet (with mean VP/VS < 1.9), the line SI = a + bAI with the
properties rather than the reflectivity sign. slope b = 1/4 and passing through both the incident- and
Unlike the case with Class II AVO, the transition from reflecting-media positions in AI-SI space will almost al-
Class III to Class IV has a more rigorous definition that ways correspond to a zero AVO gradient with approxi-
takes advantage of the isotropic, two-term version of the mately 90% accuracy. Reasonably small deviations
AVO equation presented by Fatti et al. (1994): (smaller than 0.1) from the condition that VP/VS = 2 only

04-Vernik_Ch04.indd 90 12-08-2016 20:01:24


Chapter 4: AVO Analysis: Rock-physics Basis 91

slightly affect the slope of this boundary line between upcoming modeling examples we will use strict VP, VS,
Classes III and IV, so the approximation of b ≈ 1/4 holds and bulk-density prescriptions for shale and sands/sand-
widely in siliciclastic sequences. stones saturated with different fluids according to the
It follows from the AVO classification template pre- rock-physics models discussed in Chapter 3. For illustra-
sented in Figure 1 that the shale AI-SI model line can be tion, the AI-SI plots will be compared with the VP-VS
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

referred to as the background trend in this AVO space. This plots. That will be followed by the isotropic half-space
mitigates the ambiguity and bias related to the velocity ratio reflection-coefficient-versus-­angle-of-incidence relation-
variations intrinsic to any AVO intercept-gradient plot men- ships and synthetic models for a 25-m-thick blocky sand
tioned above. As seen in Figure 1, all shale-on-shale reflec- or sandstone encased in isotropic, homogeneous shale.
tions belong to Class I or Class IV depending on the sign of Wet, oil, and gas-saturated sands will be used to empha-
the impedance contrast. That is, of course, the case for con- size that the proposed AVO classification is basically flu-
ventional shales and marls with clay-content and local pore- id-independent, unlike the conventional scheme, which is
pressure variations that are typical for clastic sequences, as geared toward gas-saturated sands.
soon as these various background lithologies plot along the
same shale trend. It should be realized that, as is shown in AVO Class III and the zero-gradient case
Chapter 3, any conventional shale line selected for the AI-SI
template will be characterized by a VP/VS value that decreas- Let us start with Class III AVO examples, because this
es dramatically with initial compaction of mud having the class is characterized by the most anomalous response with
shear modulus and S-wave velocity values close to zero. respect to the background trend (anomalous in the conven-
That decrease significantly decelerates with further com- tional-analysis sense, when we are using an intercept-gra-
paction and chemical diagenesis of mudrocks. dient crossplot). We will review two plausible cases with
More importantly, the AI-SI AVO template with rock- low and medium impedances, respectively, that are pre-
physics model lines clearly illustrates that any fluid satu- sented in Table 1. Note that both shale and sand/sandstone
ration, including water in moderate- to high-impedance properties used in these and all subsequent models present-
sands, is capable of causing noticeable deviation from the ed in this chapter are quite realistic and taken from log data
background shale trend and thereby triggering an anoma- acquired in Tertiary sedimentary basins.
lous AVO response. In fact, the only unambiguous thing Figure 2a shows these respective cases in the VP-VS
that those deviations tell us is that they are related to a template, and Figure 2b shows them in the AI-SI template.
change in lithology. Note that the vectors connecting the two shale points with
Consequently, it is grossly misleading to claim that all those of the corresponding hydrocarbon-saturated sands
brine-saturated clastic rocks — even those over a limited lie above — that is, clockwise — from the B = 0 line. On
depth range and in a particular locality — should plot along the other hand, these vectors are positioned counterclock-
the same trend in any rock-physics or AVO intercept/gradi- wise and far from the subvertical lines that pass through
ent space and, in so doing, form a background trend against the respective shale points and mark the transition from
which deviations might indicate hydrocarbons. By the same Class III to Class II.
token, no simplistic direct hydrocarbon indication (DHI) Note also that the low-impedance shale/gas-sand di-
that uses visual examination of seismic gathers is possible, pole displays a stronger deviation from the zero-gradient
in principle, unless careful calibration and quantitative in- line, predictably indicating that it should produce a stron-
terpretation are performed. Unfortunately, during the early ger negative-AVO-gradient response. This is exactly what
period of AVO’s application history, inadequate insight into is observed in Figure 3, which shows, for the Class III top
lithology, diagenesis, and saturation effects on seismic rock sand interface, the reflection coefficients that are com-
properties, together with amplitude-­preservation issues, ac- puted from the isotropic three- and two-term Fatti equa-
counted for many misuses and failures of AVO technique in tions 3 and 6 and by the isotropic two-term Aki-Richards
frontier exploration derisking. formula (equation 2).

Table 1.  Seismic rock properties used for Class III AVO
From the AI-SI template to cases (Figures 2 through 4).
synthetic-gather models Low AI values Medium AI values
In the following discussion, a detailed analysis of Lithology VP VS ρ VP VS ρ
some plausible sand/sandstone and shale properties is un- (km/s) (km/s) (g/cm3) (km/s) (km/s) (g/cm3)
dertaken in the framework of the impedance-template- Shale 2.28 0.91 2.25 3.00 1.42 2.45
based AVO classification, in order to showcase various
Sand 1.71 1.00 2.05 2.72 1.58 2.13
AVO scenarios in siliciclastic sequences. Throughout the

04-Vernik_Ch04.indd 91 12-08-2016 20:01:24


92 Seismic Petrophysics in Quantitative Interpretation

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b)
Figure 3.  Two half-space models of Class III AVO cases
corresponding to the sand/shale dipoles shown in Figure 2a.

This observation is critical in computing the aforemen-


tioned AVO attributes A, B, RP, RS and their combinations
outside the incident-angle range and/or impedance contrasts
necessary for these equations to be applicable. Notably, the
errors in the AVO gradient and S-wave reflectivity that
originate from the two-term equations can amplify during
calculation of such a useful combination attribute as the
fluid factor, so we shall return to this problem later.
Figure 4 presents the synthetic angle gathers generated
using the 25-m-thick blocky sand encased in homogeneous
shale half-spaces for the two models described in Table 1.
The plane-wave synthetics, free of the NMO-stretch ef-
fects, are generated with the Aki-Richards three-term ap-
Figure 2.  Two sand-shale dipoles (red for gas sand and green proximation and a 5/10/30/40-Hz trapezoidal wavelet. The
for oil sand) selected to demonstrate the Class III AVO case polarity convention here and throughout this book is the
in (a) VP-VS and (b) AI-SI templates. The two dashed straight one used in North America — that is, the negative imped-
lines in (b) with slopes of 1/4 indicate zero-AVO-gradient ance contrast corresponds to the negative reflection
directions, thereby presenting the Class III/IV boundary in coefficient.
the AI-SI space. Another pair of shale/sand scenarios of interest is
given in Table 2 and shown in Figures 5 through 7. These
The coefficient B in the three-term Aki-Richards
two dipoles display the zero-AVO-gradient response.
equation 1, which is equivalent to the three-term Fatti
Whereas the VP-VS plot does not reveal any consistency
equation 4, can be defined as the AVO gradient for small
­between the two dipoles in terms of their orientation
to intermediate angles only, and Figure 3 provides an un-
(Figure 5a), the impedance crossplot, shown in Figure 5b,
ambiguous confirmation of that statement. The lower-
suggest that in the AI-SI space these dipoles are approxi-
impedance shale/sand dipole is characterized by the
mately parallel (with more than 90% accuracy) to each
greater contrast in rock properties, so that the t­wo-term
other and to the line with a slope of 1/4, which designates
Aki-Richards equation in this figure starts to deviate from
the Class III/IV boundary.
the three-term formulas at angles below 20°. It should be
The lower-impedance dipole includes gas sand, and
emphasized again that the three-term Aki-Richards equa-
the higher-impedance dipole is made up of a shale-over-
tion computes the same AVO reflectivity as the three-term
wet sand couple, for illustrative purposes only. Obviously,
Fatti equation. However, it is evident from Figure 3 that
no immediate significance can be established for these
there is no equivalence observed for their respective two-
two AVO scenarios in AI-SI space in terms of fluid types,
term approximations.

04-Vernik_Ch04.indd 92 12-08-2016 20:01:26


Chapter 4: AVO Analysis: Rock-physics Basis 93

a) Vs Al
0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
RHOB Incident angle
Vp Sl
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm3) 10 0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

3200

3250

3300

b) Vs Sl
0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
RHOB Incident angle
Vp Al
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm3) 10 0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)
3900

3950

4000

Figure 4.  Plane-wave synthetic angle gathers (θ  = 0 to 40°) for two sand/shale models described in Table 1 and shown in
Figure 2b. (a) A lower-impedance sand model exemplified by gas sand; (b) a higher-impedance sand model exemplified by oil
sand. In both cases, a 25-m-thick layer of sand and a 5/10/30/40-Hz bandwidth wavelet are used.

Table 2.  Seismic rock properties used for Class III-IV AVO equation should be ­limited in many situations, such as
transition (Figures 5 through 7). those exemplified in Figures 3 and 6, to an incident angle
that does not exceed 20 to 25°.
Low AI values Medium AI values Synthetic angle gathers for these two cases are gen-
erated from the same options that are mentioned above
Lithology VP VS ρ VP VS ρ
(Figure 7). Both synthetic models are consistent with the
(km/s) (km/s) (g/cm3) (km/s) (km/s) (g/cm3)
three-term Fatti equation predictions of the half-space
Shale 2.72 1.21 2.35 3.28 1.64 2.52 models shown in Figure 6. Again, of p­ articular interest
Sand 1.80 1.07 2.05 3.10 1.69 2.24 is the observation that the high-­impedance-contrast case
reveals amplitude brightening at angles greater than 20
to 25°.
other than that their associated reflection coefficients are
much stronger for the gas-sand case. AVO Class IV
Examination of Figure 6 immediately reveals that
AVO ­gradient B, as determined by the two-term Aki- The next pair of AVO scenarios is selected to high-
Richards equation, is indeed zero in both cases. However, light some curious properties of shale/sand interfaces.
that is not so for the two-term Fatti equation at angles The first case is one in which ΔVS is zero, thereby yield-
beyond 20°, especially in the higher-contrast scenario of ing no shear-velocity contrast. That situation is some-
gas sand. Therefore, application of the two-term Fatti times quoted to result in an AVO gradient of zero

04-Vernik_Ch04.indd 93 12-08-2016 20:01:27


94 Seismic Petrophysics in Quantitative Interpretation

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 6.  Half-space models for zero-AVO-gradient cases


b) corresponding to the sand/shale dipoles shown in Figure 5a
and 5b. In both cases, the AVO gradient B is zero.

even though the amount of deviation is minor for the case


in which ΔVS is zero.
The AVO curves for the two cases are shown in
Figure 9. The plot is totally consistent with the AI-SI ob-
servations predicting a Class IV AVO response for both
scenarios, but there is a much greater positive gradient
for the case in which PR is zero. Consequently, Poisson’s
ratio reflectivity by itself does not define the AVO gradi-
ent. The PR term must always be compared with the
near-incidence reflectivity term RP, and the more appro-
priate condition for an AVO gradient of zero is for PR to
be approximately equal to RP (Verm and Hilterman,
1995).
Note also that the AVO gradient line given by the
Figure 5.  Two sand/shale dipoles (red for gas sand and blue two-term Aki-Richards equation applied to the latter case
for wet sand), selected to showcase the flat AVO response (with Δ(VP/VS) being zero) noticeably deviates from the
with zero gradient in (a) VP-VS space and (b) AI-SI space. still more accurate three-term approximations at incident
Note that the respective vectors in the impedance space are angles as small as θ = 20°. That deviation is due to the
consistent with the dashed straight lines with slopes of 1/4 marked contrast in seismic rock properties between the
that represent Class III/IV boundaries. oil sand and the overburden shale in this scenario.
These two examples clearly indicate that between
(Castagna et al., 1998). In the companion case, Δ(VP/VS) sand and shale, neither the contrast in shear velocity nor
is zero — that is, the Poisson’s ratio reflectivity term PR that in Poisson’s ratio or VP/VS by itself controls the shape
is zero in equation 3 (Table 3). of the AVO curve. On the other hand, the AI-SI space,
Both dipoles are designed to share the same shale which visualizes contrasts in both AI and SI, accurately
point on the VP-VS and AI-SI plots (Figure 8a and 8b). The predicts all the peculiarities of the AVO behavior ob-
case in which ΔVS is zero is arbitrarily chosen to include served in synthetic angle gathers (Figure 10). Indeed, the
the wet sand, whereas the case with PR equal to zero com- amplitudes at top/base interfaces show a tendency to de-
prises the oil sand and is further simplified by the condi- crease with the angle of incidence. The effect is espe-
tion that VP / VS = 2 for both the shale and the sand. It is cially strong for the case involving oil sand; the differ-
quite instructive now to examine the AI-SI plot in ence in AVO gradient between the two cases is fully con-
Figure 8b. Both vectors are rotated counterclockwise sistent with rock-physics-based AVO classification
from the Class III/IV boundary line with a slope b of 1/4, template of Figure 8.

04-Vernik_Ch04.indd 94 12-08-2016 20:01:29


Chapter 4: AVO Analysis: Rock-physics Basis 95

a) Vs Sl
0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
Incident angle
RHOB Vp Al
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

3650

3700

b) Vs Sl
0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
RHOB Incident angle
Vp Al 3
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm ) 10 0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)

4250

4300

4350

Figure 7.  Plane-wave synthetic angle gathers for two sand/shale models described in Table 2 and shown in Figure 5b.
(a) A lower-impedance sand model exemplified by gas sand; (b) a higher-impedance sand model exemplified by wet sand.
A 25-m-thick sand layer and a trapezoidal 5/10/30/40-Hz wavelet are used in both cases. Different “time thicknesses” of the
sand layer between the panels are due to the sand-velocity difference.

Table 3.  Seismic rock properties used for cases in which It is also worth mentioning here that without elabo-
ΔVS = 0 and PR = 0 (Figures 8 through 10). rate rock-physics modeling of the nearby well-log data,
proper AVO interpretation in terms of fluids and/or lithol-
ΔVS = 0 PR = 0 ogies is always a challenge in Classes I and IV. Because
Lithology VP VS ρ VP VS ρ the Class IV cases with a strong positive AVO gradient
(km/s) (km/s) (g/cm3) (km/s) (km/s) (g/cm3) typically involve dramatic AI contrasts, they can be relat-
ed to exotic lithologies (e.g., profoundly zeolitized tuffs,
Shale 3.28 1.64 2.52 3.28 1.64 2.52 organic shales, and coals) that are encased in relatively
Sand 3.08 1.64 2.24 2.00 1.00 2.10 hard conventional shales, marls, or carbonate rocks.

Class IV AVO behavior, when it is characterized by a AVO Classes I and II


strong positive gradient, can rarely be referred to as being
anomalous and indicative of sand presence in convention- The final two rock combinations considered in this
al sand/shale sequences. As is the case with Class I, Class section involve positive contrasts in acoustic impedance for
IV’s AVO response largely belongs to, or plot in the vicin- the reflecting layer (wet sandstone and hard shale or marl)
ity of, the background trend on the standard intercept/­ versus that for the incident layer, which is represented in
gradient plot (Figure 6 of Castagna et al., 1998). both cases by the same, relatively soft shale (Table 4).

04-Vernik_Ch04.indd 95 12-08-2016 20:01:31


96 Seismic Petrophysics in Quantitative Interpretation

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b) Figure 9.  Two Class IV AVO cases corresponding to the


sand/shale dipoles shown in Figure 8a and 8b. The case with
a zero-velocity-ratio contrast (or PR = 0) corresponds to the
green vector in the AI-SI plot of Figure 8b. Again, half-space
models free of tuning effects are shown.

On the other hand, the shale-on-marl-interface vector,


although it implies a strong Class I response, is direction-
ally closer to the line (containing the shale point) with a
slope of 1/4. As we have seen before, this results predict-
ably in a lower AVO gradient. Also, no sign reversal of the
reflection coefficient is observed below an incident angle
of at least 50°. Further rotation of the dipole clockwise,
until it reaches alignment with the slope of 1/4, will result
in an atypical Class I AVO curve that has a zero gradient.
Incidentally, based on the relationship between those zero
gradient vectors and rock physics model lines in the AI-SI
Figure 8.  Two sand/shale dipoles selected to illustrate AVO template, this kind of response is unlikely to be caused
behavior at no shear-velocity contrast (blue vector for wet solely by sand/shale interfaces without the involvement of
sand) and no Poisson’s-ratio reflectivity (green vector for oil
an exotic lithology.
sand) in plots of (a) VP-VS and (b) AI-SI. Note that both
Another distinctive feature of the described shale-
vectors in the impedance space deviate counterclockwise
over-marl interface is that it is characterized by a positive
from the dashed straight lines with slopes of 1/4 and, hence,
density contrast (Table 4), which manifests itself via the
belong to the Class IV sector.
two-term Fatti equation deviating increasingly positively
from the three-term equation, with increasing offset. All
These combinations are used to exemplify Class I and other examples discussed in this section show negative
Class II AVO responses. contrasts in density and negative deviations of the two-
As follows from the AI-SI plot in Figure 11a, the wet- term approximations from those of the more-accurate
sandstone point is located right on the tentative boundary three-term equation. Note, however, that this statement
line between the Class II and Class I sectors. This combi- about greater accuracy is true only when the anisotropy
nation produces the AVO curve for the top sandstone, contrast between the layers is minimal and can be ne-
with a relatively high gradient and a polarity reversal at glected. The anisotropy effects are considered in a special
approximately 22° (Figure 11b) — that is, the previously section below.
mentioned Class IIp behavior. The obvious far-offset One should keep in mind that the models presented
brightening associated with the medium-porosity wet above are based on simplified assumptions of at least
sandstone is frequently quite misleading and can be very ­locally homogeneous seismic rock properties of the inci-
problematic in oversimplified DHI analysis (Figure 12a). dent and reflecting layers, on a scale of a few tens of

04-Vernik_Ch04.indd 96 12-08-2016 20:01:33


Chapter 4: AVO Analysis: Rock-physics Basis 97

a) Vs Sl
0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
RHOB Al Incident angle
Vp
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm3) 10
Time (ms) 0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4350

4400

b) Vs Sl 3
0.5 km/s 5.5 0 (km/s) × (g/cm ) 10
RHOB Incident angle
Vp Al
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) × (g/cm3) 10 3 6 9 11 14 17 20 23 26 29 31 34 37
Time (ms)

4400

4450

4500

Figure 10.  Plane-wave synthetic angle gathers for two sand/shale models described in Table 3 and shown in Figure 8b:
(a) ΔVS = 0, wet sand; (b) Δ(VP/VS) = 0, oil sand. A 25-m-thick sand layer and a 5/10/30/40-Hz wavelet are used in both cases.

Table 4.  Seismic rock properties used for Class I and Class realistic) rock-property distributions and thickness varia-
IIp AVO cases (Figures 11 and 12). tions in both forward modeling and inversion of prestack
seismic data using well logs. Those situations will be dis-
Shale/sand Shale/marl cussed further in this chapter and in Chapter 5.
Lithology VP VS ρ VP VS ρ
(km/s) (km/s) (g/cm3) (km/s) (km/s) (g/cm3)
VTI anisotropy effect
Shale 3.00 1.42 2.45 3.00 1.42 2.45
If reliable information is available on the axial vertical
Sand/Marl 3.37 1.92 2.32 3.28 1.64 2.52
transverse isotropy (VTI) of two layers in question, then,
following Rüger (1998), any approximation of the exact
AVO equation can be generalized to include Thomsen’s
­ eters, that reduce to respective points and, thus, to
m anisotropy parameters ε and δ (Thomsen, 1986):
unique vectors or dipoles on the rock-physics templates.
Nonetheless, now that the framework of AVO interpreta- 1 2 1 2 2
tion using semitheoretical rock physics is established, it is R P (θ )ani ≈ R P (θ )iso + 2 ∆δ sin θ + 2 ∆ε sin θ tan θ ,
possible to apply it to situations with more complex (and (9)

04-Vernik_Ch04.indd 97 12-08-2016 20:01:35


98 Seismic Petrophysics in Quantitative Interpretation

a) It should be emphasized that this formula is particularly


important in unconventional shale reservoirs that are (1)
characterized by much higher values of intrinsic anisot-
ropy than most conventional shales have (Vernik and Liu,
1997; Sondergeld et al., 2000) and (2) frequently sand-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

wiched between quasi-isotropic rock formations, such


as carbonates, thereby resulting in marked anisotropy
contrasts.
Figure 13 compares results from the isotropic two-
term version of the Fatti equation with those from the
isotropic and anisotropic three-term versions, when ap-
plied to a shale/oil-saturated-sand interface. The starting
set of seismic rock properties is taken from Tables 1
(AVO Class III), 3 (Class IV), and 4 (Class II). All these
different sands and sandstones are assumed to be isotro-
pic and all the encasing shales to be anisotropic, with the
same Thomsen’s parameters of ε equal to 0.16 and δ
equal to 0.07 (as measured on shale core samples re-
b) trieved from a deepwater Gulf of Mexico well; see
Vernik et al., 2002).
Noting that Rüger’s (1998) anisotropic adjustment to
the Fatti equation 10 results in an increasingly negative
deviation from the isotropic three-term formula for these
shale-over-sand interfaces, we observe that the isotropic
two-term Fatti curve for RP(θ ) can be a more accurate rep-
resentation of AVO reflectivity in sand/shale sequences
than can any three-term approximation that fails to ac-
count for VTI. This curious observation is especially true
for the negative density contrasts that are typical of Class
III and Class IV sands.
Kim et al. (1993) investigated similar shale-over-
sand models, using the exact plane-wave anisotropic re-
flectivity expressions of Daley and Hron (1977).
However, their selection of the shale-anisotropy param-
eters (ε = 0.12 and δ = 0.13) is not consistent with the
Figure 11.  Examples of Class I and Class IIp AVO responses rule of thumb used today and based on accurate labora-
in (a) an AI-SI template, and (b) as AVO curves. The shale/ tory measurements of polar anisotropy in shale cores
wet-sandstone combination (blue dipole) is seen to gravitate (suggesting that ε > δ and, moreover, ε ≈ 2δ ). That
toward the Class IIp sector and to result in a polarity reversal. choice resulted in their anisotropic AVO curve deviating
Black dipole indicates shale-on-shale reflection with positive negatively, beginning at very small angles of incidence.
impedance contrast. Nonetheless, when comparing relative positions of the
reflectivity curves obtained by using exact isotropic and
anisotropic solutions with the various approximations of
where RP(θ)iso is given by the Aki-Richards equation 1 or those solutions, it is indeed curious that the isotropic
the Fatti equation 4. For instance, with respect to the Fatti two-term Fatti formula could fortuitously be the best ap-
approximation, one obtains: proximation of the exact anisotropic-reflectivity equa-
tion under certain conditions — such as weak anisotropy
or a small anisotropy contrast, Classes III and IV sands,
1 ∆A I  1 ∆A I ∆SI 1  2 and incident angles smaller than 35°.
R (θ ) ≈ + cos −2θ − 4K + ∆δ  sin θ
2 AI  2 AI SI 2  We should always remember another important point
related to VTI effects in synthetic forward modeling and
1 ∆ρ 1 seismic-gather analysis: Isotropic assumption-based
− (tan 2θ − 4 K sin 2θ ) + ∆ε sin 2 θ tan 2θ .
2 ρ 2 offset-to-­angle mapping computes greater angles of inci-
(10) dence for the same offset than is the case for an

04-Vernik_Ch04.indd 98 12-08-2016 20:01:36


Chapter 4: AVO Analysis: Rock-physics Basis 99

a) Vs Sl
0.5 km/s 5.5 0 (km/s) (g/cm3) 10
RHOB Incident angle
Vp Al
1.65 g/cm3 2.65 0.5 km/s 5.5 0 (km/s) (g/cm3) 10 0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)

4000
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4050

4100

b) Vs Sl
0.5 km/s 5.5 0 (km/s) (g/cm3) 10
RHOB Incident angle
Vp Al
1.65 g/cm 3
2.65 0.5 km/s 5.5 0 (km/s) (g/cm3) 10
0 3 6 9 11 14 17 20 23 26 29 31 34 37 40
Time (ms)

4100

4150

Figure 12.  Plane-wave synthetic angle gathers for two models described in Table 4 and shown in Figure 11a. (a) A wet sand
with Class IIp behavior; (b) a shale-on-shale event with a Class I response. A 25-m-thick sand layer and a 5/10/40/50-Hz
wavelet are used in both cases.

anisotropic overburden (Tsvankin, 2001). Again, the


angle mismatch is mostly a function of the strength of
shale anisotropy in the overburden. On the contrary, this
effect can be offset by layering-induced anisotropy in
thinly laminated reservoir-sand facies, which reduces the
anisotropy contrast between such facies and overlying
shales.
In certain parts of the world, such as offshore equato-
rial West Africa, the VTI of clay-rich conventional shale
sequences can be quite high (Gratwick and Finn, 2005;
Shumaker and Vernik, 2008). Thus, further investigation
and application of VTI to AVO-attribute studies is defi-
nitely warranted. Because not all shales are created equal,
it is important to note that occasionally even geographic
locations in close proximity may also require different
shale VP-VS transforms and, hence, modified AI-SI tem-
plates that have slightly steeper shale lines. The latter are
typical of those shales encountered offshore equatorial
West Africa and adjacent provinces causing them to be
Figure 13.  Class IIp, III, and IV AVO curves for a shale- quite distinct not only in terms of their VTI strength but
over-oil-sand interface, derived using the isotropic and also in terms of the impedance template construction and
anisotropic Fatti equations. AVO prediction (Chapter 3).

04-Vernik_Ch04.indd 99 12-08-2016 20:01:39


100 Seismic Petrophysics in Quantitative Interpretation

The fluid factor in prospect same sand/sandstone in a fully water-saturated state.


Nonetheless, in practical computations of the fluid-factor
risk mitigation attribute from prestack seismic data, the assumption men-
Among the numerous AVO attributes, the fluid factor tioned above is made — that wet sandstones are not dis-
tinguishable from associated shales in impedance space.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

is widely accepted to be simultaneously one of the most


instructive and most meaningful attributes (Smith and Then a scalar optimizing the background trend in RP-RS
Sutherland, 1996; Castagna et al., 1998). It was first intro- space is calculated for the seismic time window of inter-
duced by Smith and Gidlow (1987) and then reformulated est. However, because in the general case the wet-sand-
by Fatti et al. (1994) in terms of the zero-incidence reflec- stone line in VP-VS (or AI-SI) space actually deviates from
tion coefficients for P- and S-waves: the shale line, relatively high absolute fluid-factor values
can be attributed to wet sandstones. This is especially true
 SI  at the medium to low porosities where, with diagenesis,
∆F = R P − M  R , (11)
 A I  S
the wet-sandstone line deviates increasingly from the
shale line. Incidentally, a similar observation was first
made by Fatti et al. (1994) in their case history from the
where M is the slope of the linear AI = f (SI) transform in
Mossel Bay gas field (offshore South Africa).
a limited impedance range. This useful attribute is tradi-
To examine features related to the fluid-factor attri-
tionally referred to as the fluid factor (ΔF), even though
bute, let us analyze four sand/shale interface dipoles repre-
the name was originally derived from the notion that all
senting each of the AVO classes: I, II, III, and IV in AI-SI
wet sands, sandstones, silts, and shales tightly and univer-
space (Figure 14). Again, to emphasize that the fluid factor
sally fall along the so-called mudrock line of Castagna
as a seismic attribute may have little to do with actual res-
et al. (1985). In that case, the ΔF of all these wet clastics
ervoir fluids, the latter are assigned arbitrarily so that
is reduced to zero or to a relatively small value.
Class I is represented by a wet sand, Classes II and III by
Consequently, an increase in ΔF is assumed to indicate
a gas sandstone, and Class IV by an oil sand. For simplic-
the presence of hydrocarbons.
ity, all four dipoles share the same shale properties given
This notion is still quite popular, and the empirical
in Table 4 for the upper layer, so that the respective vectors
mudrock line is often used to guide computation of the
in the AI-SI plot radiate from the same point.
slope M to be approximately 1.16, without limitation on
Even though a wide dynamic range is covered by the
the dynamic range of impedance. On the basis of what we
sand/shale and sandstone/shale pairs, it should be noted
have learned so far in this book, such a notion can be quite
misleading: At very low impedances, the shale and wet-
sand lines indeed approach each other and even cross
over. However, at high impedances in medium- to low-
porosity, ­cemented sandstones, the lines tend to diverge
quite significantly (e.g., Figure 1).
Note that equation 11 can also be expressed as
AVO class
1  ∆A I  SI  ∆SI  ∆ A I − M ∆SI
∆F =  −M = . (12)
2  AI  A I  SI  2A I

That is, the fluid factor can be conveniently visualized in Shale
the AI-SI space, and its computation can be simplified by
placing a lower limit on shale’s acoustic impedance at ap-
proximately 5.5 g/cm3 ×  km/s (i.e., considering only
shales with VP > 2.4 km/s). Under that limitation, the
AI-SI relationship is indeed quasi-linear, with a slope of
approximately 0.66 for most conventional shales and
0.708 for some shales encountered in equatorial West
Africa. The respective M values range from 1.518 to Figure 14.  Illustration of the fluid-factor attribute, using
1.412; that is, they are much greater than the value of 1.16 horizontal projections of sand(stone) points onto the shale
implied by the mudrock line of Castagna et al. (1985). line. Fluid assignments among the cases with different AVO
According to Fatti et al. (1994), ideally ΔF is the dif- classes are arbitrary (red for gas, green for oil, and blue for
ference between the actual RP of a sand/sandstone with water) to emphasize that the fluid factor as an AVO attribute
arbitrary fluid saturation and the calculated RP for the generally combines the effects of fluid and lithology.

04-Vernik_Ch04.indd 100 12-08-2016 20:01:40


Chapter 4: AVO Analysis: Rock-physics Basis 101

that at least the shale line, which is representative of the incomplete without taking into account the gross thickness
background trend, is quite linear over it. Therefore, the of the reflecting layer. This is a well-known problem of both
lengths of the horizontal projections of each sand(stone) full-stack and offset-dependent reflectivity below, at, and
point onto the shale line, according to equation 12, are above the tuning thickness or vertical resolution, which, ac-
proportional to their fluid-factor values (Figure 14). cording to Widess’s rule, is usually taken to be one-quarter
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Indeed, all four cases produce measurable fluid-factor of the dominant seismic wavelength (Widess, 1973).
anomalies. Moreover, the oil-sand case with Class IV The seismic-resolution thickness can, under the ideal
AVO behavior results in fluid-factor values that are not circumstances of large elastic contrasts and a high signal-
dramatically different from those of the wet ­sandstone to-noise ratio, be as low as (VP/fc)/8, where fc is the domi-
with Class I (transitioning to Class II) AVO response. nant frequency of the reflection’s spectral bandwidth.
It is important to note that all four of these cases as- Below one-eighth of a wavelength λ –– that is, for rela-
sume homogeneous sand/sandstone layers with net-to- tively thin beds (<15 m thick, if  fc = 25 Hz and VP = 3 km/s),
gross ratios (N/G) of one, gross thicknesses of no less than the two-way traveltime “thickness” of a layer becomes con-
the vertical resolution of the seismic data set analyzed, and stant and no longer is affected by the actual thickness of the
high hydrocarbon saturation when applicable. These mean layer. At that point the only characteristic that continues to
(upscaled) sand/sandstone points in P- versus S-impedance change below this tuning thickness is the seismic amplitude
crossplot are expected to migrate toward the shale point if (Widess, 1973). Obviously, this resolution limit usually
thin shale beds or laminations are introduced into the gross worsens under less ideal conditions. For the dominant fre-
reservoir interval, and that will reduce the interval’s N/G quency and velocity combination mentioned above, it can
value. The same migration is expected to happen if gross reach 20 to 30 m — that is, 0.25 λ.
thicknesses are smaller than the tuning thickness of ap- Even though various vertical-resolution-enhancement
proximately 0.25 λ, where λ is the seismic wavelength. methods have been proposed (e.g., Partyka, 2007) that use
Another complicating factor that deserves more atten- standard 3D seismic acquisition for targets at medium and
tion is inhomogeneous hydrocarbon saturation. This situa- greater depths, their reliability in terms of quantitative
tion naturally manifests itself in compositionally inhomo- thickness prediction is yet to be adequately documented
geneous reservoirs by creating a layered saturation profile worldwide in real case histories of prestack seismic inver-
with moderate to low overall oil and gas saturation. sion or AVO analysis. Consequently, here we concentrate
Similarly low hydrocarbon saturation phenomena can also on the reflection amplitudes generated by a specific strati-
be a result of seal leakage, whereby reservoir saturation graphic layer with a changing gross thickness. This layer
may be reduced to a residual level. The latter case is often can be homogeneous or heterogeneous on the log scale.
loosely referred to as a “fizz-gas” reservoir. As we demon- However, it is effectively homogenized on the scale of
strated before, the hydrocarbon sand/sandstone points in tuning thickness during the process of convolution, so
such situations will migrate toward their wet counterparts, that it can be represented by its effective seismic rock
effectively reducing the fluid factor. However, the magni- properties having been reduced to focal points on the
tude of this reduction is, to a large degree, prescribed by the AI-SI template. Those focal points should be generally
sand porosity and/or stiffness. The effect in high-porosity consistent with the upscaled impedances of sands/sand-
sands potentially can be much greater than in low-porosity stones and shales.
sandstones, where the separation among the gas-, oil-, and In the following discussion, to better understand
wet-sandstone lines in AI-SI space is diminished. thickness effects let us consider simplified wedge models
Considering all of these complications, we can con- of sands/sandstones that have different seismic rock prop-
clude that although the fluid-factor concept is the most erties and are embedded in homogeneous shale. A zero-
robust among AVO attributes, it still needs to be carefully phase trapezoidal wavelet of 5/10/40/50 Hz, with a 25-Hz
quantified and calibrated. That can be accomplished by central frequency, is employed throughout the modeling
using seismic petrophysics methods in each specific oc- routine. Seismic rock properties of shale and two differ-
currence in an attempt to distinguish between the fluid ing gas sands, causing Class III and Class II AVO re-
and lithology contributions; only then can the fluid factor sponses, are shown in Table 5.
be used as a true DHI in oil and gas exploration projects. As is illustrated in Figure 15, which is constructed
similarly to Figure 14, these rock properties are selected
by using the most typical shale AI-SI relationship in addi-
Tuning effects in tion to that established for arenite sands, at an effective
AVO synthetic modeling stress of 20 MPa and with a critical porosity of 38% and a
consolidation porosity of 26% (see Figure 38a of
A review of a seismic reflection’s AVO behavior due to Chapter 3). The gas density is 0.25 g/cm3, and the gas
rock-property contrasts across an interface would be modulus is 0.25 GPa. Note that the effective stress of

04-Vernik_Ch04.indd 101 12-08-2016 20:01:40


102 Seismic Petrophysics in Quantitative Interpretation

Table 5.  Seismic rock properties of gas sands and shale used in AVO wedge modeling.
Lithology ρb VP VS AI SI
(g/cm3) (km/s) (km/s) (g/cm3 × km/s) (g/cm3 × km/s)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Gas sand, Class III 2.05 2.80 1.77 5.72 3.61


Gas sandstone, Class II 2.13 3.40 2.18 7.25 4.63
Shale 2.45 2.92 1.37 7.15 3.35

respect to the shale shown in Figure 15, these two hydro-


carbon reservoir models produce Class III and Class II
AVO responses.
The Class II sandstone is consolidated, and its poros-
ity is slightly lower than the consolidation porosity men-
tioned above. Because its P-wave reflectivity is close to
zero, the synthetics look “transparent” at near angles of
incidence. It is also noteworthy that for this reason, the
AVO gradient in this case is significantly greater than in
the Class III case.
When the sand thickness is reduced to less than 6 m,
both near and far amplitudes in the synthetic gathers be-
come weak and will be difficult to discern if noise is
added to make the synthetics look even more like prop-
erly conditioned CDP seismic gathers. The minimally de-
tectable thickness (for the specified impedance contrast),
which can be referred to as the seismic detection limit, in
this case is likely to be approximately 5 m. For the two
Figure 15. An AI-SI template, showing AVO cases for two cases modeled here, this limit seems to be on the order of
different gas-saturated sands(tones) encased in the same shale 0.04 λ (i.e., 1/25 of the dominant wavelength).
(Table 5). The horizontal segments are proportional to the Tuning curves for the P-wave reflectivities and fluid
fluid-factor values associated with these two lithologies. factors of these two sand-wedge models are presented in
Figure 17 and clearly indicate that their fluid factors are
20 MPa is typical of the medium depths below the mud nearly the same, which is consistent with the sand(stone)
line (2 to 3 km of TVDml) in normally pressured subsur- point projections onto the shale line as is shown in
face settings, and of greater depths in overpressured ones, Figure 15. It should also be emphasized that the tuning
so it can be used as a reference stress level in modeling. effect on the full-stack amplitude is stronger for the Class
It should be emphasized that the fluid factors depict- II AVO case than it is for the Class III case. That is, detec-
ed in Figure 15 correspond to sand layers with a net-to- tion of thin, gas-saturated sand beds showing a Class II
gross ratio of one and with a gross thickness exceeding response without prestack data analysis is expected to be
the vertical resolution given by the tuning thickness. more challenging than detection of Class III gas-saturated
According to the rule of one-quarter of the wavelength, sands would be.
the gross thickness computes as 29 m in the Class III case Finally, the wedge modeling of the two typical DHI-
and 32 m in the Class II sand case. Thus, if the sand thick- prospect sands indicates that the fluid-factor attribute is
ness exceeds 29 and 32 m in our two respective occasions, quite sensitive to the thickness of the sand(stone) layer,
it is expected that the top sand interface will have AVO especially below tuning thickness, which seems in these
characteristics (intercept, gradient, RP, RS, and ΔF) pre- ­models to be slightly lower than the one-quarter-wave-
scribed by Table 5. length rule. A similar sensitivity to sand(stone) thickness
Figure 16 shows AVO wedge models of these two can be ­established for the AVO gradient/intercept combi-
sand layers, with the thickness increasing from 6 to 36 m, nation. However, because the fluid factor presents a com-
in 6-m increments. The models are composed of the syn- bination attribute involving both the AI and the SI of the
thetic angle gathers constructed on the basis of the seis- sand(stone) layers, as well as their relationship with the
mic rock properties given in Table 5. As is expected from shale-dominated background trend in the AI-SI crossplot,
the AI-SI positions of the sand and sandstone points with fluid-factor stacks should be given precedence in

04-Vernik_Ch04.indd 102 12-08-2016 20:01:41


Chapter 4: AVO Analysis: Rock-physics Basis 103

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b)

Figure 16.  AVO wedge models for two gas sand(stones) encased in shale (Table 5), illustrating (a) Class III AVO behavior and
(b) Class II behavior. The sand thickness is increasing from left to right, in increments of 6 m, beginning at 6 m and ending at
36 m (20 to 120 ft). On each angle gather, the incident angles increase from 2 to 44°.

a) b)

Figure 17.  P-wave reflectivity and the fluid factor as functions of sand thickness for (a) AVO Class III and (b) AVO Class II.

exploration settings before simultaneous impedance in- (Figure 18). An oil-reservoir penetrated by three wells
version is undertaken. On the other hand, the P-wave re- varied from 6 to 10 m in thickness, and synthetic and
flectivity’s sensitivity to layer thickness is basically re- CDP seismic gathers both revealed a great degree of vari-
duced to zero in Class II sands and can be minor in Class ability, from Class III to Class I AVO responses, laterally
III and Class I situations adjacent to the Class II sector in across the field. The 3D seismic gathers were conditioned,
the AI-SI AVO classification space (Figure 14). muted at an incident angle of approximately 35°, phase
An interesting example of sand detection below the rotated by 90° to attempt to enhance the vertical resolu-
tuning thickness comes from the North Sea province tion, and then stacked.

04-Vernik_Ch04.indd 103 12-08-2016 20:01:45


104 Seismic Petrophysics in Quantitative Interpretation

a)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b)

Figure 18.  (a) A seismic section through a 90° phase-rotated full-stack 3D seismic volume, and (b) a fluid-factor-based
thickness AVO attribute, from a thin oil-sand reservoir in a North Sea province. The AVO attribute provides a more accurate
means of mapping this highly heterogeneous, thin reservoir. Red colors in (a) correspond to softer impedances.

A simplistic interpretation using the full-stack vol- tuning, a starting point in the workflow is to compute that
ume (or band-limited acoustic-impedance inversion) turns gradient over the entire field.
out to be erroneous because the laterally variable concen- The AVO-gradient-based attribute (shown in Figure
trations of cemented lenses in the sand cause its upscaled 18b) is calibrated to read sand thicknesses above and below
impedances at some locations to exceed the impedances the OWC; it not only unambiguously maps the top/base res-
of the shale overburden. In turn, shale velocity and acous- ervoir in space but also furnishes information for assessing
tic impedance controlled by the highly variable clay con- reservoir connectivity and aquifer support. This challenging
tent also show heterogeneity. case history presents compelling evidence of the power of
As a result, Figure 18a reveals that top/base sand ho- AVO-based attributes and demonstrates why they must be
rizons cut across the trough (red) and peak (blue) ampli- used more routinely in exploration, d­ evelopment, and pro-
tudes, causing a great deal of confusion. Indeed, the seis- duction stages, especially in thin reservoirs.
mic trough with red amplitudes can represent either soft, Use of the fluid factor in general or of the AVO gradient
high-porosity oil sands or soft, high-vcl shales onlapping in a special case of thin-bed thickness detection should al-
laterally. On the other hand, the blue peak amplitudes on ways be given preference over a full-stack tuning-curve ap-
Figure 18a can indicate either wet sands below the oil- proach. However, one must keep in mind that any quantita-
water contact (OWC) or rigid, calcite-cemented sand- tive prestack attribute analysis (including analysis of the
stone lenses and concretions responsible for Class I AVO fluid factor) in terms of lithology or saturating fluids will
responses even above the OWC. necessarily suffer when we are dealing with gross thickness-
It is obvious that some prestack attribute had to be es below one-eighth of the dominant seismic wavelength.
relied upon in this specific field in order to tie the wells, Additionally, because tuning can affect near and far offsets
pick the top/base reservoir horizons, and map the reser- differently as a result of NMO stretch and a related change in
voir thickness for oil-in-place computation. Because AVO frequency content, this phenomenon in some cases may lead
gradient is most sensitive to reservoir thickness below to AVO distortion and even to false AVO anomalies.

04-Vernik_Ch04.indd 104 12-08-2016 20:01:46


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 5: Simultaneous AI-SI Inversion and N/G Computation

The limitations of using conventional 3D seismic data reservoir’s net-to-gross and net-pay values. Integration of
for detailed reservoir characterization are well recognized seismic petrophysics (notably, AI-SI analysis of well-log
(e.g., Widess, 1973; Brown et al., 1984). These limitations data) with calibrated prestack inversion performed by
are related primarily to vertical and lateral resolution, the using well-conditioned CDP gathers is clearly demon-
seismic detection limit, and, in some cases, poor ampli- strated to be the key to success. The proposed estimation
tude fidelity and image quality. Despite such difficulties, methodology is described below with a case-study exam-
seismic reservoir characterization (also known as quanti- ple (Vernik et al., 2002).
tative interpretation or QI) can be a powerful tool for un-
derstanding and predicting reservoir quality. The key word
here is integration. And the key to properly integrating Log editing and the AI-SI template
seismic data with geologic and engineering information is
adequate conditioning of log data and application of ap- The reservoirs discussed in this section are from the
propriate rock-physics tools, both of which are the focus deepwater Gulf of Mexico (USA) and are Tertiary in age.
of seismic petrophysics (see Figure 23 of Chapter 1). On average, the sands are poorly consolidated and were
Interestingly, 3D seismic is often well suited for quan- originally deposited in a deepwater channel/levee system,
titative estimates of the net-to-gross ratio (N/G), which is a with relatively massive channel facies and laminated
critical reservoir-scale property in deepwater turbidite res- sand/silt/shale levee facies. There are two distinct oil res-
ervoirs. Because deepwater turbidites have inherent hetero- ervoirs separated by approximately 250 m of shale inter-
geneities and significant lateral variations in N/G values, it spersed with thin gas-saturated sands. The hydrocarbons
is important that we quantify this parameter before we try to are stratigraphically trapped, occur in the depth interval
gain further insight into other important reservoir properties spanning 4.0 to 4.3 km (the water depth is 1.6 km), and
(such as porosity) between the appraisal well locations. display a high degree of facies and consolidation hetero-
Since the ­publication of Brown et al.’s (1984) paper, ver- geneity, which tends to increase in the lower sand. The
tically integrated seismic amplitudes and/or poststack and reservoirs are only slightly overpressured, are subjected
prestack impedance inversions of seismic data have been to a vertical effective stress of approximately 20 to
used routinely to estimate N/G in many clastic reservoirs, 22 MPa, and are saturated with 32 to 35°API oil that ex-
with varying success. Indeed, when the shale’s baseline im- hibits significant variations in GOR (125 to 250 m3/m3).
pedance can be estimated, a departure from that trend has The log-based N/G evaluation, which includes litho-
been tentatively argued to represent a certain measure of the facies classification that uses a laminar shale volume Vsh
hydrocarbon-saturated sand volume on the reservoir scale. as the major turbidite facies identifier, is developed with-
Vernik et al. (2002), among others, have taken this technique out any input from sonic logs. Therefore, a special effort
a step farther by using both the acoustic-impedance and using seismic petrophysics is required to marry the rock
shear-impedance volumes (AI and SI) to employ those in- properties to the high-quality 3D seismic data set. The
verted data and compute the sand-fraction volume for a res- necessary steps in the process, including log anisotropy
ervoir in the deepwater Gulf of Mexico. In their study, as correction and examination of VP-VS space systematics,
many as six appraisal wells were available for calibration are briefly described below.
prior to the reservoir’s development and production. After establishing the shale and oil-sand diagenesis
Further discussion of the methodology for estimating trends employed in a calibrated AI-SI template and then
N/G is presented in this chapter by focusing on its advan- identifying their relationship with ­lithology and fluid vec-
tages and limitations. Simple strategies are proposed for tors, simultaneous AI-SI prestack inversion of the 3D seis-
improving that methodology, and additional modeling is mic is applied to generate maps of the oil-charged-sand-
carried out to shed more light on seismic expressions of a fraction volume (N/G) and net pay. As a reminder, lithology
105

05-Vernik_Ch05.indd 105 12-08-2016 20:59:06


106 Seismic Petrophysics in Quantitative Interpretation

vectors correspond to shale-over-wet-sand(stone) interface Following the workflow described in Chapter 1 (see
dipoles, whereas fluid vectors, which connect specified wet also Figure 29 of Chapter 1), the modeling of anisotropic
reservoir data points in AI-SI space with those at different effective layered media is performed to evaluate the effec-
fluid saturations (oil and gas), are always subhorizontal. tive properties of the deepwater levee facies. These facies
Due to the bulk-density effect on shear impedance fluid are composed of laminated, very-fine-grained oil sand
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

vectors always have a slight positive slope in AI-SI crossplot. (transitioning oil-saturated silt) and wet shale (e.g., Figures
A proper account of the relative positions and orientations 9 and 10 of Chapter 1). Each levee facies has its own petro-
of the diagenesis, lithology, and fluid vectors is instructive physical signature, with a core-calibrated, laminar Vsh value
in any QI workflow. ranging from the approximate median values of 30% in
In the text below emphasis is given to the lower of the proximal levees to 50% in medial levees and 70% in distal
two stratigraphic intervals representing our case history, levees. Obviously, the levee-dominated field locations are
which is designed to illustrate a methodology for convert- characterized by lower N/G values. The thickness of sand
ing the impedance volumes into the N/G attribute by using laminae in these rocks varies from 1 to 40 mm — that is, it
AI-SI space with built-in templates. is more than an order of magnitude smaller than the wave-
lengths that are typical of any sonic-log acquisition.
Anisotropy correction of sonic velocities The anisotropic Backus model (Backus, 1962), with
end-member velocities and shale anisotropy (Figure 1), pro-
In anisotropic subsurface the general rule is that verti- vides the best fit to the log-facies average values for Vsh, VP,
cal velocities should be used to achieve a good tie between and fast VS (VSH-mode-polarized in the bedding plane). Note
the log-based synthetics and the seismic a­ mplitudes. This that these values were obtained separately for straight holes
also facilitates proper construction of a low-frequency and deviated-well subsets (with relative dips near 63°).
background model, which is necessary for absolute imped- The modeling reveals peculiarities of the elastic proper-
ance inversion. Unfortunately for vertical-velocity model- ties of sand and shale layers — for example, the sand lami-
building, in many offshore appraisal and development proj- nae in thin-bedded turbidite facies have a lower shear rigid-
ects it is common practice to drill several highly deviated ity (lower VS) value and a higher VP/VS value than do the
wells, in addition to a few straight holes, to fully delineate blocky channel sands (note the mismatch at small Vsh values
and understand the reservoir. In relatively simple and gen- between the model and the actual measured VSH in channel
tle structural settings with only minor bed dips, the devia- sands in Figure 1b). This observation agrees well with core
tion angle and azimuth of a well determine the relative dip velocity data and can be explained by the ever so slightly
angle of those beds. As was discussed in Chapter 1, that better grain bonding (consolidation) found in generally
relative dip between the wellbore and the bed dip controls coarser-grained channel sands, compared with that in very-
the sonic-log velocity values in layers that can be approxi- fine-grained and more clay-rich laminated sands/silts.
mated by transverse isotropy (TI). Assuming that the shale anisotropy across the field is
In the reservoir of interest, sands are observed to be consistent with the range indicated by the laboratory mea-
quasi-isotropic, whereas the intervals dominated by thin surements on core samples mentioned above, the best-fit-
beds, laminated levee facies, and massive shale intervals ting anisotropy model for parameters ε, δ, and γ as functions
display a strong relative dip dependence of both P- and of the laminar Vsh is constructed from the data presented in
S-wave velocity. To better understand this phenomenon, Figure 1. As expected, the maximum amount of P-wave an-
Vernik et al. (2002) resorted to core measurements, which isotropy is predicted for massive shales with higher clay
indicated a moderate but variable amount of intrinsic content, followed by the lithologies with a high shale con-
polar anisotropy in those relatively compacted mudrocks tent, such as distal-levee facies (wherein the mudrock com-
with total porosity of 15 to 18%; the P-wave anisotropy ponent is often represented by marly shales or marls). The
parameter ε was found to vary from 0.05 to 0.16. effective anisotropy of the gross reservoir is also expected
This range of values corresponds to compositional to increase when the nonreservoir component is represented
variations, both vertically and laterally, from marls with by more-anisotropic shales with higher vcl values. Note that
lower clay content and weaker anisotropy to shales with vcl the models presented in Figure 1 are consistent with those
values averaging 55% and respectively higher anisotropy included in Figures 28 of Chapter 1 illustrating the angular
values due to preferred alignment of clay platelets in the dependence of P- and S-wave velocities in laminated reser-
bedding plane. The main agents impeding this alignment voirs and massive shales.
are individual silt grains and grain aggregates composed The Vsh functions of the three anisotropic parameters
of quartz and carbonate minerals. In particular, the higher are necessary for anisotropy correction of the cross-di-
impedance and less anisotropic marls are ubiquitous above pole sonic logs that were run in the deviated wells. The
and inside of the reservoir interval penetrated at a relative significance of the anisotropy correction (“verticaliza-
dip of 63° by one of the most deviated wells in the field. tion”) of the velocity logs and its dependence on Vsh can

05-Vernik_Ch05.indd 106 12-08-2016 20:59:06


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 107

Table 1.  Lithologic facies definitions for a deepwater


reservoir, using log-derived Vcl and Vsh curves.
Lithologic facies Vcl cutoff Vsh cutoff
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Channel sands (arenites) Vcl < 0.12 Vsh = 0


Splay sands and proximal n/a 0 < Vsh < 0.4
levees
Medial levees n/a 0.4 < Vsh < 0.6
Distal levees n/a 0.6 < Vsh < 0.9
Shales and marls Vcl > 0.3 Vsh = 1

Additional log editing


Comparisons between the “verticalized” sonic velocity
(with a 3- to 8-kHz frequency range) and the vertical-seis-
mic-profiling (VSP) interval velocity (with an 80- to 130-Hz
frequency range) in the reservoir of interest indicate a mini-
mal amount of frequency dispersion similar to the upper por-
tion of another well drilled in deepwater West Africa and
shown in Figure 24 of Chapter 3. Because oil-based mud
system was used to drill all the appraisal wells in the field,
mud-invasion effects are negligible above OWC, but not so
for the wet sands below OWC, where mud invasion effect is
detectable. Therefore, no additional corrections to the sonic
velocities need to be undertaken, but the density log is cor-
rected for oil-based-mud invasion prior to its use in log-
based AI and SI calculations (refer to Chapter 1).

AI-SI crossplot
Edited and verticalized well-log-based acoustic and
shear impedances are computed for all six appraisal wells,
including four wells with significant degrees of deviation
from the vertical. This well database is subdivided into lith-
Figure 1.  Average laminar-shale-fraction Vsh versus (a) VP ologic facies by imposing Vcl and Vsh cutoffs, as in Table 1.
and (b) fast-shear VSH velocities for five petrophysical facies Figure 2 shows a simplified version of the AI-SI
from two straight holes (red) and one highly deviated well ­template, which uses linear approximations to the Sand
(green), superposed by the best-fit Backus models. Adapted Diagenesis Model line computed for oil-saturated arenites
from Figure 2 of Vernik et al. (2002). Used by permission. with irreducible water saturation and Shale Model line in a
short dynamic range of interest. At the high-impedance end,
be appreciated by ­inspecting Figure 29 of Chapter 1, this linear sand model may or may not, depending on the
which shows a significant polar anisotropy effect even in effective stress, extend to the matrix properties of arenites
a well with a moderate relative dip of 47°. reported in Table 2 of Chapter 1; however, the shale line
On the other hand, the anisotropy contrasts (Δε and should always reach the clay-content-dependent bedding-
Δδ) may be rather weak between the overlying shale, with normal matrix properties of shales (given in the same table).
a Vsh value of one, and the layered reservoir, which has a The most remarkable feature of this edited and verti-
reduced N/G value. Such a minor contrast in anisotropy is calized data set is that, over the impedance range of inter-
unlikely to produce significant errors when we are using est, the slope of the linear regression through the oil-
synthetic AVO models based on isotropic Zoeppritz, Aki- saturated channel sands from appraisal wells is approxi-
Richards, or Fatti equations, all of which are frequently mately 0.70, which is the same as the slope of the Sand
employed in prestack-impedance-inversion software Diagenesis Model for arenites with the same saturation
packages. state. Moreover, it is practically the same as the slope of

05-Vernik_Ch05.indd 107 12-08-2016 20:59:07


108 Seismic Petrophysics in Quantitative Interpretation

SI − bAI − a0 (1)
Vsand ≈ ,
a1 − a0
where b is the average slope of the shale and hydrocarbon-
sand lines (with the general slopes of b0 and b1, respec-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tively), and a0 and a1 are their respective SI-axis intercepts


in AI-SI space (in impedance units). In our case, the shale
and oil-sand intercepts are –1.44 and –0.64 g/cm3 × km/s,
respectively, and b = 0.67. In the general case, these coef-
ficients are expected to vary with hydrocarbon fluid prop-
erties and, albeit less so, with the amount of overpressure.
Equation 1 virtually represents the process of axis ro-
tation in AI-SI space, aligning the abscissa to be parallel
to the background-trend (shale) line. Quakenbush et al.
(2006) described a similar axis-rotation process to com-
pute another attribute referred to as “Poisson i­ mpedance”
or PI. The Poisson impedance is a linear c­ ombination of
AI and SI; specifically, PI = AI – (1/b)SI. It is easy to dem-
Figure 2. An AI-SI template, showing major lithofacies and
onstrate that in PI terms equation 1 can be expressed as
end-member parallel trends in agreement with shale (Vsand = 0)
and oil-sand (Vsand = 1) well-log impedance data. Note that the PI 0 − PI
Vsand ≈ , (2)
shales and marls are shown uniformly in black even though PI 0 − PI 1

they tend to differentiate along the shale line, with marls
being characterized by greater impedances than are the where PI0 = –a0/b and PI1 = –a1/b are the Poisson imped-
less-carbonate-rich and more-clay-rich shales. Adapted ances of shale and sand, respectively. Equation 2 makes
from Figure 7 of Vernik et al. (2002). Used by permission. it clear that the Poisson impedance and the sand/sand-
stone volume on seismic scale are closely related; how-
ever, it is the Vsand attribute that provides for more intui-
the normal-to-bedding shale-line approximation given by tive connection with reservoir composition, including
SI = 0.67AI – 1.44 (in km/s). These observations are typi- net-to-gross, as demonstrated below.
cal of siliciclastic sequences worldwide with an exception Integration of the sand fraction Vsand over the gross
of very poorly consolidated clastics, where dramatic non- reservoir thickness (ΔZ = Ztop – Zbase) leads to an estimate
linearity of both sand and shale models in the impedance of N/G for the specific well or any x-y location where AI
space has been established in Chapter 3. and SI data are available:
Each lithofacies with varying levels of consolidation Z base
above OWC appears as an elongated cluster that finds its
own place in the AI-SI space of Figure 2. Those clusters ∫V
Z top
sand dZ
(3)
are either consistent with or sandwiched between the N G= .
∆Z
subparallel channel sand and the bedding-normal shale
lines of the template. The only exception is the water- This computation can be performed in either the time or
saturated sands from the aquifer and “perched water” depth domain. The same principle applies to the upscaled
zones inside the reservoir, which naturally plot along the Vsand curve or the inversion-derived Vsand attribute volume
wet arenite line that cuts across the shale line (but is not because the boundary conditions given by the hydrocar-
included in Figure 2). The scatter of oil-saturated-sand bon-saturated sand(stone) and shale baselines do not
data points and shale data points around the respective change, as those are prescribed by the global Sand(stone)
model lines may be related to variations in fluid proper- Diagenesis Model and Shale Model.
ties and composition, slight depth mismatches among Of course, equation 1 is a gross approximation and
VP, VS, and density logs, differences in the depth of inves- should be treated as a purely empirical means of obtain-
tigation, differences in vertical resolution, and process- ing the sand fraction over the gross reservoir interval from
ing artifacts. seismic data. Because this formula is a linear interpola-
The oil-sand fraction (Vsand = 1 – Vsh) above the oil- tion between the two known linear trend lines, the 50%
water contact on the log scale can be treated as the instan- N/G line occurs in the middle between the shale line and
taneous value of N/G. The subparallel trends for shale and a hydrocarbon-saturated arenite sand line. The accuracy
oil-sand end members in Figure 2 can be taken advantage of this computation will be tested in the following model-
of to develop a simplified linear interpolation equation: ing experiment.

05-Vernik_Ch05.indd 108 12-08-2016 20:59:08


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 109

Modeling N/G
The most straightforward theoretical method for
modeling the effect of reservoir-scale variations in Vsand
and/or net-to-gross ratio in siliciclastic hydrocarbon res-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ervoirs is provided by Backus averaging (Backus, 1962).


In that averaging, a two-component sand/shale composite
is built by using varying end-member elastic moduli, ve-
locities, bulk density, and, hence, AI and SI values pre-
scribed by the sand and shale diagenesis models (refer to
Figures 23, 25, and 39 of Chapter 3).
Because Backus averaging in the bedding-normal di-
rection represents a simple harmonic mean of elastic
moduli as a function of the sand fraction Vsand, the model-
ing results are anticipated to be different from the arith-
metic averaging of the sand and shale impedance lines
given by equation 1. The difference is expected to be the
most significant when the end-member component prop-
erties (or their trend lines) are as far apart from each
other as possible. Thus, it is appropriate to resort to mod-
eling gas sand instead of oil sand, as was done for the
reservoir in question.
A wide range of sand porosities and degrees of con-
solidation, as well as of shale compaction, are used to
arrive at three sand/shale pairs for each of the following
AVO classes: Class III, Class II, and Class IIp (Table 2).
Models 1, 2, and 3 in each AVO class correspond to shale/
sand dipoles with increasingly greater shale compaction
and sand consolidation. The following model parameters
are used for gas sand: The gas saturation is 80%, the crit-
ical porosity φc is 0.4, the consolidation porosity φcon is
0.23, the vertical effective stress σ is 20 MPa, the gas

Table 2.  Porosity and impedances (g/cm3 × km/s) of gas


sand/shale pairs used in Vsand modeling.
AVO Class Sand Shale
Class III φ AI SI φ AI SI
Model 1 0.30 4.60 2.84 0.20 6.32 2.81
Model 2 0.28 5.85 3.73 0.14 7.49 3.58
Model 3 0.26 6.74 4.31 0.11 8.33 4.13
Class II φ AI SI φ AI SI
Model 1 0.28 5.85 3.73 0.22 5.95 2.57
Model 2 0.26 6.74 4.31 0.18 6.69 3.06
Model 3 0.18 8.42 5.40 0.11 8.33 4.13
Class IIp φ AI SI φ AI SI
Figure 3.  P-wave reflection coefficient versus sin2θ of the
Model 1 0.26 6.74 4.31 0.22 5.95 2.57 incident angle, for nine shale-on-gas sand pairs used in
Model 2 0.26 6.74 4.31 0.18 6.69 3.06 modeling the N/G effect. Red, green, and blue correspond
respectively to models 1, 2, and 3 in each AVO class
Model 3 0.18 8.42 5.40 0.14 7.49 3.58 (Table 2).

05-Vernik_Ch05.indd 109 12-08-2016 20:59:09


110 Seismic Petrophysics in Quantitative Interpretation

modulus Kg is 0.2 GPa, the gas density ρg is 0.25 g/cm3, Figure 4 is an AI-SI plot illustrating the positions of
the water modulus Kw is 3.1 GPa, the water density ρw is gas sands (N/G = 1) and shales (N/G = 0), along with the
1.06 g/cm3, and the shale net clay volume vcl is 0.6. modeled N/G = 0.5 lines, where the modeled lines are
The respective half-space models showing shale-on- based on Backus averaging. This exercise assumes that
sand reflection coefficients are displayed in Figure 3a, the mean thickness of the sand and shale layers does not
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

3b, and 3c, and the end-member AI and SI values can be exceed 1/10 of the dominant wavelength, which is a nec-
identified in Figure 4. The AVO curves are built by using essary precondition for this type of effective-medium
the isotropic three-term Fatti equation, but the AVO gra- modeling. The AI-SI points for each of the nine 50%-N/G
dient is computed for a 0 to 20° spread of the angles of composite models (three for each AVO class considered)
incidence, where the gradient can be adequately repre- are arranged in the figure to be connected by straight dot-
sented by a linear approximation (see Chapter 4). As ted segments according to their AVO behavior.
Figure 3 shows, analysis of the changes in the conven- It is noteworthy that whereas on average the ­computed
tional AVO attributes (intercept and gradient) when the AI-SI positions of the 50%-N/G model points cluster
Vsand value decreases from 100 to 50% reveals two inter- around the midline between the shale and gas-sand lines,
esting observations: distinct and significant deviations from this midline occur
that are controlled by the AVO class. Those deviations are
1. The intercepts undergo a 50% reduction, in all three determined by the relative positions of the shale and as-
AVO classes modeled. sociated sand points (listed in Table 2) that form imagi-
2. The gradients decrease much more dramatically than nary dipoles in the AI-SI crossplot of Figure 4. Those di-
the intercepts do, by an average factor of 2.1 in Class poles are not shown intentionally, so the reader can draw
III, 3.5 in Class II, and 4.4 in Class IIp. the vectors on Figure 4 as an exercise by using the AI and
SI values from Table 2.
Consequently, a reduction in N/G values affects the AVO The deviations of the modeled 50%-N/G points from the
gradient in a way that is qualitatively similar to the effect midline are related to the contrast between the shear moduli
of gross-thickness reduction below tuning, which was (and shear impedances) of sands and shales. When that ΔSI
discussed in Chapter 4. It is also remarkable that the gra- contrast increases from AVO Class III toward Class IIp (as
dient is much more sensitive to N/G variations than the Fatti et al., 1994, predicted it would), the modeled 50%-N/G
intercept is. loci in AI-SI space gradually shift toward the shale line. This
is expected because the Backus model is a harmonic average
of elastic moduli, and it yields bedding-normal effective
properties of the composite that are partial toward the lower-
rigidity component. That component could be either shale
(in consolidated rocks more typically associated with Class
II AVO behavior) or sand (in poorly consolidated clastics
often associated with Classes III and IV).
As is evident from Figure 4, the N/G computation
from equation 1 can result in absolute errors as high as
15%. In consolidated sandstone reservoirs, we are likely to
overestimate N/G, whereas less-significant underestimates
are expected in less-consolidated sands. Those situations
require mitigation to improve the prediction ­accuracy of
N/G; that can be achieved by parallel repositioning of the
hydrocarbon-saturated sand line in AI-SI space to alter its
intercept a1 (refer to equation 1). How we do this depends
on the elastic contrast between the sands/sandstones and
shales in the reservoir in question. Note that even though
Figure 4. An AI-SI template, showing the zero-N/G shale the errors in seismic N/G estimates will be slightly lower
(black) and 100%-N/G gas-saturated arenite (red) impedance in oil-sand reservoirs than in gas sands, their model-line
points used to compute the model lines of the 50%-N/G gas intercepts and/or gradients in the AI-SI template may still
reservoir based on Backus (1962) averaging. Note that have to be adjusted to optimize the computation.
although on average the computed data points (purple, blue, As was mentioned above, the effect of N/G on AVO
and green dots) cluster around the “midline” trend between attributes resembles the effect of a reduction in gross sand
the shale and gas-sand lines, distinct and significant thickness below tuning thickness in a reservoir with
deviations occur that are controlled by the AVO class. N/G = 1. Accordingly, it is occasionally tempting to

05-Vernik_Ch05.indd 110 12-08-2016 20:59:10


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 111
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 5.  Plane-wave synthetic angle gathers gas-sand reservoir, generated using a (a) the 100%-N/G value and (b) the 50%-N/G
value of the gas-sand reservoir. (For modeling parameters, see Figure 16 of Chapter 4.) Note that whereas both the near-angle and
far-angle amplitudes are reduced by more than 50% with a reduction in N/G, the “time thickness” of the lower-N/G reservoir is
actually greater as a result of a reduced interval-velocity.

believe that it is net pay, as a product of the gross thick- appraisal-well-based information on gross reservoir
ness and N/G above the oil-water or gas-water contact, is thickness is also necessary.
largely responsible for variations in seismic attributes
(e.g., Connolly, 2007). However, the qualitative resem-
blance between the seismic signatures of gross thickness Seismic N/G computation from
and N/G changes can be deceptive because significant simultaneous inversion
quantitative differences may exist that are important to
recognize. AI-SI inversion
To demonstrate those differences, the Class III AVO
model that was used in Figures 15 and 16a of Chapter 4 can The 3D seismic data available over the Gulf of Mexico
be invoked. The new model is composed of the same gas-­ oil field, which is the focus of the case history considered in
saturated sand and shale (see Table 5 of Chapter 4) and the this chapter, were prestack time-migrated with a Kirchhoff
same gross thickness of 30 m, but with its N/G ­reduced to algorithm. Additional noise reduction (e.g., multiple remov-
50%. The N/G reduction is simply achieved by introducing al) was performed on the initially flattened gathers, after
four equally spaced 3.75-m-thick intrareservoir shale inter- which residual-moveout corrections were applied by using
vals separating five 3-m-thick gas-sand beds. a nonhyperbolic flattening beyond the angle of incidence of
The results of the forward modeling, which uses a 30°. That moveout correction yielded a high-quality set of
plain-wave wavelet with a 5/10/40/50-Hz bandwidth, are flat gathers that could be stacked in three selected overlap-
presented in Figure 5, which compares 100%-N/G AVO ping angle ranges (near angles of 0 to 22°, medium angles
synthetic angle gathers (Figure 5a) with 50%-N/G gathers of 20 to 35°, and far angles of 30 to 45°) in preparation for
(Figure 5b). The shale-free reservoir-sand synthetic is the simultaneous P- and S-impedance inversion.
­basically the same as the 30-m-thick fragment of the sand The relatively high fold in the overlapping angle
wedge model displayed in Figure 16a of Chapter 4. ranges produced high-S/N partial-angle stacks. The
However, the 50%-N/G model with a 15-m gas pay, near-angle and far-angle stacks in Figures 6 and 7 dis-
shown in Figure 5b, looks similar in its AVO response to play strong Class III AVO behavior in the central and
the 6-m-thick gas-sand fragment of the wedge model (the western areas of the lower reservoir (in the M-sand) but
second gather from the left on Figure 16a of Chapter 4). not in the eastern locations. It is apparent that the nega-
At the same time, the “time thickness” of the 50%-N/G tive AVO gradient becomes increasingly negative in this
model is noticeably greater than that of the 6-m-thick, main reservoir zone — a behavior that is confirmed by
100%-N/G analog, which, of course, is equal to the irre- the synthetic angle gather (see Figure 8) and typically is
ducible time thickness below ­tuning. Obviously the AVO linked to (1) the greater sand volume, (2) slightly greater
attributes cannot be unequivocally related to net pay in stiffness of the medium-grained massive channel sands,
any straightforward manner — i­ndependent geologic or and (3) the absence of high-impedance marls in the

05-Vernik_Ch05.indd 111 12-08-2016 20:59:11


112 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 7.  Horizon amplitude maps for the top of the lower
reservoir (the M-sand). Amplitude is extracted at local
minima on the waveform. Appraisal wells 1 through 7 are
shown at the horizon intersection locations. Note the
Figure 6.  An arbitrary line showing the near-angle- and far- heterogeneous AVO character, which varies from a strong
angle-stack volumes. This section runs northwest to southeast Class III response to an almost flat response (with an AVO
and contains appraisal wells 1 and 3 (see Figure 7) of this gradient close to zero) at well locations 4, 6, and 7. Adapted
deepwater Gulf of Mexico oil field. Note the strong AVO from Figure 9 of Vernik et al. (2002). Used by permission.
response at the top and base of the lower reservoir (the
M-sand). Adapted from Figure 8 of Vernik et al. (2002). Used reservoir sands and encasing shales over the specified time
by permission.
interval of interest. Anisotropy-corrected acoustic-imped-
ance and elastic-impedance logs are used for wavelet
immediate overburden. The last two factors are expected ­extractions from each seismic volume at the control well
to contribute to the shale/oil-sand dipoles’ clockwise ro- locations. To delineate the gross reservoir boundaries and
tation away from the line with a slope of 1/4 passing minimize tuning effects, a low-frequency earth model is
through the shale point and denoting a zero AVO gradi- built from the edited and low-pass-filtered AI and SI
ent in the AI-SI classification scheme (for example, see well logs. The simultaneous inversion process first gener-
Figure 2b of Chapter 4). ates band-limited AI and SI volumes, which ultimately
However, this dominant Class III AVO behavior was are transformed into absolute i­mpedances by merging
found to vary laterally in both the upper and lower reser- them with the log-based low-­frequency model (Figures 9
voirs and to include transitional, zero-gradient responses and 10).
as well as Class IV AVO signatures. Note that the main The seismic-band-limited AI and SI volumes are in-
driver for these variations is the lithology effect, which verted for deterministically by minimizing the cost func-
can be reduced to net-to-gross variation above the OWC tion that involves the P- and S-wave reflectivities, which
and, consequently, to the upscaled Vsand curve at the well are wired into the exact Zoeppritz or Fatti equation, with
locations. the scalars being constrained by the impedance-model
Inputs to simultaneous sparse-spike inversion (Pendrel lines. It is noteworthy that by muting the input gathers at
et al., 2000) typically are seismic angle stacks, the corre- approximately 40° and using the two-term isotropic Fatti
sponding seismic wavelets, the low-frequency model, and, approximation (Fatti et al., 1994), the process can yield
occasionally, several constraining AI-SI relationships for inversion results that are nearly as good as or even better

05-Vernik_Ch05.indd 112 12-08-2016 20:59:12


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 113
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 8.  An NMO-corrected synthetic angle gather for well 1, showing the upper reservoir (the J-sand, centered at 3900 m)
and the lower reservoir (the M-sand, centered at 4150 m). Computed VS is used instead of the measured one, which results in a
better dynamic tie to the seismic gathers around the well. A 5/10/30/40-Hz-bandwidth wavelet, similar to that extracted from
the seismic, is applied. Note that the AVO gradient at the top of the lower reservoir (at 4130 m) is much more negative than the
gradient at the upper one (3890 m). That is consistent with the angle stacks shown in Figure 6.

Figure 9.  An arbitrary line through the absolute AI inversion volume, showing the upper (J-sand) and lower (M-sand)
reservoirs as well as some appraisal-well geometries (wells 3, 1, 4, and 6, left to right). Note the asymmetric shale impedances
above and below the M-sand. The color code is given in g/cm3 × km/s. Adapted from an unpublished figure in a presentation by
Vernik et al. (2001). Used by permission.

05-Vernik_Ch05.indd 113 12-08-2016 20:59:15


114 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 10.  An arbitrary line (the same line as in Figure 9) through the absolute SI inversion volume, showing a highly
heterogeneous SI distribution in both the encasing shales and the M-reservoir sands. The color code is given in g/cm3 × km/s.
Adapted from an unpublished figure in a presentation by Vernik et al. (2001). Used by permission.

than those based on the exact isotropic equation. That is (Figure 11). This can be done using a simple mathemat-
the case because this approximation, under certain condi- ics tool available in most inversion packages. A 2D line
tions, fortuitously nearly accounts for the anisotropy con- through the computed 3D sand volume shown in Figure
trast between the sands and shales (Chapter 4). The qual- 11 immediately confirms a stratigraphic nature for both
ity of the density inversion in both cases can be consid- reservoirs, with a marked worsening of their quality in
ered to be largely inadequate because of the limited angle the updip, northeast direction.
range of high-fidelity traces available and the limitation As Figure 12 illustrates, the AI, SI, and Vsand curves
of the inversion engine itself, which is not designed to ac- extracted from the respective volumes along two apprais-
count for anisotropic effects. al wells are fairly consistent with the upscaled log imped-
The shear-impedance volume in Figure 10 shows ances and Vsand curves (the Vsand curves in Figure 12 are
much greater heterogeneity in both the encasing shales obtained by smoothing the well-log-model-based Vsh
and the M-reservoir sands, compared with that in the curves and subtracting them from unity). Interestingly,
acoustic-impedance volume (Figure 9). The latter, on the the match here for Vsand (and N/G) is visually superior to
other hand, reveals asymmetric impedances above and that for the impedances. Such a superior result ought to be
below the lower reservoir. The observation of asymmetric considered an additional bonus of this N/G estima-
impedances in encasing mudrock formations has a poten- tion ­approach; the systematic errors in inverted imped-
tial to distort amplitude- or even single impedance-based ances are likely minimized in equation 1 due to
reservoir quality assessment. However, the simultaneous subtraction.
P- and S-impedance inversion substantially mitigates this Another quality check is presented in Figure 13,
problem. which is an AI-SI template populated with the inverted
impedance extractions along the same well trajectories.
Sand volume computation and net/gross Notably, the position of the shale-dominated sections
mapping relative to the log-derived shale line with a slope of 0.67
­remains basically unchanged. Somewhat hypothetically,
Combining the AI and SI volumes in equation 1 if — even on the seismic scale — this reservoir con-
yields an inverted-from-seismic Vsand attribute volume tained pure-channel-sand sections of the M-sand, with

05-Vernik_Ch05.indd 114 12-08-2016 20:59:16


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 115
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 11.  An arbitrary line (the same line as in Figures 9 and 10), but through the Vsand volume. Again, wells 3, 1, 4, and 6
(left to right) are shown. Adapted from an unpublished figure in a presentation by Vernik et al. (2001). Used by permission.

Figure 12.  Well-log (black), upscaled-well-log (blue), and simultaneous-inversion-derived (red) AI, SI, and Vsand data, extracted
from wells (a) 1 and (b) 5.

05-Vernik_Ch05.indd 115 12-08-2016 20:59:18


116 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 13. An AI-SI template showing 0-, 50%-, and 100%-


N/G (Vsand) lines and superimposed inverted impedances
extracted along the two appraisal-well trajectories shown in
Figure 12. It is instructive to compare this figure with Figure
2, a similar template showing log impedances. Adapted from
Figure 11 of Vernik et al. (2002). Used by permission.

Figure 14.  (a) A thickness-average N/G map and (b) a net-


its typical gross thickness of approximately 45 m, those oil-sand-feet map for the M-sand reservoir. Adapted from
sections would have plotted in the proximity of the oil- Figure 14 of Vernik et al. (2002). Used by permission.
sand trend with Vsand = 1, provided the fluid properties
did not change appreciably. In reality, as follows from
Figure 13, the seismic-scale impedance volumes barely responses that are related to lithologic factors such as mu-
contain values reaching the oil-sand line in the AI-SI drock composition, sand lithofacies heterogeneity, and
template. That is the case primarily because the individ- various levels of consolidation. However, computation of
ual channel-sand beds and bed sets typically are far the hydrocarbon-saturated sandstone volume is expected
below the tuning thickness, which is approximately 35 to to become more challenging or even impossible in well-
40 m in this case study. Moreover, thin channel sands are consolidated, lower- p­ orosity clastic sequences. As will
often separated by intrareservoir shales or levee deposits, be demonstrated below, lithology effects overwhelm
both of which reduce the N/G value of the reservoir. those related to different fluids in most high impedance
The noted heterogeneity of the nonreservoir rocks environments.
(shales and marls) and the reservoir sands in addition to In the final steps of the QI workflow for the lower
the net-to-gross variations give rise to a variety of AVO reservoir, the Vsand volume in the time domain is inter-
behaviors that are observed both laterally and vertically in sected by the top and base time surfaces of this reservoir.
the field. Indeed, as we recall, Class III, Class IV, and This is done to extract the thickness-average value of the
zero-AVO responses are observed throughout the cube of N/G attribute by using equation 3, with the top and base
CDP gathers for both the upper and lower reservoirs. reservoir depths replaced by their respective two-way
Nonetheless, this apparent complexity turns out to be in- times. Figure 14a is a map showing this reservoir-scale
consequential in computing the oil-saturated Vsand vol- N/G distribution. The gross reservoir isochore can be con-
ume, primarily because the AI and SI variations of the two verted easily to an isopach by using a vertical P-wave ve-
end-member components occur along the quasi-linear oil- locity volume calibrated by the anisotropy-corrected
saturated arenite and shale/marl lines and are automati- well-log velocities. The gross isochore above the oil-wa-
cally accounted for by equation 1 in the virtual axis-rota- ter contact is multiplied by the average N/G value to com-
tion process. pute the net reservoir-sand thickness displayed in Figure
Importantly, that observation, which is quite evident 14b. These maps are vastly superior to those generated
in Figure 2, makes the Vsand estimation using the proposed from a single AI inversion, let alone to those from the full-
technique relatively i­mmune from the diverse AVO stack seismic amplitudes.

05-Vernik_Ch05.indd 116 12-08-2016 20:59:20


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 117

The full-offset AVO response that is taken into


a­ ccount in the simultaneous inversion undoubtedly facili-
tates identification of the channel-sand-dominated loca-
tions with high N/G values and levee-dominated ones
with reduced N/G values. The primary reason underlying
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

such a capability is the shale-content differences that give


rise to different VP/VS values above the oil-water contact
and therefore to different respective RP and RS contrasts at
the top of the reservoir. Although, as was mentioned
above, the seismic N/G inversion is relatively immune to
the plausible AVO class variations in any specific field,
spatial correlation with such AVO variations cannot be
ruled out entirely. In fact, in our example the low AVO
gradient transitioning from Class III to Class IV behavior
in the northeastern part of the field (wells 4 and 6) is ob-
served to correlate with the low N/G and net-pay esti-
mates in Figure 14.
Such a correlation between AVO responses and lat-
eral lithologic variations is associated with (1) higher- Figure 15.  Average well values of N/G from the log model
impedance marls overlying the reservoir at northeastern versus those predicted from the simultaneous inversion of
locations and (2) the presence of lower-rigidity sand seismic data. Wells under the salt canopy in the northwest
and silt components that are typical of levee deposits, as corner of the field are omitted. Adapted from Figure 15 of
opposed to channel sands. These two peculiarities in Vernik et al. (2002). Used by permission.
seismic rock properties contribute to the counterclock-
wise rotation of the nonreservoir/reservoir dipoles from caution should be exercised when exotic shale lithologies
the Class III to the Class IV AVO sector in the AI-SI (e.g., organic-rich mudstones) are present in the overbur-
classification scheme (Chapter 4). As such, those local den or in the proximity of the reservoir base. These or-
lithology effects by themselves do not have any cause- ganic shales do not plot on or even near the conventional
and-effect influence on seismic N/G attribute shale-compaction trend line prescribed by the rock-­
computation. physics model discussed above (and in Chapter 3) and
An N/G map can be used extensively in reservoir may negatively impact the accuracy of the N/G
model construction, oil-in-place estimation, and, finally, interpretation.
in selecting the locations for development and production Another limitation is that this method works adequate-
wells. Figure 15, which plots the inversion-based N/G ly only above the hydrocarbon-water contact. In the water
values above the OWC against the log-model N/G values leg below the reservoir, the relative positions of the up-
for the appraisal and development wells, is an ultimate scaled reservoir impedances in the AI-SI template change
confirmation of the proposed technique, at least in fields according to the water-saturated Sand Diagenesis Model.
in which the gross reservoir thickness does not dramati- Moreover, the line described by this wet sand(stone) model
cally deviate from the tuning thickness and that are domi- is no longer quasi-linear in the dynamic range of interest
nated by Class III or IV AVO responses. Note that the si- nor subparallel to the encasing nonreservoir lithologies un-
multaneous inversion was carried out before the develop- less a significant concentration of the residual hydrocarbon
ment wells were drilled, so that the latter can be viewed as remains in the aquifer.
“blind” wells. Finally, the technique for N/G inversion introduced
The overall accuracy of the prediction is within 8%, by Vernik et al. (2002) may not be fitting for reservoirs
which incorporates errors in estimates of the gross thick- with a gross thickness significantly below or above the
ness. The RMA regression slopes for the two reservoirs tuning thickness. There are two different reasons for those
are close to 45°. Not surprisingly, the predictive power is two cases: (1) inverted impedances for reservoirs below
much less impressive under the salt-canopy-affected areas tuning thickness are mostly not accurate, and (2) inverted
of the 3D seismic data (in the northwest corner of the impedances for reservoirs significantly above tuning
field, well 7), where the S/N value is low and the simulta- thickness typically suffer from uncertainties with low fre-
neous inversion results are largely inaccurate. quency (or initial) model. In such situations, as well as in
The technique presented here also has several other frontier exploration in general, the AVO attributes, such
pitfalls and limitations, as is the case in any aspect of seis- as intercept, gradient, and, especially, the fluid factor (see
mic exploration and reservoir characterization. First, Chapter 4), are deemed more appropriate.

05-Vernik_Ch05.indd 117 12-08-2016 20:59:21


118 Seismic Petrophysics in Quantitative Interpretation

Effects of lithology and fluid on the shale and hydrocarbon-­ saturated sand diagenesis
models are subparallel to each other, which facilitates
prestack attributes N/G computation from AI and SI volumes. The approxi-
The seismic N/G attribute described above is a combi- mation of the model lines corresponding to the end-
member lithologies being subparallel in AI-SI is ­violated
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

nation attribute and, as such, is not dissimilar to the fluid


factor, which was alluded to in Chapter 4. In both ap- only for totally wet sands and sandstones.
proaches, reservoir-quality estimation and mapping take Additionally, it is important to note that the lithology
advantage of the anomalous deviation of the extracted vectors, which are visualized in AI-SI space as dipoles
AI-SI coordinates from the background trend dominated by connecting the end-member data points of wet sands and
shale-on-shale reflections in most siliciclastic sequences. shales, always cut across the diagenetic trends at an angle.
Consequently, both techniques always present a combina- Lithology vectors are exemplified schematically on the
tion of lithology and fluid effects that may be impossible to background of the AI-SI template in Figure 16, where
separate without additional, independent information. they are denoted by solid black arrows.
Nonetheless, the differences b­ etween the fluid factor The fluid vectors are added in Figure 16 as dotted
and seismic N/G values are significant in that the latter black arrows, even though they are not strictly linear, as
are derived from seismic inversion instead of from P- follows from the Gassmann-equation modeling of sands
and S-wave reflectivities, which are more dramatically and sandstones that have equal porosity, consolidation,
affected by the wavelet effects. On the other hand, the and effective stress acting upon them (see Figure 39 of
quasi-linear and subparallel diagenesis trend lines that Chapter 3). The fluid vectors are always subhorizontal
can be identified in AI-SI space over a limited range of with minor positive slope and connect water-, oil-, and
well-log or inversion-derived impedances are not among gas-saturated sand(stone) points in AI-SI space.
the necessary conditions for fluid-factor computation, as Lithology vectors may occasionally coincide in ori-
they are for accurate N/G inversion. To reiterate, in this entation with fluid vectors (case 3 in Figure 16). The
limited impedance range, which is commonly of interest M-sand reservoir in the central and western areas of the
in quantitative interpretation of oil and gas reservoirs, oil field described above and characterized by strong

Figure 16. An AI-SI template, schematically showing five examples of lithology vectors (solid black arrows) and fluid vectors
(dotted black arrows) and their vectorial summations (red arrows) for gas-saturated arenites. Note that the lithology vector (1) is
reduced to zero in case 1, (2) is dominated by the fluid vector in cases 2 through 4, and (3) dominates the fluid vector in case 5.
Also note the directional coincidence of the lithology and fluid vectors in case 3.

05-Vernik_Ch05.indd 118 12-08-2016 20:59:22


Chapter 5: Simultaneous AI-SI Inversion and N/G Computation 119

Class III AVO behavior is a good example of such a near From this point on, in the direction of increasing dia-
coincidence. In other seismic rock-property environ- genesis (shale compaction/recrystallization and sand(stone)
ments, lithology vectors may be oblique (case 2 in Figure consolidation/cementation), the lithology vectors tend to
16) or even be quasi-orthogonal to fluid vectors (case 4 in grow in size due to the divergence of wet sand(stone) and
Figure 16). The latter situation is common in Class II shale AI-SI trends. In the same direction, the fluid vectors
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

AVO settings, where, unless in high impedance (low po- tend to gradually shorten, and these tendencies culminate
rosity) environments, an attempt to separate fluid from at the point of convergence of the wet-, oil-, and gas-sand-
lithology (net-to-gross) effects can be more successful. stone lines, where both total porosity and, therefore, the
Although it is instructive to identify the relative ori- fluid vector are reduced to zero. Note that this point is out-
entations of the lithology vectors and the fluid vectors in side of the scope of Figure 16, but it is approached in very
our QI workflow, it should be emphasized that those rela- low porosity tight sandstones.
tive orientations have no impact on the seismic N/G val- Because siliciclastic sequences undergoing advanced
ues or the fluid factor because these attributes are always stages of diagenesis, including quartz cementation in ar-
related to the sum of the fluid and lithology vectors in enites, naturally produce AVO behavior of Classes II and
AI-SI space. The resulting interface vectors (dipoles) are I, these responses, even when characterized by the rela-
denoted in Figure 16 by red arrows. tively strong negative AVO gradients, may never be con-
A curious special case is that of a seismically nonre- sidered reliable DHIs, let alone net-pay indicators. The
flective (transparent) wet-sand/shale interface with simi- level of sandstone diagenesis at which no accurate predic-
lar impedances on both sides. Obviously, this special case tion from seismic data is possible for the hydrocarbon-
corresponds to the point of intersection of the wet-sand saturated N/G values and, therefore, for net pay, is subjec-
and shale lines — the point at which the lithology vector tive. This level should be determined by the asset-team
is reduced to zero by definition (case 1 in Figure 16). geoscientists on a case-by-case basis using detailed rock
If this situation is established, any nonzero computed physics (e.g., the Sandstone Diagenesis Model in combi-
seismic N/G values can be confidently and entirely attrib- nation with fluid substitution and the Shale Model) and
uted to the effects of hydrocarbon fluids, unfortunately synthetic forward modeling of the offset-dependent
including those in low-saturation reservoirs. reflectivity.

05-Vernik_Ch05.indd 119 12-08-2016 20:59:22


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 6: Seismic Petrophysics of Unconventional Reservoirs

Just ten years ago, the use of seismic rock properties needs to be converted from weight to volume fractions, as
for seismic detection and characterization of source rocks was discussed in Chapter 1, before it can be used in rock-
was of very limited interest. However, after a decade of physics modeling.
tremendous advances in horizontal drilling and hydraulic Significant information on the organic phase of un-
fracturing, source rocks now draw much more atten- conventional reservoir rocks, including total-organic-car-
tion — for their ability to generate hydrocarbons and for bon (TOC) content and the maturity index, comes from
their role as reservoir lithologies. pyrolysis setups such as RockEval (Durand and Espitalie,
The microscale and nanoscale natures of these rocks 1976; Price, 1983; Durand et al., 1987), whereas vitrinite
have made their description and seismic characterization reflectance is used as an additional, petrographic source
much more challenging than we are accustomed to con- of data on organic-matter maturity. Microstructural and
fronting in siliciclastics and carbonates. As a conse- textural information comes from scanning electron mi-
quence, in this chapter more attention is given to the croscopy (SEM), often performed on focused-ion-beam-
­advanced petrographic imaging and organic geochemistry prepared surfaces (FIB-SEM, e.g., Loucks and Ruppel,
techniques that are useful in helping us understand elastic 2007) as well as from nanoCT scanners (e.g., Walls and
properties and their controls in unconventional shale Sinclair, 2011).
reservoirs.
Starting with analysis of composition, texture, and
thermal maturity, this chapter focuses on the petrophysi- Rock composition
cal log model and the rock-physics model of those uncon-
ventional shale-oil and gas reservoirs. Much attention is An ultra-fine-grained nature and dispersed-organic-
paid to the laboratory data on the P- and S-wave velocities matter content impart dark-gray to black colors to organic
and elastic anisotropy of these rocks in relation to their mudrocks and make these unconventional reservoir lithol-
kerogen content, mineral composition, and level of ther- ogies poor candidates for standard quantitative petrogra-
mal maturity. Thermal maturity, of course, is widely ac- phy — that is, for thin-section point counting of their
cepted as being responsible for hydrocarbon generation, porosity and mineral content. Multiple XRD analyses
­
open microcrack development, and the creation of excess show that the primary minerals comprising unconvention-
pore pressure in organic mudrocks. The velocities and al shales are quartz, silica (chert, opals), carbonates, and
seismic anisotropy are critical for mapping organic rich- clays. A schematic ternary diagram describing composi-
ness and reservoir properties from seismic amplitudes tion by volume could look like the one presented in
and/or impedances. They also aid us in 3D earth model Figure 1. The terms “shale” and “mudstone” have tradi-
building and, consequently, more accurate location of mi- tionally been employed in the general sense to refer to all
croseismic events during hydraulic-fracture stimulation. these ultra-fine-grained rocks. Alternatively, and possibly
more specifically, the term shale can be retained to de-
scribe a rock with a strongly oriented fabric made up of a
Petrologic data in unconventional shales mixture of all the major minerals mentioned above and
occupying the pentagon in the middle of the ternary dia-
Petrologic data acquisition and analysis are even gram, with a carbonate content of less than 50%, a silica
more critical to our understanding and interpretation of content of less than 25 to 30%, and a clay mineral content
the seismic attributes associated with unconventional res- of less than 50% on an organic-free basis (Figure 1).
ervoirs than of those associated with sandstones and car- Applying the schematic diagram presented in Figure 1,
bonates. The primary source of information about mineral we can adequately classify the lithologies occurring in the
composition is X-ray diffraction analysis (XRD), which most prolific organic-rich mudrock formations of the world,
121

06-Vernik_Ch06.indd 121 12-08-2016 19:59:50


122 Seismic Petrophysics in Quantitative Interpretation

Quartz, feldspars, silica where the subscript i denotes a specific mineral, wi is the
weight fraction of the mineral reported in XRD analysis,
and ρi is the mineral density. The term ρnk, which is the
Siliceous solid-grain (matrix) density on a kerogen-free basis, can
be replaced with ρnk = (ρm − ρkK)/(1 − K), where ρm is the
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

mudstones
matrix density of the rock and K is the kerogen volume
fraction of the solid phase of the rock.
50% 50%
From equation 1, the formula for the clay fraction is
Shales νcl = wcl(ρnk/ρcl), where wcl is the total clay determined
from XRD analysis. On the other hand, the kerogen vol-
ume fraction can be computed as K = (TOC /Ck )(ρm /ρk ),
Marls
Clay-rich where TOC is the total organic carbon reported from rock
shales pyrolysis in weight percent, Ck is the TOC-to-kerogen
Limestones
conversion factor (in the 65 to 90% range depending on
Carbonate Clay thermal maturity), and ρk is the kerogen density.
50%
Secondary minerals found in organic shales are feld-
Figure 1.  An organic-mudrock classification scheme. Note spars, pyrite, and hydroxyapatite. Pyrite is the most influ-
that kerogen content is excluded from this mineralogy-based ential among them because of its physical properties, with
approach. Also note that clay-rich organic shales with more ρm = 5.01 g/cm3. This mineral, even at relatively low con-
than 60% clay volume are quite rare. centrations averaging approximately 3% by volume,
causes the solid grain density to increase quite substan-
such as the Bakken Formation, Eagle Ford Group, Woodford tially until it often exceeds the average value of 2.73 g/cm3
Shale, Marcellus Formation, Barnett Shale, Monterey reported for conventional mudstones and shales in Table 2
Formation, and Niobrara Formation (all USA), the Bazhenov of Chapter 1.
Formation (Russia), the Kimmeridge Shale (North Sea and
UK), and the La Luna Formation (Venezuela). For instance,
Organic richness and thermal maturity
the majority of Bakken and Woodford mudrocks fall in the
middle portion of the diagram (shales), the Eagle Ford and The TOC concentration and maturity level of kerogen
La Luna formations lean toward the carbonate corner (marls are the key parameters in any unconventional-reservoir
and limestones), and some upper Woodford, Barnett, and characterization. Ultimately, they relate to the ability of
Monterey rocks tend to gravitate upward in the ternary dia- the organic shale to contain and potentially produce hy-
gram (to siliceous mudstones). drocarbons (Tissot and Welte, 1978; Durand and Espitalie,
The total clay content on an organic-free basis among 1976; Price, 1983). The organic geochemistry informa-
the formations mentioned varies from 5 to 50% and rare- tion acquired for the database of laboratory measurements
ly exceeds that concentration. For example, some of P- and S-wave velocities in organic mudrocks present-
Kimmeridge and Marcellus shales are more than 50% ed later in this chapter are given in Table 1, which is large-
clay minerals, but even these rocks adhere to the general ly based on data from Vernik and Landis (1996).
rule regarding an average vcl value of less than 50%. In the Specifically, Table 1 reports on the solid-grain densities,
generic sense, however, we can refer to all these various TOC contents, and maturity indicators for a subset of
lithologies as organic shales and/or organic mudrocks if samples that are included in the database.
total organic carbon (TOC) in them exceeds 1.5%. On the basis of RockEval pyrolysis information, the
The solid organic matter in organic mudrocks typi- hydrocarbon-generating potential of an organic mudrock
cally is referred to as kerogen. It is amorphous, and as sample is expressed by the hydrogen index HI, which is the
such it cannot be accounted for by XRD analyses. ratio of the amount of convertible hydrocarbon (S2) to the
Therefore, the fractional content of kerogen is obtained TOC and is expressed in milligrams per gram of organic
from rock pyrolysis and reported as TOC. Because both carbon (mg/gOC) (Tissot and Welte, 1978). The TOC val-
XRD mineral fractions and pyrolysis data are tabulated in ues in the database range from 1 to 20%, but the dominant
weight fractions or percents, they should be converted to range is smaller: 2 to 12%. The vast majority of the core
volume concentrations for s­eismic-rock-property analy- samples in Table 1 contain oil-prone (type II) kerogen.
sis. When one is operating on a kerogen-free basis, the An attempt to classify the thermal-maturation stages
conversion formula used by Vernik and Landis (1996) is of organic shales is given in Table 2 (using data from
Vernik and Landis, 1996), which lists the HI from
ρnk
vi = wi , (1) RockEval analysis and also the vitrinite reflectance Ro.
ρi According to Price et al. (1984), the HI provides a

06-Vernik_Ch06.indd 122 12-08-2016 19:59:51


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 123

Table 1.  Lithology, solid-grain density, TOC, and maturity data from the unconventional-
shale database.
Depth (m) Lithology ρm (g/cm3) TOC (%) HI (mg/g OC) Ro (%)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Bakken Formation, North Dakota, USA


2307 Shale 2.18 16.2 569 –
2632 Shale 2.15 15.9 493 –
2996 Shale 2.31 10.3 420 –
3098 Shale 2.32 12.1 436 0.61
3196 Shale 2.28 15.4 416 –
3199 Shale 2.41 9.8 425 –
3223 Shale 2.61 7.0 447 0.69
3271 Shale 2.44 9.7 282 0.80
3272 Shale 2.45 7.1 292 0.82
3332 Shale 2.46 9.5 135 1.13
3423 Shale 2.67 4.5 115 1.10
3428 Shale 2.49 9.8 161 1.19
3438 Shale 2.59 6.0 97 1.27

Eagle Ford Shale, Texas, USA


3711 Marl 2.62 2.4 30–50 1.2–1.4
3713 Limestone 2.70 0.7 30–50 1.2–1.4
3713 Marl 2.58 3.5 30–50 1.2–1.4
3716 Marl 2.61 3.0 30–50 1.2–1.4
3721 Marl 2.54 5.1 30–50 1.2–1.4
3723 Limestone 2.72 2.2 30–50 1.2–1.4
3729 Limestone 2.67 1.2 30–50 1.2–1.4
3736 Limestone 2.65 2.7 30–50 1.2–1.4
3742 Limestone 2.60 6.8 30–50 1.2–1.4
3752 Limestone 2.69 0.8 30–50 1.2–1.4
3753 Limestone 2.64 5.9 30–50 1.2–1.4
3754 Marl 2.57 2.4 30–50 1.2–1.4
3754 Limestone 2.59 5.8 30–50 1.2–1.4
3754 Marl 2.57 4.2 30–50 1.2–1.4
3755 Marl 2.55 5.9 30–50 1.2–1.4
3757 Marl 2.59 5.5 30–50 1.2–1.4

Bazhenov Formation, Western Siberia, Russia


3784 Shale 2.60 6.2 385 0.61
3787 Shale 2.47 7.4 434 0.64
3788 Shale 2.49 6.0 319 0.64
3822 Shale 2.74 2.8 279 0.78
3824 Shale 2.59 4.7 422 –
3834 Shale 2.65 2.8 288 –
3842 Shale 2.62 3.6 276 0.61
(Continued)

06-Vernik_Ch06.indd 123 12-08-2016 19:59:51


124 Seismic Petrophysics in Quantitative Interpretation

Table 1.  Continued.


Depth (m) Lithology ρm (g/cm3) TOC (%) HI (mg/g OC) Ro (%)

Woodford shale, Oklahoma, USA


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

1960 Siliceous 2.63 2.3 540 0.49


shale
2154 Shale 2.64 6.6 582 0.49

Niobrara Formation, Colorado, USA


1121 Marl 2.79 1.2 107 0.88
1159 Marl 2.74 1.9 120 1.15
2248 Marl 2.67 2.1 19 1.15
2323 Marl 2.77 1.6 7 1.31
2350 Marl 2.72 0.5 15 1.36

Kimmeridge Shale, North Sea


2133 Shale 2.36 8.6 499 0.49
4412 Shale 2.53 4.4 103 1.10
4449 Shale 2.63 5.0 67 1.25

Table 2.  Stages of organic shale maturation and hydrocarbon generation. Level of organic metamorphism
(LOM) refers to oil-prone types of organic matter. Data are from Vernik and Landis (1996).
Stage Description Ro (%) HI (mg/gOC) LOM
I Compaction; early methane <0.3 >700 <4.8
II Onset of oil generation, H2O expulsion 0.3–0.5 500–700 4.8–7.2
III Advanced oil generation 0.5–0.75 300–500 7.2–9.5
IVa Main stage of oil generation 0.75–0.9 200–300 9.5–10.4
IVb Final stage of oil generation 0.9–1.3 50–200 10.4–11.6
V Condensate and wet gas 1.3–2.0 5–50 11.6–13.5
VI Dry gas >2.0 <5 >13.5

quantitative measure of the level of maturity of kerogen well-log observations on sonic velocities in unconventional
and is complementary to Ro. shales, as well as to the levels of organic metamorphism
The stages of thermal maturity gradually succeed (LOM) introduced by Hood et al. (1975). The organic ma-
each other in a mudrock’s burial history and result in the turity expressed in the LOM scale is a principal part of the
onset and progression of kerogen transformation into liq- log-based TOC evaluation method of choice in this book
uid hydrocarbon. The maturation stages presented in (Passey et al., 1990). The gamma-ray and density-log-based
Table 2 are conceptually similar to those proposed for transforms are much less accurate TOC indicators in gen-
type II kerogen (Tissot and Welte, 1978), with two addi- eral but can be useful for specific geographic locations.
tions that are appropriate for rock-physics characteriza- Figure 2 is a crossplot of TOC versus HI values for a
tion: (1) An Ro value of 0.75% is used as the onset of stage significant subset of the database. This plot illustrates the
IV, the main stage of petroleum generation and primary wide range of organic richness, oil-generating potential,
migration; (2) an Ro value of 0.90% is used as an internal and thermal maturity sampled in the database. The matu-
boundary that divides stage IV into two substages: IVa rity stages of Table 2 are implicitly mapped onto the TOC-
and IVb. versus-HI plot using the isomaturity lines described by the
As will be discussed later in this chapter, the subdivi- ad hoc exponential saturation function HI = HImax
sion of stage IV into substages allows us to better relate [1 − exp(−0.5TOC)], where HImax values were selected at
thermal-maturity information to the laboratory and 500, 370, and 50 mg/gOC for Ro values of 0.5, 0.75, and

06-Vernik_Ch06.indd 124 12-08-2016 19:59:51


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 125
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 2.  TOC content versus the RockEval hydrogen index Figure 3.  Plot of the hydrogen index (HI) versus vitrinite
(HI) for Bakken, Woodford, Niobrara, Bazhenov, and reflectance. The two measures of organic maturity are well
Kimmeridge shale mudrocks, showing isomaturity contours correlated and can be used interchangeably. Note that
that map the maturation stages with respect to oil. Adapted from equation 1 is an inverse of the exponential regression line on
Figure 3 of Vernik and Landis (1996). Reprinted with permission this plot. Adapted from Figure 4 of Vernik and Landis (1996).
from AAPG, whose permission is required for further use. Reprinted with permission from AAPG, whose permission is
required for further use.
1.3%, respectively. The function is based on the empirical
relationship between the two maturity indicators, HI and wet-gas window, which is overmature with respect to oil
Ro (Figure 3) — a relationship that reflects the tendency (stage V). It should be noted that vitrinite reflectance is
for the organic matter to deplete with maturity. The basis widely considered a more reliable maturity indicator, and,
for the HI and Ro coupling is discussed next. as a consequence, it is given more weight in the matura-
Those two parameters — the hydrogen index from tion-stage mapping of Figure 2. Notably, the isomaturity
pyrolysis and the vitrinite reflectance — are well-accept- lines are defined by vitrinite reflectance and dramatically
ed thermal-maturity indicators (e.g., Tissot and Welte, converge to zero with reduction in organic richness.
1978). However, vitrinite reflectance is more difficult to For rock-physics modeling, it is important to distin-
obtain, for a variety of reasons that include the availabil- guish the volumetric fraction of the solid organic m
­ atter —
ity of well-developed vitrinite grains within kerogen sites. the kerogen or K value — from the rock-forming minerals
Relating these two measures of maturity to each other and pores. As Vernik and Milovac (2011) and Alfred and
often helps us to more consistently predict the maximum Vernik (2012) reported, the kerogen concentration on the
burial temperature that a rock has been subjected to dur- solid-rock basis is computed as
ing its history of deposition, burial, and possible uplift.
TOC ρm TOC ρnk
The plot of HI versus Ro in Figure 3 displays a strong, K = = . (3)
Ck ρk TOC ( ρnk − ρk ) + Ck ρk
negative, exponential dependence between the two pa-
rameters, which fits the samples with type II kerogen rea-
sonably well despite the large scatter. The empirical In this formula, ρm, ρnk, and ρk are the solid-grain density,
­relationship that relates vitrinite reflectance to HI is an the nonkerogen-component density, and the kerogen den-
­inverse of that shown in Figure 3: sity, respectively. The term Ck is the carbon concentration
(in percent) in the solid organic matter; as is the case with
7.82 − ln( HI ) kerogen density, the Ck value is a function of kerogen ma-
Ro = , (2)
3 turity, which will be discussed later. The expression on
the right-hand side of equation 3 is obtained by substitut-
where Ro is in percent and HI is in mg/gOC. Figures 2 and ing the solid-grain density, in the original formulation of
3 indicate that the majority of the organic shales in the the equation, with the summation ρm = (1 − K)ρnk + Kρk.
database fall into maturity stages III, IV, and V — that is, The nonkerogen-solid densities can be quite accu-
essentially in the oil window (stages III and IV) and the rately derived from XRD analyses in combination with

06-Vernik_Ch06.indd 125 12-08-2016 19:59:52


126 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 4.  Estimation of kerogen density from vitrinite Figure 5.  The empirical relationship between measured TOC
reflectance. Adapted from Figure 5 of Alfred and Vernik content and the kerogen volume fraction K (computed from
(2012). Used by permission. equation 3 with inputs from XRD mineralogy and vitrinite
reflectance). R here is the coefficient of correlation.

the mineral densities reported in Table 2 of Chapter 1.


Kerogen density is hard to measure, but fortunately it can subject of the next section, is indispensable to a complete
be predicted on the basis of its maturity (Figure 4) by understanding of tensor properties, such as elastic moduli
using data from the literature (e.g., Okiongbo et al., 2005) and the sonic/seismic velocities derived from them.
supported by unpublished measurements (P. R. Craddock,
personal communication, 2015), as compiled by Alfred Rock texture
and Vernik (2012). The relationship between kerogen
density and Ro may not be as monotonic as Figure 4 sug- In general, most organic-rich mudrocks (shales, marls,
gests, especially at the kerogen/graphite transition, but it and siliceous mudstones) were deposited in low-energy,
can be fine-tuned just with additional direct and well-con- low-oxygen environments and were subjected to a minor
strained indirect estimates of ρk (Rudnicki, 2015). degree of bioturbation. Therefore, the majority of such
The kerogen volume fraction for each sample in the mudrocks are characterized by a thinly laminated texture
database was computed from equation 3 and plotted that may or may not persist on the submicroscale, depend-
against TOC in Figure 5. The plot shows that for this quite ing on the mineral and organic-matter (kerogen) contents.
diverse database, the relationship is strong (the coefficient Individual laminae can be a few millimeters to a few cen-
of correlation is R = 0.99) and can be substituted by a timeters thick and may comprise beds that are 0.5 to 2 m
simple linear regression: K = 2.6TOC/100. This trans- thick. Thin-section observations typically reveal interlaced
form is deemed sufficient if the input parameters for strips of organic matter and elements of directional texture
equation 3 are not available. caused by a bedding-parallel arrangement of grains (e.g.,
The information on K in our database will be used the Bazhenov shale, West Siberia; see Figure 6).
extensively in both the petrophysical modeling designed However, key textural details in mudrocks can only
for computing log-based reservoir parameters and in be discerned with scanning electron microscopy (SEM),
rock-physics modeling, especially for derivations of the which images elemental contrasts between submicrosco-
solid matrix density and elastic moduli. On the other pic constituents by using ion-beam-polished rock surfac-
hand, the TOC content will be employed sporadically — es (Figures 7 through 10). The images help resolve low-
primarily in certain useful models and crossplots and/or density organic matter (black or dark gray) from illite
for illustrative purposes in Chapter 8. clay, calcite, and silt grains. They also confirm the pres-
Whereas the compositional data are important in fur- ence of interlaced strips of kerogen (in a latticework), on
ther quantitative seismic analysis, they can be considered a submicroscopic level, that apparently insulate other par-
adequate in log interpretation just of scalar rock proper- ticle assemblages composed primarily of illite clay, silt
ties, such as density. Textural information, which is a grains, and carbonates in tight organic mudrocks.

06-Vernik_Ch06.indd 126 12-08-2016 19:59:54


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 127
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 6.  Photomicrograph of a typical Bazhenov shale core


(from West Siberia, Russia, with TOC = 6% and a vertical
scope of 400 µm), showing interlaced strips of black organic
matter laminations. Courtesy of the Stanford Rock Physics
Laboratory. Used by permission.

It is important to emphasize that the textural charac-


teristics of organic-rich mudrocks, such as the relative
positions and preferred orientations of different particles
and their aggregations, were formed in response to com-
paction and clay mineral transformation prior to the onset
of hydrocarbon generation (Kaarsberg, 1959; Vernik and
Nur, 1992a; Hornby et al., 1994; Wenk et al., 2007).
Porosity reduction and compaction/recrystallization
processes are exemplified in Figure 7, which shows a
poorly compacted Eagle Ford marl (Figure 7a) and its
highly compacted compositional analog from the wet-gas
window (Figure 7b). The role of mineral transformation
reactions (e.g., Hower et al., 1976) is deemed secondary Figure 7.  Field-emission scanning-electron-microscopy
at the early stages of diagenesis, but it can be increasingly (FE-SEM) images of (a) a complex pore system of an Eagle
relevant at advanced stages of organic metamorphism. Ford marl that was not subjected to any significant
Note that the bedding-parallel fabric alignment is absent compaction, and (b) a highly compacted Eagle Ford marl
in high-porosity organic marl due to its very limited com- with a weakly oriented fabric caused by subordinate amounts
paction (Figure 7a) and is quite vague in the well consoli- of clay and kerogen (TOC = 3.1%). Courtesy of K. Milliken,
dated organic marl displayed in Figure 7b — here a result Bureau of Economic Geology, University of Texas, Austin,
of subordinate clay and kerogen concentration in this and T. Kosanke. Used by permission.
sample.
On the contrary, Figure 8 shows SEM images of the from the Eagle Ford Formation. Calcite content in this
Woodford Shale, Oklahoma, and the Marcellus Shale, typical Eagle Ford marl is approximately 55% by vol-
Pennsylvania (both USA), each with a relatively elevated ume because of the abundance of coccolith shells,
clay/mica content. Both images focus on clay-rich micro- whereas the illite content is approximately 17%, with
layers and show a strongly preferred orientation of illite- only 10% organic matter. Nonetheless, the laminated
clay platelets, which, in combination with organic matter, texture of the marl is very well-developed and is attrib-
effectively separate silt grains and grain assemblages. Of uted to the lenticular distribution of both the clay and
note in Figure 8a is the very fine latticework created by kerogen. The high resolution achieved with Ar-ion-
interspersed kerogen and illite flakes permeating the rock. based surface preparation and the SEM imaging tech-
A new generation of backscatter SEM images using nique reveals oriented fabric-controlled microcracks,
Ar-ion cross-section polishing and related techniques is which have a tendency to occur adjacent to or inside of
exemplified by Figure 9. The core sample here is taken illite/mica platelets.

06-Vernik_Ch06.indd 127 12-08-2016 19:59:54


128 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 9.  High-resolution FE-SEM image of an Eagle Ford


marl, prepared by Ar-ion cross-section polishing and showing
a lenticular latticework of kerogen (dark gray), calcite/clay
matrix, and minor quartz silt. Of note are the fabric-
controlled microcracks (black). Courtesy of K. Milliken,
Bureau of Economic Geology, University of Texas, Austin.
Used by permission.

Figure 8.  FE-SEM images of compacted organic shales with


elevated clay contents. (a) Woodford Shale (Oklahoma, USA,
TOC = 5.7%), showing kerogen latticework interspersed
with illite-clay-dominated texture with a strong preferred
orientation of both kerogen and clay platelets. Courtesy of
T. Kosanke. (b) Marcellus Shale (Pennsylvania, USA), with
strongly oriented illite/mica platelets (gray) and kerogen sites
(black) separating quartz silt grains. Courtesy of K. Milliken,
Bureau of Economic Geology, University of Texas, Austin.
Used by permission.

Figure 10.  An FE-SEM image confirming the absence of


The Eagle Ford Group is largely composed of marls any preferred-orientation texture in this Eagle Ford limestone
with a clay content that varies laterally from 12 to 30%, (vcarb > 0.85). Courtesy of T. Kosanke. Used by permission.
and that compositional heterogeneity definitely compli-
cates the formation’s seismic petrophysics characteriza-
tion. In a vertical cross section, the marls are interbedded Similarly quasi-isotropic submicroscopic textures are
with subordinate limestones that rarely reach 1 to 2 m in also typical of siliceous mudstones and cherts of the upper
thickness and often remain below log resolution. On aver- portion of the Woodford Formation, as well as of biogenic
age, the limestones are characterized by lower TOC val- opal-dominated diatomites/cherts of the Monterey
ues and a quasi-isotropic texture, without any dominant Formation, California, USA (Vernik and Liu, 1997). A
preferred orientation of particles (Figure 10). common compositional feature of those unconventional

06-Vernik_Ch06.indd 128 12-08-2016 19:59:55


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 129

mudrock types is insignificant clay content, albeit they the rock-physics measurements of velocity and anisot-
are still quite laminated on microscopic and larger scales. ropy in organic mudrocks.
The widespread application of SEM imaging in mu-
drock studies has revealed a very broad compositional
spectrum, ranging from rocks dominated by siliciclastic Log model for unconventional shales
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

material of extrabasinal derivation (particles of clay min-


erals, quartz, and feldspar) to those with significant ad- The formation-evaluation aspect of unconventional
mixtures of intrabasinal grains composed of carbonate or shale reservoirs is still evolving, although some consensus
biogenic opal (Milliken, 2014). With only few exceptions, appears to have been reached regarding the necessity to
a strong, fabric-controlled clay-mineral and solid-organ- partition the rock into two phases (Sondergeld et al., 2010;
ic-matter preferred orientation in bioturbation-free organ- Alfred and Vernik, 2012) — the organic phase and the in-
ic shales is observed ubiquitously and is conducive to organic phase — each with its associated pore systems.
multiple observations of elastic, electrical, and mechani- These phases (or compositional domains) are not simply
cal anisotropy in these rocks. an arbitrary construct; they are directly suggested by the
To summarize what we have learned from the petro- nanoscale images that have been acquired for these rocks.
logic data on unconventional shale reservoirs, the follow- The log model advanced by Alfred and Vernik (2012) fol-
ing conclusions, bearing on their elastic properties, can be lows an approach used in the past for understanding other
drawn: geologic materials with a heterogeneity scale that is orders
of magnitude smaller than the measurement scale.
• Despite the great mineralogic variety found in organ- Alfred and Vernik’s unconventional-shale log model is
ic mudrocks, certain groups can be distinguished, based on the premise that hydrocarbon fluids occupy the
such as shales, marls, limestones, siliceous mud- kerogen-related porosity, whereas water almost exclusively
stones, and less-ubiquitous clay-dominated shales. fills the nonkerogen matrix porosity. This premise effec-
• Organic richness, which typically is defined by TOC tively eliminates the need to compute water saturation with
content for a general rock description, should be better the conventional, resistivity-based methods that are deemed
described for rock-physics purposes by using the questionable in organic shales.
­solid-organic-matter (predominantly kerogen) volume The innovative aspect of this approach is that the model
fraction. That fraction, of course, correlates strongly solves for the kerogen porosity created by the organic
with the TOC content, from which it is derived. diagenesis that invokes the solid-to-liquid organic-
­
• In combination with the kerogen volume fraction, the matter-transformation reaction. To corroborate the m ­ odel’s
thermal maturity of organic matter — that is, its or- ­assumptions, we again resort to microstructural observa-
ganic metamorphism — should have a significant im- tions — this time, focusing on micropore and nanopore
pact on mechanical and reservoir properties. (It distributions in organic mudrocks. Discussion of the model
should be noted that the database used in this book is details and applications will follow.
strongly partial toward oil and condensate windows,
and that should be kept in mind in the rock-physics Microstructural observations
analysis to follow.)
• The submicroscopic rock fabric in tight unconven- The examples of high-resolution nano-computerized
tional shales, with TOC values ranging from 2 to tomography (nano-CT) and focused ion beam-scanning
12%, is almost exclusively characterized by a bed- electron microscopy (FIB-SEM) imaging provided in
ding-parallel preferred orientation, which is ampli- Figure 11 suggest that two principal kinds of pores coex-
fied by, but not entirely dependent on, the clay vol- ist in mature organic shales: (1) organic-matter site-relat-
ume fraction. The exception to this rule is seen only ed pores, and (2) inter- and intragranular pores that are
in limestones and siliceous mudstones with reduced associated with mineral components. The latter can have
clay content and TOC values below 3% (and possibly, various shapes, from stiff and triangular to softer and
but rather speculatively, above 15 to 20%). more elongated (both in 2D).
• Kerogen displays a marked diversity in textural ap- An aspect ratio α in the range from 0.05 to 0.30 is most
pearance, but generally its prevailing texture in high- typical of the second type of pores. Varying in size from
er-TOC shales is a lenticular latticework, whereas nanopores to larger, 0.1- to 0.2-μm intergranular voids,
scattered microlenses and other inclusion shapes these pore types typically are hosted by clay platelets, but
dominate in lower-TOC shales. they also can be associated with rock-fabric disruption by
• Petrologic data in general, including rock composition, silt grains of quartz, feldspar, and pyrite. All these pore
texture, organic richness/maturity, and fabric-con- types appear to be independent of the solid organic matter
trolled microcracks, are indispensable in interpreting sites. Overall, these FIB-SEM observations clearly justify

06-Vernik_Ch06.indd 129 12-08-2016 19:59:55


130 Seismic Petrophysics in Quantitative Interpretation

a) (1 – φk)
Vk
φk

φnk
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Vnk
(1 – φnk)

b) K ρk
ρm
(1 – K) ρnk

Figure 12.  Schematic representation of the dual nature of


organic shale on (a) the bulk-rock basis and (b) the solid-rock
(porosity-free) basis, illustrating the partitioning of porosity
and solid-phase density. Black is the solid organic matter
volume; green is the bulk hydrocarbon volume; blue is
the bulk water volume; gray is the volume of mineral
components. Adapted from Figures 3 and 4 of Alfred
and Vernik (2012). Used by permission.

c­ hanges in coal properties, such as luster, elasticity, and


brittleness. The type of hydrocarbon that is generated de-
pends on the type of kerogen involved (marine, lacustrine,
or terrestrial) and its thermal maturity.
It is apparent from microstructural observations that
maturation-induced pores might be interconnected and
provide flow paths for hydrocarbons via continuous ker-
ogen latticework permeating organic shales. The pore
pressure generated in the process of solid-to-liquid or-
ganic-matter transformation could also result in micro-
hydraulic-fracture-induced subhorizontal microcracks.
The latter may enhance conductivity for initial flow
(Vernik, 1994; Vernik and Liu, 1997), but barely contrib-
ute to the porosity of organic shales if the dominant as-
Figure 11.  High-resolution FE-SEM images showing (a) pect ratio of these microcracks does not exceed 0.01.
kerogen pores, and (b) other pore types not associated with This hydrocarbon-filled system of kerogen pores and
kerogen, in organic shales. These images suggest the dual- microcracks coexists with the nonkerogen phase that has
porosity nature of mature organic shales. Courtesy of negligible (nanodarcy) permeability. Maturation-induced
K. Milliken, Bureau of Economic Geology, University of pores and microcracks are likely to be at least partially
Texas, Austin. Used by permission. responsible for the microdarcy-range horizontal permea-
bility that typically is invoked from the flow capacity of
partitioning the pore system into o­ rganic-matter and min- production wells. Although it is hard to totally rule out the
eral phases, with their respective porosities. possibility that liquid hydrocarbons (e.g., bitumen) gener-
Kerogen pores form in response to thermal matura- ated by the kerogen sites could diffuse into the adjacent
tion and a solid-to-liquid transformation reaction, as was inorganic pores (Tissot and Welte, 1978), such a process
envisioned by Tissot and Welte (1978). The process ulti- seems more likely to have a limited impact on the rock’s
mately results in dewatering and wettability reversal of storage capacity.
those kerogen sites to create an arguably oil-wet system Figure 12a is a conceptual representation of the dual-
that is saturated with hydrocarbon fluids (e.g., Price et al., phase system with a porosity partitioning that is suggested
1984). This consistently leads to an increase in electrical by microstructural images of organic shales. Alternatively,
resistivity, which normally is associated with mature or- the sum of the kerogen-domain and nonkerogen-domain vol-
ganic shales (Meissner, 1984). The residual solid kerogen umes containing the partitioned porosities can be expressed
undergoes densification, as is illustrated in Figure 4, in as Vk + Vnk = 1. Figure 12b schematically depicts partition-
a way that is similar to coal-metamorphism-related ing on the solid-grain basis with solid kerogen K and solid

06-Vernik_Ch06.indd 130 12-08-2016 19:59:56


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 131

inorganic (1 − K) volume fractions and their respective maturity through LOM results in the log-scale TOC
solid-phase densities ρk and ρnk. curve, which can be calibrated by core data. Alternative
The petrophysical model, described next, is based on methods of log-based TOC computation, such as regres-
these schematics and is designed primarily for computing sion of core-based bulk density versus TOC (Schmoker,
the total porosity, kerogen porosity, and water saturation 1979), is less effective, as will be discussed below.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

of organic mudrocks with widely varying maturities. The Proper calibration of the TOC log is paramount for the
inputs to the computations are standard well data such as overall accuracy of the log model, because any errors in
resistivity, neutron, density, and sonic logs, complement- the calculation of the kerogen fraction propagate and am-
ed with sporadic core measurements of organic richness plify in further evaluation of reservoir parameters. Errors
and maturity. in the TOC log model are exaggerated by any uncertainties
with the organic carbon fraction Ck in kerogen and kero-
gen density. As Figures 3 and 4 show, kerogen density can
Kerogen-fraction log be estimated from the vitrinite reflectance or the HI. The
As is implied by the schematic description of the du- Ck value also can be constrained by the same parameters
al-phase organic shale system shown in Figure 12, the because it has been shown to increase, with maturity, from
volume fraction of the organic (or kerogen) phase Vk in- approximately 65% to roughly 90% of the solid kerogen,
cludes oil-wet porosity φk saturated almost exclusively by during the process of liquid hydrocarbon generation. As a
the hydrocarbon fluid, whereas the inorganic (nonkero- first approximation, Ck = 15 + 50ρk can be used, but this
gen) phase is characterized by its own porosity φnk, which relationship generally is rather “soft” and always should
is fully water-saturated. In addition, it is instructive to be verified by empirical log-to-core comparisons.
introduce two bulk densities, ρbk and ρbnk, two fluid densi-
ties, ρfk and ρ fnk, and two water saturations, Swk and Swnk, Total porosity and kerogen porosity
to describe the two respective phases of the whole rock.
All these additional variables are depicted schematically As was mentioned above, the kerogen-fraction log is
in Figure 13, which is designed to illustrate more details computed on the basis of the solid matrix. If φ is the total
than Figures 12a and 12b. porosity of the rock, the kerogen fraction on the bulk-rock
It follows from Figure 13 that computation of the basis can be found from
solid-kerogen volume fraction K is a principal part of the
petrophysical model; the K value can be derived from vk = K (1 − φ ), (4)
either the TOC log or the core data by using the exact
equation 3 or the empirical approximation mentioned and the bulk-organic-phase volume fraction is
above and shown in Figure 5. In turn, the standard log- K (1 − φ )
based TOC-estimation technique is employed and then Vk = . (5)
1 − φk
calibrated with core pyrolysis data (Passey et al., 1990).
This method relies on the premise that porosity logs By substituting for K from equation 3, the organic-phase
(density and/or sonic logs) respond to a low-density ker- volume can also be expressed as
ogen volume and total porosity, whereas resistivity logs
are largely affected by the type of fluid. Subtracting the TOC ρnk (1 − φ )
Vk = , (6)
properly scaled versions of those logs from one another A
and taking into account the impact of the thermal
where

A = (1 − φk )[TOC ( ρnk − ρk ) + Ckρk ]. (7)


vk = K(1 – φ) ρk
Vk ρbk
φk Swk ρfk The bulk density of the rock is computed through phase
φnk Swnk ρfnk (domain) volumes and their densities:
Vnk ρbnk
ρnk ρb = Vk ρbk + (1 − Vk )ρbnk . (8)
vnk = (1 – K )(1 – φ)

Because by definition Swk ≈ 0 and Swnk ≈ 1, it is possible


Figure 13.  Schematic representation of the two-phase to express the organic and inorganic phase densities as
composition of organic shales on a bulk-rock basis, including
all the variables for the log-model computations. Color ρbk = ρk − φk ( ρk − ρhc ),
definitions are the same as in Figure 12. Adapted from Figure (9)
ρbnk = ρnk − φnk ( ρnk − ρw ).
6 of Alfred and Vernik (2012). Used by permission.

06-Vernik_Ch06.indd 131 12-08-2016 19:59:57


132 Seismic Petrophysics in Quantitative Interpretation

The combination of equations 6 through 9 leads to the uncertainty in the kerogen-porosity prediction from the
bulk-density expression bulk-density-TOC template constructed from equation 10,
in the absence of XRD data, is the grain density of the non-
TOC ρnk (1 − φnk )( ρbnk − ρbk ) kerogen phase. Therefore, having at least one or two cored
ρb = ρbnk − , (10)
A + TOC ρnk (φk − φnk ) wells in the specific geographic location is always a plus.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/


Schmoker (1979) presented a relationship between
which should be regarded as the fundamental relationship bulk density ρb and TOC values from core measurements
between bulk density and TOC in organic mudrocks. That in a Devonian organic shale play, and it is included in
relationship is illustrated in Figure 14 as a template for a Figure 14. It is obvious that this empirical model can be
plausible range of kerogen porosity values, from 10 to matched accurately by the equation 10 realization for ker-
50%, that are calculated on the basis of the bulk kerogen ogen porosity of 17 ± 1%, a kerogen-free matrix density
volume Vk. ρnk of 2.765 g/cm3, and a nonkerogen porosity of 3.8%.
Inorganic (nonkerogen) porosity in unconventional Therefore, Schmoker’s transform is nonunique in organic
reservoirs typically ranges from 3 to 4%, as is calculated shales and can be applied only in a certain narrow range of
on the basis of the bulk inorganic volume Vnk, when this thermal maturities. Clearly, equation 10 provides a more
model is applied in unconventional shale plays. The solid complete characterization of the bulk-density-versus-TOC
density of the inorganic phase ρnk is best estimated from relationship in organic mudrocks. Of particular interest in
XRD mineralogy data, whereas the kerogen density ρk practical applications is the observation that the numerical
and organic carbon fraction Ck can be derived from ther- values of φnk and ρnk selected for the template in Figure 14
mal-maturity indicators. are also applicable to many other unconventional reser-
Note that at zero TOC content, the model lines con- voirs that may not necessarily fall into the category of
verge on the bulk density of the inorganic phase of the rock shale in the classification scheme of Figure 1.
ρbnk, which, from the specified input parameters and equa- The total porosity φ of the rock is a linear function of
tion 9, is 2.70 g/cm3. The lines of constant kerogen porosity its bulk density and is given by
in Figure 14 are likely to correspond to constant maturity
levels; these lines help us visualize the kerogen porosity ( ρm − ρb )
φ = , (11)
estimated by the model from the TOC and bulk-density ( ρm − ρf )

logs. Aside from the TOC log discussed above, the largest
where the solid matrix density ρm can be conveniently ex-
pressed, on the basis of equation 3, as

C k ρk ρnk
ρm = . (12)
TOC ( ρnk − ρk ) + C k ρk

From the estimated kerogen porosity, and assuming


­perfect water and hydrocarbon-fluid partitioning, the
effective fluid density of the rock can be computed and
plugged into equation 11 for a total-porosity evalua-
tion. Alternatively, the total porosity can be expressed
through ρbk and ρbnk, which explicitly contain fluid-den-
sity information:

A ( ρbnk − ρb)
φ = 1− . (13)
TOC ρnk ( ρbnk − ρbk )

Figure 15 shows this relationship between the total po-


rosity and TOC for the same range of kerogen porosities
and using the same input parameters as in Figure 14. Note
Figure 14.  Model lines showing the relationship between that the model lines converge on the porosity intercept of
bulk density ρb and TOC content in organic mudrocks with 3.8%, which is the chosen porosity of the nonkerogen
varying kerogen porosities φk. The input-parameter values phase. Because of this convergence, at a low level of or-
(densities are in g/cm3 and Ck is in percent) are chosen for a ganic richness (TOC < 2%) the stability of total-porosity
particular gas shale play in North America. Adapted from and kerogen-porosity estimates is predictably reduced.
Figure 7 of Alfred and Vernik (2012). Used by permission. Even so, the TOC-porosity template in Figure 15 illustrates

06-Vernik_Ch06.indd 132 12-08-2016 19:59:58


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 133
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 15.  Model lines showing the total-porosity-versus- Figure 16.  Model lines showing the water-saturation-versus-
TOC-content relationship in organic mudstones with varying TOC-content relationship in organic mudstones with varying
kerogen porosities. The input parameter values (densities are kerogen porosities. The input parameter values (densities are
in g/cm3 and Ck is in percent) are the same as in Figure 14. in g/cm3 and Ck is in percent) are the same as in Figure 14.
Adapted from Figure 8 of Alfred and Vernik (2012). Used by Adapted from Figure 9 of Alfred and Vernik (2012). Used by
permission. permission.

in a quantitative manner how the total porosity of organic


mudrocks increases with thermal maturity. φk (1 − φ )TOC ρnk
Sw = 1 − . (15)
As previously mentioned, that increase in porosity φA
is almost entirely attributed to new porosity generation
through kerogen transformation, in agreement with the ob- The crossplot in Figure 16 illustrates the relationship be-
servation by Loucks et al. (2009) on Barnett Shale cores. tween the bulk water saturation and TOC values, again as a
However, it would be incorrect to propose that kerogen po- function of kerogen porosity. The plot suggests that for a
rosity is a monotonic function of thermal maturity, because given TOC, an increase in thermal maturity (and hence in
organic pores may also tend to collapse with greater burial kerogen porosity), to at least stage V (i.e., the wet-gas win-
depths and maturity. That is especially true when matura- dow), results in a simultaneous decrease in formation-water
tion-induced overpressure dissipates as a result of subse- saturation. It also indicates that although Sw always decreas-
quent natural fracturing and hydrocarbon migration (e.g., es with organic richness in oil- and wet-gas-mature shales,
the Marcellus Shale, Pennsylvania, USA). the rate of that decrease is greater in lower-TOC rocks.
Finally, it is noteworthy that kerogen porosity com-
puted from the model presented can be physically con-
Water saturation strained. By rearranging equation 15 in accord with equa-
On the basis of the simplifying assumption of perfect tion 5, we can write that
saturation partitioning, it is evident that the bulk-volume
φk (1 − φ )K
fraction of hydrocarbons in our log model is given by the Sw = 1 − . (16)
product φShc = φkVk. The resistivity-independent water φ (1 − φk )
saturation can then be derived from
Because by definition Sw ≤ 1, we obtain a useful inequality:
φV
Sw = 1− k k . (14) φ
φ φk ≤ . (17)
(1 − φ )K + φ
By using equations 6 and 14, the Sw function can be ex-
pressed in terms of the more explicit variables that are It follows that there is always an upper bound on such
available at this stage of log-model application: an elusive parameter as kerogen porosity. That bound

06-Vernik_Ch06.indd 133 12-08-2016 19:59:59


134 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 17.  Log model applied in an Eagle Ford well, USA. Adapted from Figure 10 of Alfred and Vernik (2012).
Used by permission.

is imposed by the total porosity and the solid kerogen Figures 17, 18, and 19, which display log models for three
­volume, both of which can be readily verified by core cored pilot wells from different fields in the USA. The un-
measurements. conventional shale formation in each well is at thermal-
maturity stages V or VI, which correspond respectively to
Model applications the wet-gas (condensate) and dry-gas windows (Table 2).
All three wells share the same log-display layout:
The petrophysical log model developed by Alfred and track 1: GR, CALI; track 2: DEPTH; track 3: DTC, RT;
Vernik (2012) and described above in more detail was ap- track 4: RHOB, NPHI; track 5: TOC (core), TOC; track 6:
plied to many vertical pilot wells with adequate log sets, in PHIT (core), PHIT; track 7: BVH, PHIT; and track 8: SW
a variety of unconventional shale reservoirs saturated with (core), SW. Core data are shown as black dots. The log
different fluids, including oil, gas, and condensate. The mnemonics are as follows: GR is gamma ray, CALI is
model predicts porosity, saturation, and TOC distribution caliper, DTC is P-wave traveltime, RT is horizontal resis-
with reasonable accuracy, notably in cored wells, where tivity, RHOB is bulk density ρb, NPHI is neutron porosity
core measurements of TOC, crushed core porosity, and φn, PHIT is total porosity φ, BVH is bulk volume hydro-
water saturation are available. The results are presented in carbon, and SW is water saturation Sw.

06-Vernik_Ch06.indd 134 12-08-2016 20:00:01


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 135
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 18.  Log model applied in a Woodford well, USA. Adapted from Figure 11 of Alfred and Vernik (2012). Used by
permission.

It is important to realize that some of these organic formations in most cases are composed of tight lime-
shale formations are more heterogeneous than others, on stones with subordinate conventional, low-TOC shales.
core and log scales. Therefore, most of the discrepancies Moreover, the development of kerogen porosity causes
between the log-model-computed curves and core mea- the total porosity to correlate well with organic richness
surements are related to differences in vertical resolution, inside unconventional shale units. The kerogen porosi-
which can be lower than is required to adequately evalu- ties for the three representative wells selected vary with-
ate thin beds. For example, Figure 17 shows several low- in the 20 to 40% range, independently of TOC content,
TOC cores, from the middle portion of the Eagle Ford but, as multiple applications of the log model to other
interval, that were sampled from thin limestone beds that organic mudrock reservoirs reveal, they are controlled
are below vertical resolution of logs employed in TOC by the shale’s maturity level. These results are consistent
computation, such as resistivity and sonic. with observations made earlier by Loucks et al. (2009)
In Figures 17, 18, and 19, the total-porosity track, and Sondergeld et al. (2010).
which always spans the 0 to 15% range, shows that or- It is quite surprising to observe in all three log-model
ganic shales generally tend to have noticeably higher displays (Figures 17 through 19) a very close match be-
porosities than do encasing geologic formations. Those tween the water-saturation log, which is computed on the

06-Vernik_Ch06.indd 135 12-08-2016 20:00:03


136 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 19.  Log model applied in a Haynesville well, USA. Adapted from Figure 12 of Alfred and Vernik (2012).
Used by permission.

basis of the perfect-fluid-partitioning assumption for kero- do not impact the volumetrics). On the other hand,
gen and nonkerogen phases of the rock, on one hand, and the inorganic phase of the mudrock is water-wet and,
core measurements of water saturation on the other. It is hence, is fully water-saturated due to the extremely
still quite possible to have oil and gas in subhorizontal mi- high local capillary pressure that must be e­ xerted by
crocracks that are expected to propagate into the inorganic the nonwetting hydrocarbons to penetrate the non-
rock components, but because of an exceedingly low crack kerogen pore network.
porosity (see Chapter 3) those microcracks may not impact • The key input parameters that bear on the accuracy
the volumetrics of unconventional shale reservoirs. of the model predictions for porosity and saturation
A few concluding remarks are warranted on the log are the TOC content (or the volume fraction of kero-
model presented above: gen in the solid-rock matrix) and the solid-grain den-
sity of the inorganic phase of the rock. These param-
• The key reservoir parameters (TOC, porosity, and eters should be evaluated independently from XRD
water saturation) of an unconventional shale reservoir mineralogy for any new geographic location. In gen-
can be fairly accurately evaluated by using a standard eral, model predictions become less stable at a TOC
wireline log suite of gamma-ray, resistivity, sonic, content below 2%.
and density logs.
• The petrophysical log model is based on the assump- The accuracy of the TOC, porosity, and saturation
tion that the hydrocarbon fluid is almost exclusively computations can only be verified by measurements on
contained in oil-wet kerogen (with the possible ex- crushed core samples. Such a calibration is also a critical
ception of overpressure-induced microcracks, which element in experimental rock-physics data analysis and

06-Vernik_Ch06.indd 136 12-08-2016 20:00:05


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 137

modeling of seismic rock properties in organic mudrocks. adding confusion to the problem of shale identification in
Those subjects will be discussed in detail in the remainder sedimentary geology. Third, the methodology of ultrasonic
of this chapter. measurements differs in time from one laboratory to an-
other, making it difficult to identify measured parameters
and determine their accuracy. Additional problems arise
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Rock physics of unconventional shales because of insufficient control of core inhomogeneity,


which in lab measurements can be easily mistaken for ve-
In principle, seismic petrophysics focuses on the locity anisotropy when individual plugs, drilled out of core
same problem for organic shales as it does for conven- samples, are used. Some of these issues are addressed next.
tional reservoirs — relating seismic attributes to key res-
ervoir parameters. The major difference is the parame- Core measurements of velocity and anisotropy
ters of interest. For instance, the organic richness of
thermally mature mudrocks (measured as TOC or K) is An ultrasonic pulse transmission technique has been
of particular importance because of its strong correla- used predominantly to measure elastic wave velocities
tion with hydrocarbon-filled porosity, as was discussed and anisotropy in laboratory experiments, and Vernik and
in the previous section. It is also critical to specify the Nur (1992a) and Vernik and Liu (1997), among others,
kerogen maturity level, which is used to predict hydro- have employed it for velocity measurements in mudrocks
carbon fluid types (oil, condensate, or dry gas). All of with varying organic contents. The technique uses tablet-
these factors are important targets for any seismic petro- shaped, stacked piezoelectric transducers with a 1-MHz
physics investigation. central frequency for P-waves and a 0.7-MHz frequency
Textural characteristics of organic mudrocks require for S-waves. The accuracy of P-wave velocity measure-
not only standard velocity measurements but also elastic ments is approximately ±1%, and that for S-wave veloci-
anisotropy evaluation. In fact, the latter can shed light on ties is ±2%. Assuming that there is transverse isotropy
the still debatable issue of the mechanism for primary hy- (TI), which is often also referred to as polar anisotropy,
drocarbon migration from micropores and nanopores to with the symmetry axis normal to bedding (which is typi-
preexisting natural fractures and hydraulically induced cally confirmed by data-acquisition redundancy), the ve-
ones. As Vernik and Nur (1992a) demonstrated, all three locity anisotropy is calculated from velocities measured
known causes of elastic anisotropy operate in organic on three plugs cut from each core sample: (1) normal to
shales: small-scale interlayering, preferred orientations of bedding (at 0° to the symmetry axis), (2) at 45° to bed-
kerogen and clay platelets, and nonrandomly oriented mi- ding, and (3) parallel (90°) to bedding. Vernik and Nur’s
crocracks. One task of rock physics is to try to separate cores were 5 to 7 cm long and 10 cm in diameter, whereas
the effects of these three factors and to interpret seismic their plugs were 30 to 40 mm long and 20 to 25 mm in
attributes in terms of organic richness, maturity, and per- diameter. Figure 20 is the core-plug schematic, showing
haps even the effective stress tensor. wave-propagation and polarization directions with re-
Modeling of unconventional-shale elasticity has been spect to the bedding plane.
extensively discussed in the geophysical literature for the Velocity was measured in room-dry and partially wa-
last two decades (e.g., Vernik and Nur, 1992a; Johnston ter-saturated rocks at a range of effective (confining)
and Christensen, 1995; Vernik and Liu, 1997; Hornby,
1998; Sondergeld et al., 2000; Lucier et al., 2011; Vernik VSH(0)
VP(0) VP(45) VP(90)
VSH(45) VSH(90)
and Milovac, 2011; Sayers, 2013; Allan et al., 2014; VSV(0) VSV(45) VSV(90)
Khadeeva and Vernik, 2014). Even though our under-
standing of the microstructural control on velocities and
anisotropy of organic shales has advanced, some mea-
surements and conceptual models are still controversial
and cause persistent discrepancies in the interpretation.
Some of the problems with laboratory experimental
studies are listed briefly here. First, shales are rarely cored, Figure 20.  The three-plug principle, illustrating P- and
and when they are, core-preservation requirements with S-wave propagation and polarization directions with respect
respect to native fluids are not strictly adhered to and to the bedding plane of an organic shale core sample. Nine
core plug preparation, notably in mature organic-rich independent velocity measurements are conducted on each
shales, is adversely affected by the excessive brittleness core (three per plug). The numbers in parentheses indicate the
of the core material. Second, petrophysical studies in the phase-velocity angle θ with respect to the bedding-normal
laboratory are rarely accompanied by detailed petro- symmetry axis. Adapted from Figure 1 of Vernik and Nur
graphic observations and compositional analyses, thereby (1992a). Used by permission.

06-Vernik_Ch06.indd 137 12-08-2016 20:00:06


138 Seismic Petrophysics in Quantitative Interpretation

pressures from 5 to 70 MPa, in jacketed/drained tests, with room-dry conditions, for all wave modes and orienta-
with a velocity-equilibration time often exceeding 3 tions except the P-wave normal to bedding. It is possible
hours, especially for plugs that were normal to bedding. that the chemical softening due to incompatible water
The pore fluid was allowed to drain upon hydrostatic chemistry is countered by mechanical stiffening of water-
loading, so that the initial pore-pressure buildup decayed filled, bedding-parallel, low-aspect-ratio pores and micro-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

exponentially. That pore-pressure dissipation was moni- cracks, thereby resulting in the observed negligible net ef-
tored in time as a corresponding increase in velocity. fect on VP(0°). Note, however, that the normal-to-bedding
Therefore, the effective pressure (i.e., the isotropic stress) velocity ratio VP/VS is dramatically altered (increased) in
in fluid-saturated plugs was approximately equal to the the process. Such an increase, however, is not consistent
confining pressure — that is, σ ≈ Pc. with sonic-log data acquired in the same well from which
Vernik and Liu (1997) give an overview of the “satu- those shale cores were extracted, thereby suggesting that
rated” measurements, including the chemical softening care must be exercised when saturating even relatively
­effect of saturating fluids on certain organic shales. An inert, illite-dominated shales with nonnative brine.
­example of that effect is illustrated in Figure 21, which In this book the focus is on room-dry measurements
shows brine-saturation-induced shale softening, compared for the following reasons: (1) the plugs contain viscous
bitumen and liquid oil in their oil-wet micropores and
nanopores, and those heavy hydrocarbons are difficult to
extract without pyrolysis; (2) despite some water loss from
inorganic pores, and an arguably incomplete forced water
saturation due to mixed wettability, the ultrasonic wave
pulse is unlikely to be conducive to a pore-pressure relax-
ation that would be capable of softening the rock 1 MHz
frequency, (3) forced saturation with nonnative brine may
induce undesirable chemical softening effects, and (4) ve-
locities and velocity ratios measured normal to bedding on
room-dry core plugs, which arguably contain measurable
amounts of liquid hydrocarbons, are much more consistent
with sonic-log data obtained from vertical pilot wells.
For all three core-plug orientations depicted in
Figure 20, the S-wave polarization direction with re-
spect to bedding is consistent with SH- and SV-modes,
with the latter always being polarized in the plane per-
pendicular to bedding. This provides the redundancy
necessary to confirm the TI assumption and/or cross-
validate traveltime picks for all three wave modes.
Density measurements on each plug (with an accuracy
of ±0.01 g/cm3) are used as checks on core homogenei-
ty; only the samples with a density variation of less than
2% among the three plugs were chosen for a subsequent
calculation of anisotropy.
Transversely isotropic media can be adequately de-
scribed by five independent elastic constants: C11, C33,
C44, C66, and C13. After White (1965), the three phase ve-
locities of waves propagating at an angle θ to the TI sym-
metry axis are given by

2 C 66 sin 2 θ + C 44 cos2 θ
V SH = ,
ρ
(18)
A + B ± ( A − B )2 + 4G 2
Figure 21.  P-wave and SH-wave velocities versus confining 2
V P,SV = ,
pressure for a black shale containing 6% smectite, showing a 2ρ
dramatic effect from incompatible brine saturation (Bazhenov
shale, core from 3788 m). Adapted from Figure 7 of Vernik where A = C11 sin 2 θ + C44 cos2 θ , B = C44 sin 2 θ + C33 cos2 θ ,
and Liu (1997). Used by permission. and G = (C13 + C44 ) sin θ cos θ . Velocities VP and VSV in

06-Vernik_Ch06.indd 138 12-08-2016 20:00:07


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 139

anisotropic media correspond to mixed modes of wave


propagation — to the quasi-compressional (qP) and qua-
si-shear (qS) modes, respectively. Equations 18 reveal
that from the pulse traveltime measurements in the direc-
tions normal (θ = 0°) and parallel (θ = 90°) to bedding,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

four elastic constants can be calculated (see Chapter 1),


whereas at least one measurement at an oblique angle is
required to estimate C13. The phase velocity VP(45°) is
often chosen as an input for calculating C13 (Lo et al.,
1986), which is the most elusive and difficult to measure
element of the TI elastic tensor.
Among the challenges associated with proper rock-
physics modeling of P- and S-wave velocities in organic
shales, accurate evaluation of the shale’s porosity is ar-
guably the largest. Shale porosity is not easily and un-
ambiguously determined from core data or log data. For
instance, the total porosity is grossly underestimated for
the tight organic shales that Vernik and Liu (1997) mea-
sured in a helium porosimeter at ambient conditions on
cores oven-dried at 80°C. More-accurate porosity values
can be obtained with the crushed-core technique, using a Figure 22.  Bulk-density versus TOC values for several cores
so-called GRI protocol developed by the Gas Research sampled from organic shale formations worldwide,
Institute (currently the Institute of Gas Technology). superposed by model realizations for 5% porosity (green) and
Vernik and Milovac (2011) employed the bulk-density- 10% porosity (red). The uncertainty range for mineral density
versus-TOC model, a simplified version of equation 10, on a kerogen-free basis (ρnk) is shown for each realization.
to analyze Vernik and Liu’s (1997) results, and they con- Adapted from Figure 4 of Vernik and Milovac (2011). Used
cluded that the true total porosity of the tight black shale by permission.
cores they tested ranges from 4 to 10% (Figure 22).
The phase velocities of P- and SV-waves for small angles
of incidence can be approximated by the elliptical func-
Phase velocity versus group velocity tions that hold for any degree of anisotropy (Vernik and
The phase velocity of wave propagation in an aniso- Nur, 1992a; Tsvankin and Thomsen, 1994; Alkhalifah and
tropic formation is defined as the velocity of the wave- Tsvankin, 1995):
front. Generally, this velocity is different from the v­ elocity 1
of elastic energy that is traveling in the same formation
along a ray connecting a point source and a receiver point,
(
V P (θ ) ≈ V P (0 ) 1 + 2 δ sin 2 θ ),
2

1
which typically is referred to as the group velocity. The
direction normal to the wavefront deviates noticeably ( )
V SV (θ ) ≈ V S (0 ) 1 + 2 δ SV sin 2 θ 2
, (20)
1
(θ ) = V (0 ) (1 + 2 γ sin θ ) .
from the ray direction in highly anisotropic rocks.  2
However, the velocity values and their respective angles V SH S
2

coincide in the principal directions in TI media, such as
the symmetry axis and the symmetry plane. The velocities and the Thomsen’s anisotropy parameters
Conversion from the phase velocity V(θ) and phase are defined in Chapter 1, except for δSV, which is given by
angle θ to the group velocity v(ψ) and group (or ray) angle Vernik and Nur (1992a):
ψ should be done using exact relations (Thomsen, 1986):
(C11 − C44 )(C33 − C44 ) − (C13 + C44 )2
δ SV = . (21)
2 C33 (C33 − C44 )
 dV 
2
2
v(ψ ) = V (θ ) +  ,
 d θ 
The five elastic constants for TI rocks are also defined in
1 dV (19) Chapter 1. Equations 20 disclose the meaning of the an-
tan(θ ) +
tan(ψ ) = V dθ . isotropy parameters δ and δSV as the ratios of the axes of
tan(θ ) d V the ellipses approximating the P-wave and SV-wave
1−
V dθ phase-velocity surfaces in the small range of angles from

06-Vernik_Ch06.indd 139 12-08-2016 20:00:07


140 Seismic Petrophysics in Quantitative Interpretation

0 to approximately 25°. Note also from equations 20 that


the velocity surface of the SH-wave is elliptical over the
entire range of angles and that the shear-wave anisotropy
parameter γ amounts to the eccentricity of this ellipse.
Curiously, at θ = 90°, equations 20 yield the normal
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

moveout (NMO) velocities of P- and SV-waves, even


though the latter are evaluated at θ = 0° in seismic
applications:
1
vP = VP (0 ) (1 + 2 δ ) 2,
NMO
(22)
1
vSV = VS (0 ) (1 + 2 δ SV ) . 2
NMO

Equations 19 imply that the respective small-angle ap-


proximations for the group velocities are (Vernik and Liu,
1997):
1

 2δ sin 2 ψ  2
vP ≈ V P (0 ) 1 − , (23)
 1 + 2δ 

1

 2δ sin 2 ψ  2
v SV ≈ V S (0 ) 1 − SV  ,
 1 + 2δ SV 

where ψ is the group angle — that is, the incident angle of


the seismic ray, as opposed to the angle normal to the
wavefront, which is the phase angle θ. The nonelliptical
functions given by equations 23 converge with expres-
sions 20 for the incident angle of 90° to the symmetry
axis (θ = ψ = 90°).
Small-angle approximations for the phase velocities
(equations 20) and the group velocities (equations 23),
labeled as elliptical velocity functions Ve(θ) and Ve(ψ),
respectively, are added to the velocity surfaces that fit the Figure 23.  P-wave phase-velocity and group-velocity
phase-velocity measurements on a highly anisotropic surfaces for a highly anisotropic black shale (Bakken
sample of the upper Bakken shale in Figure 23. The true Formation, 3438 m, dry), computed from velocity anisotropy
NMO velocity is recovered from these functions at 90°, as measurements at confining pressures of (a) 5 MPa and (b)
was mentioned above. Figure 23a refers to the measure- 70 MPa. Note the decreasing difference between phase and
ments and computations at the low stress of 5 MPa, group velocities with decreasing anisotropy, from (a) ε = 0.78
whereas Figure 23b presents the results at the high stress to (b) ε = 0.23, and the very good fit provided by equations
of 70 MPa. Analysis of these plots indicates that for inci- 20 and 23 to the phase- and group-velocity surfaces at small
dent angles as high as 25°, equations 23 approximate the angles (Ve and ve). Adapted from Figure 4 of Vernik and Liu
group-velocity surfaces of the P-waves and the SV-waves (1997). Used by permission.
to within 1%, even at an anisotropy as high as ε = 0.78.
It also follows from Figure 23 that the group velocity on the 45° core plug: Do those measurements represent
of P-waves (shown as thick blue lines) measured along the phase velocity or the group velocity, or, perhaps,
the rays in nonprincipal directions is lower than the phase something in between?
velocity (shown as thick red lines) measured along the According to the definitions of phase velocity and
wavefront-normal direction. The difference is determined group velocity mentioned above, ultrasonic traveltimes
by the strength of anisotropy and reaches a maximum at an- measure either the anisotropic group velocity (if the
gles of approximately 45°. This observation introduces sig- piezoelectric transducers are very small relative to their
nificant uncertainty into the traveltime-pick interpretation separation given by the sample length) or the phase

06-Vernik_Ch06.indd 140 12-08-2016 20:00:09


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 141

velocity (if the transducers are relatively wide relative to receiving transducer (Figure 24). Those erroneously slow
the sample length) (Auld, 1973). The phase vs. group ve- VP(45°) values result in C13 and δ values that are too low,
locity issue was first addressed by numerical modeling of so minor positive corrections to a P-wave velocity mea-
the ultrasonic P-wave pulse-transmission experiment on a sured on 45° plugs longer than 35 mm are justified to bring
45° plug of the specified geometry extracted from the the values “up to the speed” of the true phase velocity.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Bakken shale core (Dellinger and Vernik, 1994). Their Such velocity corrections, which were neglected by
modeling approach is described below. Vernik and Liu (1997), are included in the database
Figure 24 presents a snapshot of the 2D numerical ­compiled for the purposes of this book. Consequently,

wavefield excited by the 12-mm-diameter virtual P-wave the d­atabase tables presented here contain improved
transducer, simulated as an array of adjacent point sourc- phase-­velocity estimates of VP(45°), thereby allowing for
es positioned on the flat surface of the shale plug oriented more-accurate computations of the elastic stiffness C13
at 45° to bedding. Note that the flat portion of the wave- and the anisotropy parameters δ and δSV (see equations
front, which mimics the transducer geometry, is deflected 35 of Chapter 1, and equation 21 herein, to appreciate the
in the direction of the bedding plane, which is the fastest effect of C13 on δSV).
traveltime direction in the shale. Indeed, because the It should be emphasized that difficulties exist in pre-
P-wave travels faster along the laminations than across paring plugs for measuring elastic constants in typical or-
them, the flat part of the wavefront containing the main ganic shales with significant mechanical anisotropy and
focus of energy is naturally deflected in the direction in- brittleness, and those problems cause some investigators
dicated by those laminations. to conduct experiments on whole cores by placing the
The most important finding of the numerical simula- ­ultrasonic transducers in three orientations with respect to
tion undertaken by Dellinger and Vernik (1994) is that the the symmetry axis (e.g., Hornby, 1998; Wang, 2002a,
12-mm-diameter P-wave transducers, which are attached 2002b; Ortega, 2010). Although that approach may help
to the plug that is shorter than 35 mm in length, trigger a simplify the issue of sample preparation, it can be quite
first break at the traveltime corresponding to the plane deficient for the following two reasons. First, the full ten-
wavefront’s arrival. By definition, this is the phase veloc- sor description, including derivation of the off-symmetry
ity of the sample. A few longer samples (35 to 40 mm) in stiffness C13 on the basis of inversion schemes, becomes
the collection may generate small negative errors in uncertain due to the confusion between group velocity
VP(45°) (errors typically less than 0.5% in the most highly and phase velocity when one is using relatively small
anisotropic shales), simply because the deflection of the transducers to measure at oblique angles on large cores.
flat portion of the wavefront results in a partial miss of the Second, sample heterogeneity (e.g., laminations thicker
than a few millimeters), if not accounted for, may cause
uncertainty in the bedding-parallel velocities measured
on large cores and, consequently, in the velocity anisot-
ropy of the rock.

Intrinsic velocity and anisotropy


Velocity measurements at high confining pressures of
50 to 70 MPa, which are three times greater than vertical
effective stress typically encountered in unconventional
shale reservoirs, essentially reveal intrinsic elastic prop-
erties of organic shales. The reason is that at those stress
levels the vast majority of low-aspect-ratio microcracks
(α < 10−2) are closed and the direct effects of rock compo-
sition and texture can be evaluated. Table 3, the database
of seismic rock properties of organic mudrocks with
widely varying kerogen concentrations, primarily com-
Figure 24.  Snapshot of a numerical simulation of the P-wave prises the results reported by Vernik and Liu (1997) with
propagation in a 45° core-plug experiment, taken at 10 µs. aforementioned corrections and with additions from
The yellow lines show the width and orientation of the core Tosaya (1982), Lo et al. (1986), Johnston and Christensen
plug (25 mm in diameter), and the thick green segments at (1995), Hornby (1998), and Sondergeld et al. (2000).
the bottom and top depict the sizes and positions of the Data on tight, low-TOC (inorganic) mudrocks, such
P-wave source and receiver transducers. Adapted from Figure as the Chicopee Shale (Lo et al., 1986) and the Cotton
4 of Dellinger and Vernik (1994). Used by permission. Valley Formation (Tosaya, 1982), are added in an attempt

06-Vernik_Ch06.indd 141 12-08-2016 20:00:09


142 Seismic Petrophysics in Quantitative Interpretation

Table 3.  Database of physical properties of mudrocks, measured under a high confining pressure of 50 to 70 MPa, on room-
dry samples. Sources: Vernik and Liu (1997), Sondergeld et al. (2000), Lo et al. (1986), Hornby (1998), Tosaya (1982),
Johnston and Christensen (1995), and M. Aldin (personal communication, 2013).
Depth ρb Kerogen VP(0°) VP(90°) VS(0°) VS(90°) ε γ δ
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

(m) (g/cm3) (v/v) (km/s) (km/s) (km/s) (km/s)


Bakken Formation, North Dakota, USA
2307 2.09 0.444 3.13 3.94 1.88 2.30 0.29 0.25 –
2632 2.06 0.415 3.02 4.14 1.76 2.50 0.44 0.51 0.10
2996 2.21 0.279 3.41 4.16 2.07 2.51 0.24 0.24 0.12
3098 2.22 0.326 3.38 4.19 2.12 2.57 0.27 0.23 0.10
3196 2.18 0.411 3.18 4.20 1.93 2.51 0.37 0.35 –
3199 2.30 0.277 3.62 4.36 2.22 2.63 0.23 0.20 –
3223 2.49 0.202 3.36 4.48 2.06 2.55 0.39 0.27 –
3271 2.33 0.249 3.46 4.49 2.00 2.62 0.39 0.27 0.14
3272 2.34 0.181 3.51 4.42 2.03 2.75 0.29 0.42 0.18
3332 2.35 0.217 3.29 4.67 1.86 2.74 0.51 0.59 0.30
3423 2.55 0.113 4.21 4.97 2.52 2.91 0.20 0.17 0.08
3428 2.38 0.222 3.72 4.36 2.25 2.59 0.19 0.16 0.10
3438 2.48 0.139 3.89 4.71 2.42 2.88 0.23 0.21 0.13

Eagle Ford Shale, Texas, USA


3711 2.51 0.061 4.44 5.00 2.52 2.69 0.13 0.17 0.07
3713 2.66 0.017 5.61 5.62 3.05 3.12 0.00 0.00 0.00
3713 2.47 0.086 4.06 4.70 2.40 2.56 0.17 0.18 0.08
3716 2.50 0.076 4.55 4.86 2.55 2.65 0.07 0.14 0.03
3721 2.40 0.125 4.11 4.68 2.47 2.60 0.15 0.14 0.07
3723 2.64 0.059 5.33 5.33 3.00 3.06 0.00 0.04 0.00
3729 2.59 0.030 5.41 5.60 2.92 3.13 0.04 0.10 0.02
3736 2.55 0.068 4.78 4.92 2.70 2.81 0.03 0.07 0.02
3742 2.49 0.169 4.65 4.87 2.66 2.75 0.05 0.08 0.02
3752 2.63 0.020 5.24 5.18 2.90 2.90 –0.01 0.06 –0.01
3753 2.51 0.149 5.00 5.24 2.84 3.05 0.05 0.11 0.02
3754 2.44 0.060 4.31 4.78 2.58 2.87 0.12 0.17 0.06
3754 2.53 0.144 4.91 5.18 2.83 3.04 0.06 0.11 0.03
3754 2.45 0.104 4.18 4.75 2.47 2.77 0.15 0.20 0.07
3755 2.40 0.145 4.25 4.71 2.55 2.70 0.11 0.07 0.06
3758 2.44 0.138 4.19 4.70 2.47 2.83 0.13 0.18 0.07

Bazhenov Formation, Western Siberia, Russia


3784 2.44 0.187 3.34 4.45 2.18 2.80 0.39 0.32 0.14
3787 2.32 0.209 3.11 4.46 1.97 2.75 0.53 0.48 0.10
3788 2.34 0.171 3.28 4.41 2.10 2.76 0.40 0.36 0.23
3822 2.57 0.082 3.82 4.90 2.56 3.08 0.32 0.22 0.24
3824 2.43 0.143 3.41 4.79 2.28 2.99 0.49 0.36 0.16
3834 2.49 0.081 3.60 4.84 2.48 3.08 0.40 0.27 0.24
3842 2.46 0.110 3.69 4.78 2.53 3.08 0.34 0.24 0.22

06-Vernik_Ch06.indd 142 12-08-2016 20:00:09


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 143

Table 3.  Continued.


Depth ρb Kerogen VP(0°) VP(90°) VS(0°) VS(90°) ε γ δ
(m) (g/cm3) (v/v) (km/s) (km/s) (km/s) (km/s)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Monterey Shale, California, USA


0 2.15 0.203 3.62 4.29 2.27 2.44 0.20 0.08 0.06
0 1.56 0.291 2.58 2.90 1.69 1.78 0.13 0.06 0.00
0 1.44 0.280 2.40 2.67 1.51 1.62 0.12 0.08 0.12
0 1.56 0.048 2.68 3.22 1.79 2.05 0.22 0.16 0.05
0 2.23 0.030 4.44 4.67 2.66 2.95 0.05 0.11 0.02
0 1.66 0.087 2.48 2.90 1.69 1.88 0.18 0.12 0.24
0 2.56 0.051 4.52 5.09 2.86 2.96 0.13 0.04 0.06
1390 1.92 0.058 3.91 4.23 2.57 2.61 0.09 0.02 –
1393 1.93 0.055 3.76 3.93 2.49 2.63 0.05 0.04 0.02
1415 1.99 0.202 3.45 3.83 2.12 2.33 0.12 0.10 0.00
1428 1.86 0.190 2.68 3.28 1.72 1.94 0.25 0.14 –
1431 1.60 0.363 2.56 3.36 1.58 2.10 0.36 0.39 0.11
1454 2.02 0.147 3.01 3.64 1.90 2.14 0.23 0.13 0.06
1462 1.93 0.224 2.85 3.37 1.92 2.04 0.20 0.06 –
1670 2.14 0.182 3.70 4.12 2.43 2.65 0.12 0.10 0.01
1692 2.40 0.016 5.03 5.46 2.99 3.38 0.09 0.14 –

Kimmeridge Shale, North Sea


2133 2.26 0.057 3.26 3.89 2.10 2.37 0.21 0.14 –
4449 2.51 0.135 3.86 4.70 2.22 2.66 0.24 0.22 0.08
4412 2.42 0.114 3.20 4.02 2.00 2.38 0.29 0.21 0.19
0 2.65 – 3.71 4.60 1.97 2.67 0.27 0.42 0.14
3993 2.32 0.203 3.41 4.16 1.92 2.41 0.24 0.29 0.12
3996 2.35 0.153 3.55 4.32 1.98 2.48 0.24 0.29 0.08

Niobrara Formation, Colorado, USA


1121 2.62 0.032 4.55 5.39 2.76 3.17 0.20 0.16 0.12
1159 2.57 0.053 4.31 5.02 2.50 2.94 0.18 0.19 0.10
2248 2.50 0.057 4.06 4.83 2.36 2.91 0.21 0.26 0.08
2323 2.60 0.045 4.23 4.96 2.43 2.90 0.19 0.21 0.00
2350 2.55 0.014 4.40 5.17 2.56 3.02 0.19 0.20 0.10

Woodford Shale, Oklahoma, USA


2154 2.52 0.207 3.76 4.90 2.25 3.07 0.35 0.43 0.11

Lockatong Formation, New York, USA


553 2.64 0.017 4.79 5.49 2.76 3.28 0.16 0.21 0.04

Chicopee Shale, Pennsylvania, USA


0 2.72 0.003 5.25 5.65 3.13 3.28 0.08 0.04 0.03

(Continued )

06-Vernik_Ch06.indd 143 12-08-2016 20:00:10


144 Seismic Petrophysics in Quantitative Interpretation

Table 3.  Continued.


Depth ρb Kerogen VP(0°) VP(90°) VS(0°) VS(90°) ε γ δ
(m) (g/cm3) (v/v) (km/s) (km/s) (km/s) (km/s)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chattanooga Shale, Tennessee, USA


0 2.34 – 3.14 4.55 1.83 2.82 0.55 0.68 0.16

New Albany Shale, New York, USA


0 2.27 – 3.21 4.10 1.99 2.55 0.31 0.32 0.13
0 2.36 – 3.21 4.21 2.08 2.67 0.36 0.33 0.17
0 2.37 – 3.33 4.30 2.18 2.70 0.33 0.27 0.17

Antrim Shale, Illinois, USA


0 2.21 – 2.97 3.89 1.82 2.40 0.36 0.37 0.14

Cotton Valley Formation, Texas, USA


0 2.64 0.010 4.61 5.09 2.85 3.26 0.11 0.16 0.06

to widen the dynamic range of organic richness in the da- Notice that the P-wave NMO velocity for the Bazhenov
tabase. The ultrasonic velocity and anisotropy were mea- shale is very close to its bedding-normal velocity. This
sured on these rocks by using the same three-plug observation is easily explained by the relatively low value
­approach illustrated in Figure 20. Several high-quality of δ. It is also curious that the VNMO for SV-waves exceeds
data points from the Devonian organic shales found in the the VNMO for P-waves in this Bazhenov shale core. This
eastern provinces of the USA (the Chattanooga, New situation is readily explained by taking into account that
Albany, and Antrim Shales) are added from Johnston and this specific core is characterized by a high ε value of
Christensen (1995), although those data were obtained on 0.53, thereby giving rise to an anomalously high δSV value
whole core samples and are subject to more phase-­ of 1.068.
velocity/group-velocity uncertainty. Nonetheless, John­
ston and Christensen did carefully invert P- and S-wave Effect of kerogen on velocities and anisotropy
data to verify that their traveltime picks in nonprincipal
directions yield a P-wave velocity that is quite close to the Several publications on organic shales in recent years
phase velocity. Unfortunately, no information on organic have tended to sum the kerogen volume with that of clay
richness was reported in that study. Note that the kerogen in order to characterize their elastic and inelastic proper-
volume fractions given in Table 3 for the rest of the data- ties (e.g., Ortega, 2010; Prasad et al., 2011; Sone, 2012).
base are derived by using equation 3 from TOC values It is important to mention that such a summation is unde-
measured on colocated cores and included in Table 1. sirable and should be avoided because of the different ef-
Typical wave-velocity surfaces obtained from the tri- fects that the rock-forming constituents such as kerogen
directional phase-velocity data at 70 MPa on low-porosity and clay have on the mechanical properties of organic
Bakken and Bazhenov organic shales are shown in Figure mudrocks.
25. These are computed from equations 18 and display Table 4 presents a subset of the larger database and
the strong elastic anisotropy that is typical of the vast ma- comprises the intrinsic elastic properties measured at 50 to
jority of tight organic-shale cores (Vernik and Nur, 70 MPa on low-porosity mudrocks (φ < 0.1) that have a
1992a). The meaning of all four anisotropy parameters in clay content vcl, albeit rather simplistic value of 15% or
a TI medium, as well as that of the NMO velocities for greater on the solid-rock basis. Values for the TOC content
P- and SV-waves, becomes obvious when the elliptical and, hence, for the kerogen concentration of each sample
functions given by equations 20 and approximating the in this subset are also included. Nearly the same subset,
phase velocity surfaces at small angles are also added measured at 20 MPa, which is much closer to the vertical
onto Figure 25 as dashed lines. effective stress in situ, is listed in Table 5.
At 90° incidence those elliptical surfaces shown in The reasons for selecting the subset presented in Table
Figure 25 yield NMO velocities of P- and SV-waves inde- 4 are twofold: (1) to separate the effects of kerogen on the
pendent of whether phase velocity or group velocity (like velocities and anisotropies of shales from the effects of
in a seismic reflection survey) is measured (Figure 23). clay content and porosity, and (2) to distinguish organic

06-Vernik_Ch06.indd 144 12-08-2016 20:00:10


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 145

respective directions are plotted in Figure 26b. The four


data clouds (C11, C33, C44, and C66) display similar varia-
tions, thereby revealing that kerogen exerts a strong effect
on elastic-stiffness/rigidity reduction in both of the prin-
cipal directions of the shale texture — perpendicular and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

parallel to the bedding-controlled particle alignment. This


corroborates earlier observations that were based on a
more limited data set (Vernik and Landis, 1996; Vernik
and Liu, 1997).
The question now arises: How can we model this ef-
fect explicitly? Could the well-established anisotropic
Backus averaging scheme for thinly laminated shale be a
starting point? The generalized Backus equations for TI
formations with locally anisotropic layers are given by
equations 34 of Chapter 1. Inspired by limited SEM im-
ages of Bakken Formation shales with a high kerogen con-
tent of 25 to 30%, Vernik and Nur (1992a) used such an
approach by intercalating locally anisotropic inorganic
laminae (generated by the clay particles’ preferred orienta-
tion in the bedding plane) and locally isotropic organic
laminations. Both types of inhomogeneities show SEM-
determined thicknesses at least ten times smaller than the
dominant ultrasonic wavelength of 3 to 4 mm, so using the
effective-medium theory of Backus (1962) is justified.
Let us revisit this promising approach. The individual
mineral and kerogen isotropic parameters and the TI
properties of the clay mineral aggregates are given in
Table 6. The logic behind the clay aggregate parameter-
ization and the underlying assumptions are given below.
Because of uncertainties with the porosities, pore
shapes, and fluid saturations of kerogen, its elastic prop-
erties — especially for unrelaxed ultrasonic wave propa-
Figure 25.  P, SH, and SV phase-velocity surfaces (solid gation — are hard to separate from the effects of its nano-
lines) of (a) a typical Bakken shale core with ε = 0.29, pores. Therefore, the elastic parameters of kerogen should
γ = 0.42, δ = 0.18, and δSV = 0.32, and of (b) a Bazhenov be treated as the effective properties of porous organic
shale core with ε = 0.53, γ = 0.47, δ = 0.10, and δSV = 1.068. matter, at least in mature unconventional shales. It can be
Apparent anisotropy surfaces culminating at respective VNMO assumed that thermal maturation is unlikely to strongly
velocities are shown as dashed lines. Solid circles and open
influence the effective microstiffness of porous organic
circles represent actual measurement results for P-, SH-, and
matter, because solid kerogen undergoes densification si-
SV-wave velocities, respectively, at 70 MPa. Adapted from
multaneously with nanopore formation. It is noteworthy
Figure 4 of Vernik and Nur (1992a). Used by permission.
that the estimates of the P-wave and shear moduli of kero-
gen given in Table 6 are slightly greater than those re-
shales and marls from organic limestones and siliceous ported by Bandyopadhyay (2009), possibly emphasizing
mudstones, which have distinctly different mineralogic the uncertainties related to variations in kerogen porosity,
and textural characteristics (e.g., a weaker preferred orien- pore shapes, and saturating fluids.
tation of mineral components, due to calcite or silica The bedding-normal stiffnesses of the illite/mica clay
domination). aggregates (C33 and C44) are extrapolated from the Shale
Based on the definitions below equation 34 of Chapter Model (Vernik and Kachanov, 2010, and Chapter 3). A
1, the velocity and density data from Table 4 for organic strongly preferred orientation of clay platelets in these
mudrocks with vcl values greater than 15% were first con- zero-porosity clay aggregates is implied by the velocity-
verted to TI elastic stiffnesses and then plotted against anisotropy measurements on tight, conventional shales
kerogen content in Figure 26. Compressional stiffnesses with a clay content of 40% or less (e.g., the Cotton Valley
that are normal and parallel to the bedding plane are and Chicopee shales). Anisotropy-parameter values of
shown in Figure 26a, and shear rigidities in these ε = 0.30, γ = 0.25, and δ = 0.5ε are assumed for those

06-Vernik_Ch06.indd 145 12-08-2016 20:00:11


146 Seismic Petrophysics in Quantitative Interpretation

Table 4.  Subset of ultrasonic velocity and anisotropy measurements on mudrocks with clay content vcl values of 0.15 or
greater, at high confining pressures (50 to 70 MPa). Data are from Vernik and Liu (1997), Sondergeld et al. (2000), Lo et al.
(1986), Hornby (1998), and Tosaya (1982).
Formation/ Depth Maturity Ro (%) ρb TOC K (v/v) VP(0°) VS(0°) ε γ δ
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Lithology (m) (g/cm3) (wt%) (km/s) (km/s)


Bakken shale 2307 II 0.49 2.09 16.2 0.444 3.13 1.88 0.29 0.25 –
2632 II 0.54 2.06 15.9 0.415 3.02 1.76 0.44 0.51 0.10
2996 III 0.59 2.21 10.3 0.279 3.41 2.07 0.24 0.24 0.12
3098 III 0.61 2.22 12.1 0.326 3.38 2.12 0.27 0.23 0.10
3196 III 0.60 2.18 15.4 0.411 3.18 1.93 0.37 0.35 –
3199 III 0.59 2.30 9.8 0.277 3.62 2.22 0.23 0.20 –
3223 III 0.69 2.49 7.0 0.202 3.36 2.06 0.39 0.27 –
3271 IVa 0.80 2.33 9.7 0.249 3.46 2.00 0.34 0.36 0.14
3272 IVa 0.82 2.34 7.1 0.181 3.51 2.03 0.29 0.42 0.18
3332 IVa 1.13 2.35 9.5 0.217 3.29 1.86 0.51 0.59 0.30
3423 IVb 1.10 2.55 4.5 0.113 4.21 2.52 0.20 0.17 0.08
3428 IVb 1.19 2.38 9.8 0.222 3.72 2.25 0.19 0.16 0.10
3438 IVb 1.27 2.48 6.0 0.139 3.89 2.42 0.23 0.21 0.13
Bazhenov shale 3784 III–IVa 0.61 2.44 6.2 0.187 3.34 2.18 0.39 0.32 0.14
3787 III–IVa 0.64 2.32 7.4 0.209 3.11 1.97 0.53 0.47 0.10
3788 III–IVa 0.64 2.34 6.0 0.171 3.28 2.10 0.40 0.36 0.23
3822 III–IVa 0.78 2.57 2.8 0.082 3.82 2.56 0.32 0.22 0.24
3824 III–IVa 0.59 2.43 4.7 0.143 3.41 2.28 0.49 0.36 0.16
3834 III–IVa 0.72 2.49 2.8 0.081 3.60 2.48 0.40 0.27 0.24
3842 III–IVa 0.61 2.46 3.6 0.110 3.69 2.53 0.34 0.24 0.22
Niobrara marl 1121 IVb 0.88 2.62 1.2 0.034 4.55 2.76 0.20 0.16 0.12
1159 IVb 1.15 2.57 1.9 0.048 4.31 2.50 0.18 0.19 0.10
2248 V 1.15 2.50 2.1 0.052 4.06 2.36 0.21 0.26 0.08
2323 V 1.31 2.60 1.6 0.039 4.23 2.43 0.19 0.21 0.00
2350 V 1.36 2.55 0.5 0.012 4.40 2.56 0.19 0.20 0.10
Eagle Ford marl 3711 IVb 1.23 2.51 2.4 0.061 4.44 2.52 0.13 0.17 0.07
3713 IVb 1.26 2.47 3.5 0.086 4.06 2.40 0.17 0.18 0.08
3716 IVb 1.30 2.50 3.0 0.076 4.45 2.55 0.10 0.14 0.05
3721 IVb 1.28 2.40 5.1 0.125 4.11 2.47 0.15 0.14 0.07
3754 IVb 1.35 2.44 2.4 0.060 4.31 2.58 0.12 0.17 0.06
3754 IVb 1.34 2.45 4.2 0.104 4.18 2.47 0.15 0.20 0.07
3758 IVb 1.34 2.44 5.5 0.138 4.19 2.47 0.13 0.18 0.07
Kimmeridge shale 4449 IVb 1.10 2.51 5.0 0.123 3.86 2.22 0.24 0.22 0.08
4412 IVb 1.25 2.42 4.4 0.100 3.20 2.00 0.29 0.21 0.19
– – – 2.65 2.2 0.050 3.71 1.97 0.27 0.42 0.14
3993 IVa – 2.32 7.8 0.203 3.41 1.92 0.24 0.29 0.12
3996 IVa – 2.35 5.9 0.153 3.55 1.98 0.24 0.29 0.08
Woodford shale 2154 II 0.49 2.52 6.6 0.218 3.76 2.25 0.35 0.43 0.11
Lockatong shale 553 VI – 2.64 0.8 0.018 4.79 2.76 0.16 0.21 0.04
Chicopee shale 0 – – 2.72 0.0 0.000 5.25 3.13 0.08 0.04 0.03
Cotton Valley shale 0 – – 2.64 0.5 0.010 4.61 2.85 0.11 0.16 0.06

06-Vernik_Ch06.indd 146 12-08-2016 20:00:11


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 147

Table 5.  Subset of ultrasonic velocity and anisotropy measurements at 20 MPa on mudrocks
with clay content vcl ≥ 0.15. Sources: Vernik and Liu (1997), Sondergeld et al. (2000), Tosaya
(1982), and M. Aldin (personal communication, 2013).
Depth ρb VP(0°) VS(0°) ε γ δ
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

(m) (g/cm3) (km/s) (km/s)


Bakken shale
2307.3 2.09 3.00 1.84 0.36 0.28 –
2631.6 2.06 2.89 1.72 0.52 0.55 0.18
2996.5 2.21 3.31 2.02 0.28 0.27 0.17
3098.0 2.22 3.28 2.07 0.31 0.27 0.14
3196.4 2.18 3.05 1.88 0.44 0.38 –
3198.9 2.30 3.47 2.16 0.28 0.23 –
3271.4 2.33 3.32 1.95 0.39 0.38 0.20
3271.7 2.34 3.37 1.98 0.34 0.45 0.23
3331.8 2.35 2.96 1.75 0.71 0.70 0.46
3422.9 2.55 3.62 2.33 0.42 0.26 0.19
3427.8 2.38 3.20 2.08 0.40 0.26 0.23
3438.1 2.48 3.31 2.22 0.51 0.27 0.35
Bazhenov shale
3784.1 2.44 3.09 2.09 0.49 0.38 0.21
3787.1 2.32 2.88 1.89 0.65 0.54 0.20
3788.1 2.34 3.03 2.02 0.51 0.42 0.33
3821.9 2.57 3.53 2.46 0.42 0.27 0.34
3824.0 2.43 3.15 2.19 0.60 0.41 0.25
3834.1 2.49 3.33 2.38 0.51 0.32 0.34
3842.0 2.46 3.41 2.43 0.44 0.29 0.31
Niobrara marl
1121.4 2.62 4.19 2.64 0.29 0.21 0.19
1159.2 2.57 3.97 2.39 0.26 0.24 0.15
2248.2 2.50 3.74 2.25 0.30 0.32 0.13
2323.2 2.60 3.89 2.32 0.28 0.27 0.07
2349.7 2.55 4.05 2.44 0.28 0.25 0.17
North Sea shale
4448.9 2.51 3.57 2.13 0.33 0.26 0.13
4412.3 2.42 2.96 1.92 0.39 0.25 0.28
Woodford shale
1960.2 2.51 5.00 3.22 0.03 0.04 –
2154.3 2.52 3.62 2.17 0.40 0.49 0.16
2882.2 2.37 3.29 2.12 – – –
3570.4 2.40 3.58 2.33 – – –
3940.8 2.39 3.35 2.06 – – –

(Continued )

06-Vernik_Ch06.indd 147 12-08-2016 20:00:11


148 Seismic Petrophysics in Quantitative Interpretation

Table 5.  Continued.


Depth ρb VP(0°) VS(0°) ε γ δ
(m) (g/cm3) (km/s) (km/s)
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Eagle Ford marls


3711.3 2.51 4.25 2.45 0.22 0.26 0.13
3713.1 2.47 3.73 2.19 0.29 0.28 0.22
3715.9 2.50 4.02 2.38 0.17 0.22 0.12
3721.4 2.40 3.75 2.30 0.26 0.22 0.19
3754.1 2.44 3.92 2.38 0.21 0.27 0.15
3754.5 2.45 3.79 2.32 0.26 0.32 0.14
3757.7 2.44 3.81 2.25 0.22 0.28 0.14

Haynesville shale
3510.4 2.56 3.42 1.97 0.48 0.62 –
3525.0 2.50 3.42 2.10 0.45 0.44 –
3536.3 2.51 3.54 2.19 0.30 0.26 –
3542.2 2.59 4.13 2.53 0.12 0.08 –
3563.6 2.44 3.19 2.05 0.44 0.36 –
3567.6 2.58 3.66 2.37 0.34 0.23 –
3584.8 2.54 3.70 2.31 0.31 0.26 –
3596.1 2.58 3.54 2.16 0.48 0.49 –

Cotton Valley shale


– 2.64 4.36 2.50 0.14 0.18 –

Figure 26.  Intrinsic elastic stiffnesses computed from bedding-normal and bedding-parallel ultrasonic velocity measurements
in tight black shales at a high confining pressure of 70 MPa, versus the kerogen volume fraction for (a) normal stiffnesses and
(b) shear stiffnesses. Data are taken from the Table 4 subset (vcl > 0.15). Curves represent model predictions from Backus
(dashed lines for C11 and C66) and modified Backus models (solid lines). Note the outliers well below the model predictions,
which come from the Kimmeridge shales, North Sea, with vcl ≫ 0.25. Sample breakdown: upper/lower Bakken shale –13, Eagle
Ford marl –8, Bazhenov shale –7, Kimmeridge shale –5, Niobrara marl –5, Woodford shale –1, Cotton Valley shale –1,
Chicopee shale –1, and Lockatong shale –1.

06-Vernik_Ch06.indd 148 12-08-2016 20:00:12


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 149

Table 6.  Elastic properties of minerals, mineral aggregates, and kerogen (in GPa) used in modeling organic
mudrock measurements.
Phase M(C33) M(C11) G(C44) G(C66) λ (C13) Reference
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Quartz 95.6 95.6 44.0 44.0 7.6 Mavko et al. (2009)


Calcite 108.9 108.9 29.0 29.0 50.9 Mavko et al. (2009)
Feldspar 57.5 57.5 15.0 15.0 27.5 Mavko et al. (2009)
Clay-free, kerogen-free matrix 91.9 91.9 37.5 37.5 16.9 the Hill averaging method
Illite/mica aggregate 33.4 53.4** 8.5 12.7** 21.0 Vernik and Kachanov (2010)
Kerogen* 9.5 9.5 4.2 4.2 1.4 Vernik and Landis (1996)
*Effective properties of kerogen with oil-filled pores (φk = 0.3)
**Assuming that ε = 0.30, γ = 0.25, and δ = 0.15

clay-matrix aggregates, so that the other three anisotropic with data. This match signifies that the effect of inorganic
stiffnesses (C11, C66, and C13) can also be evaluated (Table 6). porosity on the elasticity of tight unconventional shales is
The local elastic properties of the inorganic phase can likely to be relatively small.
be estimated by using Hill averaging, a decision that is On the other hand, as kerogen content increases, the
supported by the SEM images indicating no segregation Backus model increasingly overpredicts the bedding-­
of the individual mineral components into microlamina- parallel stiffnesses. A heuristic modification of this model
tions. The primary mineral components to consider are for C11 and C66 was introduced by Vernik and Liu (1997)
quartz, calcite, and clay. Such an approach results in the in an attempt to explain the discrepancy. Again, inspired
inorganic phase being isotropic, which is dubious. by the SEM images showing a lenticular pattern of inor-
Alternatively and perhaps more accurately, Hill aver- ganic lenses separated by an anastomosing latticework of
aging of nonclay components can be combined with a kerogen (even at its moderate concentration) in both prin-
subsequent addition of clay, using the anisotropic Backus cipal textural directions, Vernik and Liu replace the bed-
model. From that approach, the following sets of elastic ding-parallel effective elastic moduli in equations 34 of
stiffnesses (in GPa) are obtained for the solid inorganic Chapter 1 with the following modifications denoted by an
phase with a vcl value of 25% and for that with a value of asterisk * below and in Figure 26:
50%. For vcl = 0.25, C11nk = 80.0, C33nk = 63.9, C44nk = 20.2,
−1
C66nk = 28.2, and C13nk =  27.3; and for vcl = 0.50, −1
*
c11 + C11
C11nk = 70.1, C33nk = 49.0, C44nk = 13.9, C66nk = 22.0, and C11 = ,
2
C13nk = 26.2. −1
(24)
−1
Figure 27 shows the effect that clay in the inorganic *
c66 + C66
phase of carbonate-poor unconventional shales exerts on C66 = .
2
its anisotropic elastic moduli. Note that all diagonal ele-
ments of the elastic stiffness tensor, such as C11, C33, C44, These modifications amount to the arithmetic average
and C66, decrease with clay content, as expected, whereas of Reuss’s composite stiffness (the first term in the nu-
C13 initially increases until it reaches a maximum, after merator) and Backus’s composite stiffness (the second
which it slowly decreases. Note also that the elastic mod- term). Replacing the bedding-parallel elastic moduli in
uli of the inorganic phase with vcl values of 25 and 50% are equations 34 of Chapter 1 with expressions 28 results in
expected to be close to those of the kerogen-free shales an acceptable fit to the experimental data (Figure 26) and
with respective clay contents and negligible “nonkerogen” reinforces the SEM observations of predominantly dis-
porosities (less than 4%), which typically is the case. continuous microlaminations of the inorganic phase in the
This locally anisotropic, inorganic phase can be intro- bedding plane that soften the bedding-parallel elastic
duced into the Backus model given by equations 34 of moduli of organic shales.
Chapter 1, together with the kerogen volume fraction K, It must be emphasized that the apparent continuity of
to compute the final model for the intrinsic elasticity of the kerogen latticework is inferred not just laterally but
organic mudrocks. In Figure 26, the intrinsic bedding- also vertically, although to a lesser degree. Otherwise, the
normal elastic stiffnesses (C33 and C44) from Table 4 are bedding-parallel elastic stiffnesses as a function of kero-
superposed by the Backus model with a vcl value of 25%, gen would be much higher than they are and more consis-
which is close to the average clay content in the subset. tent with the stiffness values of C11 and C66 computed
Obviously, the model lines are in good visual agreement from the true Backus model. More recently, Sayers (2013)

06-Vernik_Ch06.indd 149 12-08-2016 20:00:12


150 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 27.  Bedding-parallel (C11nk, C66nk) and bedding-


normal (C33nk, C44nk) elastic stiffnesses of the inorganic phase
of carbonate-poor organic shales, versus clay content, using
the modified Backus (1963) averaging scheme. C13nk was
computed assuming δ  = 0.5ε.

investigated the effects of kerogen distribution and con-


centration on the elastic properties of organic shales by
using the anisotropic-­effective-field theory of Sevostianov
et al. (2005), and he confirmed the earlier conclusion by
Vernik and Liu on the existence of a pervasive latticework
of kerogen that tends to separate the inorganic-phase lens-
es at least in organic-rich shales.
The presence of that kerogen latticework, appar-
ently interconnected in 3D in organic-rich shales, is
hard to ascertain merely with thin-section (Figure 6) or
even SEM images (e.g., Figures 8 and 9), so it is ex-
tremely valuable to obtain an independent confirmation
Figure 28.  Bedding-normal and bedding-parallel ultrasonic
of this microstructural feature by using ultrasonic data
velocity measurements in tight black shales at a high
analysis and m ­ odeling. First, we now understand that confining pressure of 70 MPa, versus kerogen content, for
hydrocarbons generated by kerogen maturation can be (a) VP and (b) VS. Data are from Table 4. Model curves are
interconnected in the rock matrix, which is consistent selected to approximately represent the upper and lower
with the ideas of McAuliffe (1979), who proposed a bounds (vcl = 0.25 and 0.50, respectively) for the data set,
model for oil-phase diffusion through a three-dimen- as well as the separation line between bedding-parallel and
sional oil-wet kerogen network. The kerogen lattice- bedding-normal data points.
work also explains the substantial brittleness of these
lithologies; a detailed discussion of this important topic plot are designed to separate the data clouds rather than
is deferred to Chapter 7. to match them.
Another way to analyze the data set from Table 4 is With that in mind, the modified Backus model, as
presented in Figure 28, which plots the bedding-normal shown in Figure 28, is computed with a vcl value of 50%,
and bedding-parallel velocities, instead of the elastic whereas the true Backus model for bedding-parallel ve-
moduli, against kerogen content. Again, both the locities is realized for a vcl value of 25%. The latter pro-
Backus and modified Backus models are included in vides an upper bound to the data, and 0° and 90° velocity
the figure, realized here for two different clay contents models with a high clay content provide reasonable lower
(25% and 50%). Note that the models selected for this bounds to the respective data clouds. One observation

06-Vernik_Ch06.indd 150 12-08-2016 20:00:14


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 151

a) It should be emphasized that the theoretical models


C11 data that can be helpful in conventional rock physics are less
C33 data clearly defined and/or parameterized in terms of textural
or microstructural observations on organic mudrocks.
Even the well-established and robust Biot-Gassmann the-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ory, including its anisotropic extension (Brown and


Korringa, 1975), may not be applicable to these complex
rocks. Furthermore, various effective-media schemes are
not always adequate even in conventional applications
(see Chapter 3). Consequently, one should always keep in
mind that under such circumstances, heuristic, curve-­
*
C11 fitting models may be unavoidable.
*
C33
By using the model predictions shown in Figures 26,
28, and 29, it is easy to calculate the elastic anisotropy
parameter ε and plot it as a function of kerogen content K
for the tight organic shales from Table 4 (Figure 30a).
Both the moderate-clay-content shales (vcl = 0.25) and the
higher-clay-content shales (vcl = 0.50) demonstrate that
b) the modified Backus model’s anisotropy values increase
C44 data with kerogen content and reach a maximum value of ε of
C66 data approximately 0.35 to 0.40 at a K value of 35 to 45% (a
TOC content of 15 to 20%). Because the higher-clay-con-
tent tight shales are quite anisotropic even at low kerogen
concentrations, their rate of anisotropy increase with or-
ganic richness is slower than that in moderate-clay-content
shales. As also is shown in Figure 30a, the true Backus
model once again provides an upper bound on the values
of ε measured on organic mudrocks under high stress.
A subset of data points from the large database in Table
*
C66 3 and corresponding to the group of much-less-­compacted
organic mudrocks from the Monterey Formation,
*
C44
California, USA, is plotted in Figure 30b and overlain with
the same model predictions for the tight shale that were
used in Figure 30a. Clearly, the undercompacted organic
mudrocks are characterized by lower anisotropy values
compared with those from their tight analogs at similar
Figure 29.  Intrinsic elastic stiffnesses versus kerogen kerogen concentrations. Despite the wide range of thermal-
content in unconventional shales with a low carbonate maturity stages sampled in the subset used in Figures 26
content, showing modified Backus model realizations for and 28 through 30 (Table 4), thermal maturity is not found
vcl = 0.25 (solid lines) and vcl = 0.50 (dashed lines). Data are to impact the intrinsic velocities and elastic anisotropies of
from Table 4, but with the Kimmeridge and Eagle Ford data tight organic shales. That evidence suggests that the rock
filtered out. Adapted from Figure 6 of Vernik and Landis textures formed during the process of strong mechanical
(1996). Reprinted with permission from AAPG, whose compaction, prior to thermal maturation.
permission is required for further use. Intrinsic anisotropy parameters of organic mudrocks
for the entire database at high stresses are crossplotted in
Figure 31 (Table 3). Two important observations can be
from Figures 26 and 28 is that some organic mudrock li- made: (1) parameter δ is always smaller than ε, and the in-
thologies, such as carbonate-rich Eagle Ford marls and equality 0.75ε > δ  > 0.15ε identifies a plausible range for δ
clay-rich Kimmeridge shales, significantly add to the data as a function of ε, and (2) parameter γ can be greater than,
scatter of elastic moduli at approximately the same kero- equal to, or smaller than ε , but overall the two are quite
gen content. If these lithologies are filtered out, as is the well correlated with the linear regression line near the ε = γ
case in Figure 29, the data scatter is reduced and the mod- line. As will be discussed later, the aforementioned bounds
ified Backus model for the specified clay-content range on the anisotropy parameter δ can be shifted toward slight-
matches the data quite accurately. ly greater fractions of the anisotropy parameter ε at the

06-Vernik_Ch06.indd 151 12-08-2016 20:00:15


152 Seismic Petrophysics in Quantitative Interpretation

a) Low-porosity shale data a)


Modified Backus model, vcl = 0.25
Backus model, vcl = 0.25
Modified Backus model, vcl = 0.50
Backus model, vcl = 0.50
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

b)
b)

Figure 30.  Intrinsic anisotropy parameter ε at 70 MPa as a


function of kerogen content in (a) tight unconventional shales Figure 31.  Intrinsic anisotropy parameters (a) ε versus δ and
and (b) high-porosity organic mudrocks of the Monterey (b) ε versus γ for the entire database of organic mudrocks, at
Formation, California, USA, overlain by the modified Backus a high confining pressure of 70 MPa. Linear regression
model predictions for low-porosity shales. (dotted black line) with a coefficient of correlation R = 0.902
is compared in (b) with the ε = γ line (red).
lower effective stresses that are typical of in situ conditions
in many overpressured unconventional shale plays.
As Kaarsberg (1959) suggested, shale anisotropy is texture that is subparallel to the bedding (Johnston, 1987).
strongly related to the preferred orientation of clay parti- However, this kind of lenticular arrangement implies that
cles. Apparently, the effect of kerogen on velocity anisot- kerogen lenses and strips are effectively separated by the
ropy is also related to the texture of unconventional mu- interconnected inorganic phase of these rocks. That as-
drocks — notably, to their compaction and associated flat- sumption is not confirmed either by the high-resolution
tening of clay particles and aggregates. An approach to SEM images or by the ultrasonic data modeling presented
separate the effects of compaction, kerogen, and clay min- here, both of which instead show that the kerogen is re-
erals on velocity anisotropy will be presented in Chapter 7. sponsible for creating the interconnected latticework that is
The process of compaction has been previously as- disrupting the continuity of the inorganic constituents in
sumed to have resulted in vertical deformation with lateral both principal directions. This is the case even in tight mu-
extension of the originally soft organic-matter microlayers drocks that have a moderate organic-matter content
and thereby to have created a finely laminated lenticular (3% < TOC < 5%), such as organic marls from the Eagle

06-Vernik_Ch06.indd 152 12-08-2016 20:00:18


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 153

Ford formation. Incidentally, the latter are characterized by amounts to the effective stress σ, has historically been the
the intrinsic anisotropy parameter values similar to the con- most valuable experimental rock-physics activity. Such
ventional, medium-porosity Gulf of Mexico shale men- experiments not only reveal rock elasticity to be a func-
tioned in Chapter 5, even though those marls have no more tion of stress (ultimately reaching and exceeding in situ
than 1/3 of the clay content found in the conventional shale. conditions), they also provide critical information about
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This is a good example of the relative significance of clay the rock’s microstructure, which is difficult to evaluate
and kerogen on velocity anisotropy of mudrocks. even with the most sophisticated petrographic means. In
In their paper on the seismic anisotropy of shales, the majority of field applications of elastic waves to sub-
Johnston and Christensen (1995) attribute the high anisot- surface characterization, such as reflection seismology
ropy values they measured to a preferred clay-platelet and sonic logging (but with the possible exception of the
alignment in the bedding plane. It should be noted, how- 4D techniques), one simply does not have the luxury of
ever, that their shale cores were almost exclusively taken obtaining velocity data at different stress levels in situ.
from organic shales that are typical of the known source- Although they are quite consolidated, tight uncon-
rock formations of North America, so a subset of their ventional shales are rather brittle rocks that easily split
experiments is actually included in our database (Table 3). along planes of weakness. In consolidated rocks, core
It can be argued that in addition to the clay alignment, the measurements of elastic properties, whether static or dy-
kerogen latticework is primarily responsible for the con- namic, typically encounter a problem of intergranular
sistently high anisotropy that Johnston and Christensen and intragranular microcracks that do not exist in situ but
(1995) measured in their samples. are introduced during core recovery as a result of stress
Vasin et al. (2013) employ several effective-medium relief and cooling (Vernik et al., 1994; Holt et al., 2005).
theories, X-ray-diffraction microtomography-based evi- As is demonstrated in the work of Vernik at al. (1994)
dence for preferred orientation of clay particles, and single- on crystalline rocks from one of the world’s deepest verti-
crystal anisotropic elastic constants of illite (Militzer et al., cal wells (the Kola scientific borehole), stress relief and
2011), to predict the five elastic stiffnesses of a Kimmeridge cooling effects increase with depth. However, the impact
shale sample (that was tested by Hornby, 1998, and includ- of those factors on laboratory measurements of ultrasonic
ed in Table 3) with a high clay-weight fraction (wcl = 0.63). velocities varies as a function of lithology. The most dra-
Vasin et al. were able to match the experimental data fairly matic manifestation of core-recovery-related microcracks
closely by using the assumption of 10% porosity uniformly was found in quartz-rich rocks. Quartz is chara­cterized by
represented by spheroids with aspect ratios α = 0.01, which an anomalously high, and also highly anisotropic, ther-
are more typical for microcracks than are the shale pores mal-expansion ­coefficient, so the quartz-rich gneisses re-
discussed above. A simple c­ omputation of the crack-density covered from depths greater than 9000 m, when measured
parameter η for this ­numerical model yields an excessively on room-dry core plugs at ambient conditions in the labo-
high value — 2.4 (see Chapter 3) — that is totally inconsis- ratory, show nearly a 50% reduction in P-wave velocity
tent with the consolidated nature of tight organic shales compared with no more than a 30% reduction in quartz-
such as the Kimmeridge Shale. Again, by taking into ac- poor amphibolites, relative to the sonic-log velocity in
count the effects of kerogen, those authors could have these rocks. The nonlinear stress sensitivity of ultrasonic
avoided introducing such an ill-constrained numerical velocities measured on crystalline rocks recovered from
model and could have matched the experimental data in a great depth is dramatic and is related to the introduction
much more straightforward manner. of open m­ icrocracks, which are obviously absent in situ
To summarize, we can postulate that TI (polar an- according to sonic-log and VSP-interval-velocity values.
isotropy) typically found in most shales is related to Differentiating the effects of natural cracks from
their mechanical compaction. However, because com- core-recovery-induced ones is always a challenge, al-
paction affects both conventional and unconventional though shales and carbonates are less subject to cooling
shales, the reason for greater anisotropy observed in the effects than quartz-rich sandstones are. A more detailed
latter is the presence of aligned solid organic matter discussion of the issue of microcrack generation in un-
(kerogen) in addition to clay platelets with preferred ori- conventional mudrocks is provided in Chapter 7.
entation. Consequently, the main agents of polar anisot- Vernik and Nur (1992a) and Vernik and Liu (1997)
ropy in shales in general are clay and TOC contents. measured P- and S-wave velocities as a function of stress
in room-dry organic mudrocks, and their results show the
Stress dependence of velocity and anisotropy following patterns. (1) Tight shales that are subjected to
stage IV thermal maturity (0.75 to 1.30% Ro), and notably
Conducting traveltime measurements in rock samples to stage IVb maturity (0.90 to 1.30% Ro), are character-
under gradually increasing or decreasing confining pres- ized by a high-gradient nonlinear increase in the bedding-
sures, which in hydrostatic and drained conditions normal velocity at low stress, followed by a gradual

06-Vernik_Ch06.indd 153 12-08-2016 20:00:18


154 Seismic Petrophysics in Quantitative Interpretation

flattening of the velocity curves at stresses greater than 30 maturation (Ro = 0.5 to 0.75%). Two explanations may
to 50 MPa. This important observation is captured in apply to this observation: (1) kerogen-to-oil transformation-­
Figure 32. (2) Tight shales that are subjected to maturity induced pore-pressure buildup reaches the maximum dur-
stages II (immature) and III (early mature) typically do ing the latest stages in the oil window, and (2) according to
not display a high-gradient increase in bedding-normal Price’s (1994) hypothesis, a significant amount of oil is lost
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

velocities, even at low pressures (Figure 32). (3) Bedding- during core retrieval of stage IV shales, due to the hydro-
parallel P-wave velocities of tight shales of any maturity carbon-gas-phase volume expansion, and that implies a
stage always show a low stress sensitivity (Figure 33) higher oil saturation in situ. In any case, the core-recovery-
(Vernik and Liu, 1997; Vanorio et al., 2008). related origin of the bedding-subparallel microcracks de-
Figure 33 exemplifies the laboratory data acquired on tected in some organic shale cores can be safely ruled out.
the Upper Bakken organic shale core that belongs to stage Further evidence of their presence in situ will be discussed
IVb maturity (the final stage of oil generation). The differ- later in this chapter and in Chapters 7 and 8. It is also inter-
ence between the gradient of the increase in the bedding- esting to note that immature shales, and some postmature
normal P-wave-velocity with stress (dVP(0°)/dσ) at low (with respect to oil) shales, have a reduced sensitivity to
stresses and the gradient in the bedding-parallel direction stress, especially in normal-pore-pressure environments.
can be attributed to the bedding-plane low-aspect-ratio mi- Figure 34 presents the general trend of decreasing an-
crocracks present in the tight organic shales of certain matu- isotropy with increasing stress — a trend caused by stress-
rity levels. Note also that, as is typical of TI media with an- induced microcrack closures in the highly anisotropic
elliptic P-wave anisotropy, the SH-wave velocity at 45° bi- Upper Bakken shale characterized by thermal maturity
sects the measurements in the principal directions, whereas stage IVb. Note that at stress levels greater than 50 MPa,
the P-wave velocity does not. Instead, VP(45°) is found to the anisotropy values stabilize by gradually reaching the
gradually migrate closer to the bedding-normal velocity asymptotes that define the intrinsic anisotropy. Because
with increasing stress. The Bakken shale sample displayed the most productive unconventional shale reservoirs (the
in Figure 33 is somewhat extreme in the tendency of its Bakken, Eagle Ford, Woodford, Bazhenov, and the like)
VP(45°) to migrate toward the bedding-parallel velocity at are overpressured, they normally are subjected to effec-
lower stresses; we shall revisit this set of measurements tive vertical stresses that vary from 14 to 25 MPa, mean-
later. Here, it is sufficient to mention that the majority of ing that many of the maturation-induced microcracks may
other cores in the database show a less noticeable drift. be open and may enhance the polar anisotropy in situ.
Vernik and Landis (1996) found an offset between the The measured stress dependence shown in Figure 34
peak microcrack generation (stage IVb) and the peak bitu- highlights an anomalous behavior briefly alluded to in
men saturation, which, on the basis of core pyrolysis re- Figure 33: The ε and δ curves cross each other at the inter-
sults, takes place earlier during stage III of thermal mediate effective-stress level. This crossover is directly

Figure 32.  Bedding-normal P-wave velocity versus stress Figure 33.  P-wave (red) and SH-wave (green) velocity
(confining pressure) curves (normalized to VP at 70 MPa), with measurements on three plugs at different orientations to the
the breakdown according to the main stages of thermal maturity symmetry axis of the upper Bakken shale core (from 3438 m)
(stages II to VI). Data are from Vernik and Landis (1996). from maturity stage IVb (Table 2).

06-Vernik_Ch06.indd 154 12-08-2016 20:00:19


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 155

in TI media. Vernik and Milovac (2011) considered the


bedding-normal elastic stiffnesses of one of the most stress-
sensitive shale cores in the compiled database (Figure 33)
and were able to match them, with reasonable accuracy, by
using the noninteraction-approximation (NIA) model ap-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

plied to the TI background with bedding-parallel cracks


(Sevostianov and Kachanov, 2001). That model is formu-
lated in terms of the bedding-normal engineering parame-
ters (the Young’s modulus E3 and shear modulus G31 = G13
described in detail in Chapter 7) and can be extended to the
full set of elastic constants by introducing the second-rank
crack-density tensor to account for microcrack distribution
with a preferred orientation parallel to the bedding plane.
Because the vast majority of low-aspect-ratio cracks
(α < 10−2) are closed under the high confining stress of
70 MPa, and because of the low total porosity of even ma-
ture organic shales, the intrinsic anisotropic background
Figure 34.  Anisotropy parameters ε and δ, measured at a
stiffnesses can be used as a starting model. Then, the best-
confining stress that varies from 5 to 70 MPa in an organic fit values of the isotropic-equivalent Poisson’s ratio νiso and
shale that is at maturity level IVb. the moduli Eiso and Giso can be calculated (Sevostianov and
Kachanov, 2008) for this microcrack-free background,
related to the visible shift of VP(45°) with increasing stress thereby making it virtually isotropic.
from the bedding-parallel to the bedding-normal velocity That step is followed by application of the NIA model
values. The crossover seen for this Upper Bakken core with the crack-density tensor η. The tensor is characterized
sample is quite anomalous for the database, which shows by three principal components: η11, η22, and η33, the first
that even at low stresses the rule still holds that δ < ε, al- two of which are related to the bedding-parallel orientation
though the relative difference between the two values may and are equal in TI media.
be slightly reduced (refer to Figure 31a). Let us postpone The model can be further simplified if we choose to
this important discussion until a theoretical modeling of ignore the difference between the best-fit isotropic and
stress dependence is carried out. bedding-normal Young’s and shear moduli for a crack-free
To summarize, results of angle-dependent ultrasonic background, so that the elastic parameters can be written as
measurements on tight organic shales at different stresses
show that at stage IV thermal maturity (Ro values of 0.75 to E 3 int
E3 = ,
1.3%) these rocks exhibit a stress dependence that is typi- 1 + A(ν iso )η33
cal of low-porosity rocks with bedding-plane-­controlled E1 int
microcracks. Microcrack closure caused by increasing E1 = ,
1 + A(ν iso )η11
confining stress produces the high-gradient slope of the
2
VP(0°) and anisotropy-parameter curves seen in Figures 32 32(1 − ν isoo)
A(ν iso ) = ,
through 34, at low stresses. The relatively low in situ effec- 3( 2 − ν iso )
tive stress that is acting in many unconventional shales al- G13 int
lows for the majority of higher-aspect-ratio microcracks to G13 = ,
1 + B (ν iso )(η11 + η33 )
be open at depth. A continuous, anisotropic kerogen net-
work combined with spatially associated texture-controlled 16(1 − ν iso )
B (ν iso ) = , (25)
microcracks should provide an effective drainage system 3( 2 − ν iso )
that hydrocarbons can use for short-range charging of spa- G12 int
tially adjacent natural or hydraulic fractures. G12 = ,
1 + B ( viso )(η11 + η22 )
v13 int
Modeling stress dependence v13 = E1 ,
E 1 in t
Available theoretical results can reinforce the straight- E1
forward but qualitative interpretation of oriented, texture- v12 = − 1,
2 G12
controlled microcracks in certain mature organic shales to
E3
help us model, in a quantitative way, the effect of those v31 = v13 ,
microcracks on elastic stiffnesses and velocity anisotropy E1

06-Vernik_Ch06.indd 155 12-08-2016 20:00:20


156 Seismic Petrophysics in Quantitative Interpretation

with the subscript “int” meaning “intrinsic.” Note that the a)


ratio ν13/E1 in this model equals that of the TI background
medium — that is, the bedding-plane microcracks have
no added effect on the bedding-normal strain when they
are compressed parallel to bedding. This allows computa-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tion of all five of the independent elastic stiffnesses nec-


essary to fully describe the crack-induced anisotropy
enhancement.
This simple model is only reasonably accurate if the
amount of anisotropy of the background material is limit-
ed to E1int/E3int ≤ 1.5, which is largely the case for the high-
stress data set of Table 4. Because for the average shale in
the data set the assumption that Eiso = E3int is only accurate
to within approximately 15 to 20%, it should be empha-
sized that a crack density inverted using equations 25 re-
tains that uncertainty as a cost for model simplification.
Note also that the A(νiso) and B(νiso) functions in equations
25 are approximately equal to 5.70 ± 0.02 and 2.30 ± 0.07,
respectively, so they are quite stable. That stability is pro- b)
vided by the relatively tight distribution of intrinsic, best-
fit isotropic-equivalent Poisson’s ratios in the data set.
Note that the subscripts of the Poisson’s ratios ν and
shear moduli G specify the active (first subscript) and re-
active (second subscript) deformation directions, whereas
orientation 3 always denotes the TI symmetry axis. By
using the relationships among TI stiffness tensor elements
(e.g., Fedorov, 1968), we can calculate all five indepen-
dent elastic moduli and, hence, the velocities:

E1 (1 − ν13ν 31 )
C11 = ,
(1 + ν12 )(1 − ν12 − 2ν13ν 31 )
E3 (1 − ν12 )
C 33 = ,
(1 − ν12 − 2ν13ν 31 )
(26)
C 44 = G13 ,
Figure 35.  NIA models of the anisotropic (a) Young’s
E1
C 66 = G12 = , modulus and (b) shear modulus, optimized to match data
2(1 + ν12 ) (filled circles) on the Bakken shale core (3438 m) and shown
C13 = 2ν 31 (C11 − C 66 ). in Figure 33. Adapted from Figure 2 of Vernik and Milovac

(2011). Used by permission.
In the framework of the noninteraction approxima-
tion (see NIA in Chapter 3), elastic compliances are pro- work reasonably well in a variety of unconventional mu-
portional to the components of the crack-density tensor. drocks tested in laboratory experiments.
Progressive crack closures with increasing stress gradu- Figure 35 shows the results of the NIA predictions for
ally change the crack-density components of the η-tensor, the Bakken shale core from 3438 m, which was noted ex-
and each component is allowed to decay exponentially ac- perimentally to have (1) velocity and anisotropy values
cording to the formula: that are typical of mature, high-TOC shales, but (2) an
anomalous ε-versus-δ relationship with stress (Figure
η = η0 exp( − dσ ), (27) 34). A reasonable match to the experimental data for
anisotropic Young’s and shear moduli is achieved for this
where η0 is the zero-stress crack-density tensor and d is a organic shale by using η11 = η22 = 0.02 and η33 = 0.19
0 0 0
function of the crack aspect ratio controlling the crack- (Figure 36). That supports the earlier interpretation of a
closure stress. Khadeeva and Vernik (2014) have empiri- strongly preferred crack orientation in the bedding (sym-
cally demonstrated that the value for d of 0.06 seems to metry) plane.

06-Vernik_Ch06.indd 156 12-08-2016 20:00:21


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 157
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 36.  NIA-model-based P- and S-wave velocity Figure 37.  Measured and NIA-model-based anisotropy
predictions for measurements from the Bakken shale parameters ε versus δ for the Bakken shale core (from
core (from 3438 m) (red and green filled circles). The best- 3438 m), as a function of stress.
fitting crack-density-tensor components are η33 = 0.19 and
η11 = η22 = 0.02.
the velocity-measurement-based ones in Figure 37. The
results demonstrate that the crossover of ε and δ values
It is important to realize that the zero-stress bedding- displayed in Figure 34 is unlikely to be real but is, instead,
normal crack-density tensor component η33 value, which an artifact of the exaggerated δ. No crossover is seen in
0
is needed to explain why, even after a +20% correction, the model predictions for this seemingly anomalous core,
the dramatic bedding-normal stress sensitivity of the even though the value of δ does change with stress —
Bakken shale sample in question is still smaller than 0.25 from 0.54ε at 70 MPa to 0.80ε at 5 MPa.
— which is a high value for the crack-density parameter. The stress dependence of the crack-density tensor
Furthermore, in the retrospect, this value of 0.25 is still given by equation 27 is depicted in Figure 38, which
an order of magnitude smaller than the totally unreason- shows the bedding-normal and bedding-parallel tensor
able value implied by the differential-effective-medium components η33 and η11 for the core sample of Bakken
(DEM) model of Vasin et al. (2013) for the Kimmeridge shale discussed above. Both of those crack-density
Shale core measured by Hornby (1998) at a very high components exponentially reduce to negligible values
stress of 80 MPa, and reported above. It is also important at 70 MPa, at which point the crack-free organic shale
to realize that in our database the majority of organic with intrinsic velocities and anisotropy is revealed. The
shales that are at a thermal maturity other than stage IV η33 values at low stress, inverted from the NIA model-
can be modeled by using an even lower value of η33 0
ing for this sample, are among the highest for our data-
(typically below 0.15) for their stress sensitivity. base. Crack-density values in excess of 0.6, let alone
Again, because the vertical effective stress σ for the greater than 1.0, obviously are not realistic even at a
location of the modeled Bakken shale core is estimated low enough effective stress of 5 to 10 MPa to represent
to be approximately 23 MPa, the vast majority of the consolidated rocks such as unconventional shale
maturation-induced cracks inferred from the laboratory formations.
experiments may actually be open in situ, and their ef- Because crack density is strongly affected by the
fects can be detectable by sonic logs. Even more in- magnitude of effective stress, it is instructive to compute
triguing is the fact that our stress-dependence modeling the velocity surfaces for different stress levels. This is
implies that the VP(45°) value is erroneous at low stress- implemented in Figure 39, which shows the modeled
es in this specific core sample (Figure 36), whereas phase velocities of the same Bakken core, here at low,
the model’s predictions (equations 25, 26, and 27) intermediate (close to in situ), and high stress levels.
closely match all other velocity measurements from this This graph helps us to better appreciate the influence of
sample. predominantly subhorizontal microcracks on the P- and
The model-based anisotropy parameters computed SV-wave velocities and, especially, on the P-wave an-
for the stress range in the experiment are compared with isotropy changes with stress.

06-Vernik_Ch06.indd 157 12-08-2016 20:00:22


158 Seismic Petrophysics in Quantitative Interpretation

can we predict and interpret sonic logs that typically are


acquired in vertical pilot wells in unconventional plays?
For this purpose, Khadeeva and Vernik (2014) adopted
the NIA model, described in Chapter 3 and originally re-
fined and employed by Vernik and Kachanov (2010) for
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

modeling consolidated sedimentary rocks with realistic


pore geometries and microcracks. The Vernik-Kachanov
version of the NIA model takes into account the pore-shape
factor and the crack-density parameter, where the latter is
set to decay exponentially with stress.
When adopted for the bedding-normal stiffnesses in
organic mudrocks, this model contains, by necessity, the
shape factors of inorganic pores, which can be empirically
adjusted. The amount of intrinsic anisotropy must be lim-
ited to ε < 0.25 (E1int/E3int ≤ 1.5) for the model to be reason-
ably accurate, and the exponential decrease in the crack-
density parameter is assumed to be due to gradual crack
closure with stress.
Figure 38.  Crack-density-tensor components as functions of In principle, the NIA model is applicable only for
effective stress, inverted from matching the NIA model to the very-low-porosity rocks (Chapter 3) and must be replaced
measurements on the Bakken shale core (from 3438 m). Note with the Mori-Tanaka model for nonnegligible total po-
that the values η11 = η22 = 0.02 and η33 = 0.19 indicate a strong, rosities. Such an adjustment is not necessary for tight or-
preferred alignment of microcracks along the bedding plane. ganic shales. Again, when adopted for TI rocks, such as
organic shales, the model can be formulated exactly, in
terms of the anisotropic Young’s and shear moduli (as in
equations 25). However, a heuristic simplification of the
model for bedding-normal stiffnesses can be expressed as

C33m
C33d = ,
1 + pφnk′ + A (ν iso ) η33 exp (− d σ )
0
(28)
C44 m
C44 d = ,
1 + qφnk′ + B (ν iso ) (η11 + η33 ) exp(−dσ )
0 0

where C33d and C44d are the room-dry equivalent stiffnesses


of the organic shale with hydrocarbon-filled kerogen pores
and partially dry inorganic pores, C33m and C44m are the ma-
trix stiffnesses incorporating the effective organic phase of
the shale (solid kerogen with pores), p and q are the inor-
ganic pore-shape factors, and φnk′ is the contribution from
inorganic pores to the total porosity of the rock. The coef-
ficients A(νiso) and B(νiso) are quite close to those given in
equations 25 and can be taken to be equal to 5.7 and 2.3,
Figure 39.  P-wave and SV-wave phase-velocity surfaces respectively, when the intrinsic, best-fit isotropic-equiva-
for different stress levels, computed on the basis of the
lent Poisson’s ratio ranges from 0.24 to 0.28, which is the
NIA model. Note the microcrack-induced spread of the
case for tight organic mudrocks under high stresses.
P-wave velocity with a decreasing angle.
As was mentioned above, this model implies that the
user cannot separate out the effect of hydrocarbon-filled
Rock-physics model kerogen nanopores on the overall elastic properties, so the
effective stiffnesses of porous kerogen (e.g., Table 6)
Now that we have gained insight into the intrinsic, must be used in computing the values for C33m and C44m.
p­rimarily composition-controlled elastic properties of Estimating these bedding-normal solid-matrix stiffness
low-porosity organic mudrocks and also into the stress- components presents a challenge, notably in the absence
dependent microcrack effects therein, the question is: How of data on mineral fractions from XRD.

06-Vernik_Ch06.indd 158 12-08-2016 20:00:24


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 159

One way to evaluate C33m and C44m is to use the averaging, the bedding-normal elastic stiffnesses of the
Backus model — that is, harmonic averaging of the ­inorganic phase are computed first, and the overall matrix
components (isotropic kerogen and anisotropic inorgan- stiffnesses C33m and C44m are derived by another batch of
ic phase Cijnk), performed in the same fashion as was Backus averaging of the inorganic phase with kerogen. In
discussed above in modeling the effect of kerogen on doing so, the variation in kerogen porosities is ignored
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

the intrinsic elastic moduli. A good starting point for and the constant kerogen stiffnesses from Table 6 are
the inorganic phase stiffnesses is the Hill model, or the introduced.
Backus model, or a combination thereof, depending on The φnk′ porosity that enters equations 28 is not the same
the abundance of clay minerals and textural characteris- as that schematically depicted in Figure 12 and included in
tics of the mudrock. The zero-stress crack-density-­ equation 9, because here it constitutes the portion of the
tensor components η11 and η33 can be constrained by
0 0
total porosity represented by the inorganic-phase pores.
using available core measurements of velocities as Hence, it should be computed by using φ = Vkφk + (1 − Vk)
functions of the effective stress σ. φnk in combination with equation 5 as
Alternative micromechanical approaches have also
been applied in the modeling of organic shales. The ma- K(1 − φ )
φnk′ = φ − Vk φk = φ − φ. (29)
jority of those alternatives use the DEM scheme (e.g., 1 − φk k

Hornby et al., 1994; Draege et al., 2006; Vasin et al.,
2013). The DEM scheme is not explicitly formulated in Assuming a kerogen porosity of 40%, based on the
closed-form expressions and assumes that hypothetical log-model runs for the Eagle Ford marl (e.g., Figure 17),
ellipsoidal inclusions with a precisely known aspect-ratio the inorganic portion of the total porosity (of 10%) is 3%,
distribution are present and completely account for the meaning that the hydrocarbon-saturated porosity of the
total porosity of the rock. As was discussed in Chapter 3, marl is 7%. So, in this example the value that enters equa-
the assumption of ellipsoidal or penny-shaped pore geom- tions 28 for the Eagle Ford marl computations is 0.03.
etry makes these models quite controversial and may lead Figure 40 illustrates how limited core data can be used
to an inaccurate representation of the shale microstructure to calibrate the zero-stress crack-density-tensor compo-
and/or its microcrack-density parameter. It is also obvious nents η11 and η33 as well as the pore-shape factor in
0 0

from SEM observations that organic shales contain a va- equations 28. It follows from the best-fitting inversion that
riety of pore shapes, including quite isometric ones that p = 4.0, η11 = 0.095, and η33 = 0.13. Note that a similar
0 0
may slightly deform but never entirely close even under exercise carried out above for the Bakken shale (Figures
an effective stress as high as 70 MPa. 35 and 36) yields a much stronger preferred orientation of
An attractive feature of the model represented by microcracks compared with this Eagle Ford example. That
equations 28 is that it explicitly relates the bedding-nor- ­finding is consistent with (1) a generally less developed
mal moduli and, therefore, the P- and S-wave velocities oriented fabric in Eagle Ford marls relative to the other
measured in vertical pilot wells, to the mudrock composi- organic shales seen in SEM images (e.g., Figures 7b and
tion, the microstructure, and the vertical effective stress in 8) and (2) a lower intrinsic anisotropy of Eagle Ford marls
situ. Application of this model to Eagle Ford marls (from relative to that of the Bakken, the Bazhenov, and most of
Texas, USA) is discussed next. the Woodford shales due to lower clay and kerogen con-
Table 7 lists average compositional parameters, po- centration (Table 4). Note also that the crack-density pa-
rosities, effective stresses, and sonic-log velocities from rameter of 0.6 that Khadeeva and Vernik (2014) used to
four cored wells in vastly different geographic locations match the Eagle Ford core velocity measurements is much
for the Eagle Ford Formation. By using the constituent greater than the values mentioned above. Their high crack-
properties given in Table 6 and performing Hill/Backus density estimates were an artifact of underestimating the

Table 7.  Composition (from XRD and pyrolysis analyses converted to volume fractions), crushed
sample porosities φ, mean sonic velocities (in km/s), and effective stress σ values (MPa) from four wells
in the Eagle Ford Formation.
Well φ K Calcite vcl Quartz VP VS σ
#1 0.10 0.10 0.60 0.14 0.16 3.98 ± 0.32 2.26 ± 0.21 20.0
#2 0.10 0.10 0.63 0.13 0.14 3.95 ± 0.29 2.24 ± 0.19 20.0
#3 0.11 0.09 0.52 0.22 0.17 3.58 ± 0.22 2.07 ± 0.15 13.0
#4 0.11 0.10 0.43 0.30 0.20 3.29 ± 0.18 1.79 ± 0.14 13.0

06-Vernik_Ch06.indd 159 12-08-2016 20:00:24


160 Seismic Petrophysics in Quantitative Interpretation

5.0
Eagle Ford Fm.
νcl = 0.14 4 wells
4.5
0.22
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

4.0
0.30

VP (km/s)
3.5
σ = 20 MPa

3.0 σ = 13 MPa
Wells: 1-cyan
2-green
3-blue
2.5 4-red

2.0
0 2 4 6 8 10
TOC (%)

Figure 40.  Bedding-normal P- and S-wave velocities Figure 41. Vertical-sonic-velocity-versus-log-model-TOC


measured in the laboratory on an Eagle Ford marl and fitted values for four Eagle Ford wells. Carbonate beds and lenses
using the Vernik-Kachanov model for TI rock, with a are excluded. Three model realizations are superposed,
preferred orientation of microcracks along the bedding plane. matching composition and effective stress in Eagle Ford
marls. (1) vcl = 0.14 at 20 MPa, (2) vcl = 0.22 at 13 MPa,
and (3) vcl = 0.30 at 13 MPa (Table 7); pore-shape factors
coefficients A(νiso) and B(νiso) and ignoring the tensorial are p = 4, 5, and 11, for the green, blue, and red lines,
properties of η in their study. In fact, the proper way to respectively; total porosity is assumed to be constant at
calculate the aforementioned coefficients for nonrandom 0.1, φnk′ = 0.03, and η33 = 0.13.
crack distributions is given in equations 25.
0

Modeling the bedding-normal elastic stiffnesses by plot. Even though equations 28 consider a partially dry
using equations 28 and comparing the results with the well inorganic phase, the formulas can be used with ­adequate
data listed in Table 7 reveals that, if the crack-density pa- accuracy for log modeling under in situ conditions, be-
rameters are kept constant, the pore-shape factors p and q cause the effect of water saturation in low-­porosity mate-
must be quite different (q > p) and dependent on clay con- rial with predominantly stiff pore geometries is not ex-
tent. This clearly indicates that the dominant pore shapes pected to be significant. The bulk density necessary for
(not to be confused with low-aspect-ratio microcracks) in velocity computation is given by ρb = [(1 − K)ρnk + Kρk]
organic marls are not ellipsoidal (see Figure 3 of Chapter (1 − φ) + ρfφ, where ρf is the fluid density and porosity φ
3). To fit the log averages in Table 7, the clay-content de- is given by equation 13.
pendence of the shape factors ought to be nonlinear, but as The option to filter out thick carbonate beds, which
a general guideline the following two linear functions can are resolved by log data, is used for all four wells plotted.
be suggested: p ≈ 37.5vcl – 1.0, and q ≈ 100vcl – 6.0, where The two well pairs are characterized by substantially dif-
vcl is the clay content recalculated on a kerogen-free basis. ferent levels of overpressure and, given approximately the
As was discussed in Chapter 3, where the shape fac- same overburden gradient, of effective stress. The pore-
tors in arenite sandstones are shown to be porosity-depen- pressure gradient tested in wells 1 and 2 is 18.0 kPa/m
dent, the lithologic and textural controls on p and q in or- (0.78 psi/ft), and in wells 3 and 4 it is 20.3  kPa/m (0.88  psi/
ganic mudrocks seem to be natural. However, these very ft). In addition to the difference in the Eagle Ford forma-
uncertain empirical relationships for the shape factor in tion depths encountered between these two well pairs,
organic shales are not at all global (as their counterparts this pore-pressure mismatch gives rise to a stress differ-
might be in arenites) and should be adjusted for each spe- ence as great as 7 MPa for the effective vertical stress
cific case. Moreover, it should be noted that inaccuracies listed in Table 7.
in computing matrix moduli also affect the numerical val- Mineral composition, especially with respect to clay
ues of the shape factors in these complex lithologies. and calcite contents (mean vcl = 0.14 and vcalc = 0.62), of
The Vernik-Kachanov model adopted for the normal- organic marls – the dominant lithology in the Eagle Ford
bedding stiffnesses of organic shales in equations 28, Shale interval – in the first two wells shown in Table 7 is
with calibrated parameter values, is shown in Figure approximately the same. In combination with the same
41, a sonic-velocity-versus-log-model-based-TOC cross­- mean kerogen volume (K = 0.1), that mineralogy results

06-Vernik_Ch06.indd 160 12-08-2016 20:00:26


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 161

in approximately the same solid-matrix stiffnesses and The figure shows the Vernik-Kachanov model predictions
velocities. The vertical effective stress in wells 1 and 2 for both VP(0°) and VS(0°), compared with the sonic logs
averages 20 MPa. On the other hand, the Eagle Ford marls recorded in wells 1 and 4. Given the difficulties with ac-
in wells 3 and 4 are subjected to much smaller effective curately accounting for TOC-content, clay-content, and
stresses, averaging 13 MPa, because of a larger overpres- laminar-limestone-fraction variations in the vertical sec-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

sure. Additionally, they are characterized by greater com- tion, the model works reasonably well.
positional variations, notably with respect to a clay con- Even so, the question remains regarding whether,
tent averaging 22% in well 3 and 30% in well 4. Note that without adequate well control and core analysis, it is
there are at least two different subfacies with different but practical to use seismic data to simultaneously predict
overlapping distribution of clay content in wells 3 and 4, compositional variations (organic richness, clay and cal-
and the dominance of one over the other is responsible for cite content, and the like) and pore-pressure and stress
the mean-clay-content differences between the two wells. changes in organic shales. The nonnegligible stress sensi-
The log data in Figure 41 are superposed by the model tivities of VP(0°) and VS(0°) in mature Eagle Ford marls
realizations for a constant total porosity of 10%, φnk′ = 0.03, are obviously affected by the in situ presence of subhori-
two effective stresses (20 and 13 MPa), and three respec- zontal microcracks. Consequently, there may be prerequi-
tive levels of clay content (14.5, 22.5, and 30.5%) to sites for pore-pressure mapping from sonic-log-calibrated
achieve a more meaningful comparison with well data. seismic data in unconventional shales, if their lateral com-
The TOC content is implicit in equations 28 through the positional variations can be accounted for on the basis of
relationship with the volumetric kerogen content K (equa- independent information such as core measurements of
tion 3), which, in turn, enters the matrix-stiffnesses ex- organic richness and mineralogy. This important issue is
pression (the Backus) and the bulk-density expression, alluded to in more detail in Chapter 8.
both of which are necessary for velocity calculation.
The model line with vcl = 0.14 and σ = 20 MPa pro- VP-VS and AI-SI relationships
vides a good match to wells 1 and 2, whereas the model
lines with vcl = 0.22 and 0.30 at the same stress of 13 MPa Because of the strongly anisotopic nature of nearly all
give a proper account of the log VP-versus-TOC distribu- unconventional mudrocks, one always needs to specify the
tion in wells 3 and 4. All model realizations show a strong- direction in which P- and S-wave velocities and imped-
ly nonlinear effect of organic richness on velocity that is ances are being measured or for which they are being in-
consistent with the purely empirical curve-fitting model verted. The data acquired for characterizing unconvention-
introduced by Løseth et al. (2011) in AI-versus-TOC al reservoirs typically are available only from vertical pilot
space. wells drilled prior to the horizontal lateral wells placed for
In the dynamic range of stress and clay-content varia- stimulation and production. Therefore, it is customary to
tions presented in these wells, the main conclusion re- work with velocities measured in the direction very close to
garding the rock physics model application to compute the axis of symmetry of these low-structural-dip TI forma-
velocities in unconventional shales is that the composi- tions. As was discussed above, the velocities vary quite
tional effect clearly dominates other factors shaping their slowly at angles smaller than 7°, which is often within the
elastic response. However, the stress effect may also be range of the formation dip angle encountered in situ.
quite strong and should be accounted for in forward mod- Consequently, sonic velocities measured in vertical
eling. Crossplotting sonic velocities versus kerogen con- wells can approximate VP(0°) and VS(0°) values by ignor-
tent (or TOC content, as in Figure 41) and superposing ing minor structural dip angles and assuming a subsurface
pertinent model lines in the form of a template is an im- characterized by the vertical TI (VTI) symmetry. In this
portant step in any seismic petrophysics investigation of section, and later, in Chapter 8, just these velocities and
unconventional reservoirs. Such a template is instructive their corresponding acoustic and shear impedances refer-
in visualizing the effects of kerogen, of mineralogy, and ring to the symmetry axis of the VTI will be considered. In
of maturation-induced microcracks via the crack-density fact, it is convenient to focus on the bedding-normal direc-
parameter (which is, in turn, related to the effective stress tions in both synthetic forward modeling and also in simul-
in situ). taneous inversion of prestack seismic data, which must be
Because the model presented here takes into account tied to the existing vertical well control.
almost all variables bearing on sonic velocity in uncon- Figure 43 plots, in VP(0°)-versus-VS(0°) space, two
ventional mudrocks, it should be suitable for predicting subsets of the experimental core measurements listed in
sonic P- and S-wave velocity logs from information ac- Tables 4 and 5 (i.e., low-porosity organic mudrocks with
quired regarding the rock composition, porosity, and vcl values greater than 15% and tested at effective stresses
stress state. An attempt by Khadeeva and Vernik (2014) to of 50- to 70-MPa, and of 20-MPa, respectively). The very-
reconstruct the velocity logs is displayed in Figure 42. low-TOC shales and siltstones (Lockatong, Cotton Valley,

06-Vernik_Ch06.indd 161 12-08-2016 20:00:26


162 Seismic Petrophysics in Quantitative Interpretation

Well #1 Well #4

Vp_mod 20 MPa V s_mod 20 MPa Vp_mod 13 MPa V s_mod 13 MPa


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Vp Vs Vp Vs

Upper Upper
Eagle Ford Eagle Ford

Lower
Lower
Eagle Ford
Eagle Ford

Figure 42.  Vertical P- and S-wave velocity logs (black) for Eagle Ford wells 1 and 4 (Table 7), with markedly different
overpressure regimes superposed by the realization of the rock-physics model for the respective in situ effective-stress states
(red). Adapted from Figure 4 of Khadeeva and Vernik (2014). Used by permission.

and Chicopee) were excluded as nonreservoir lithologies. This curious tendency is confirmed again in Figure
Note that the effect of increasing stress is to shift data 43. Moreover, additional observations can be made on
points in the direction of a slightly greater VP(0°)/VS(0°) this more complete data set: (1) the majority of organic
value, relative to the dominant trend of the data cloud. The shales with vcl values greater than 15% plot between the
VP(0°)/VS(0°) isolines in Figure 43 are converted to pseu- pseudo-isotropic Poisson’s ratio lines of 0.15 and 0.25;
do-isotropic dynamic Poisson’s ratio νdyn values by using (2) the outliers to this rule are some Bazhenov and Kim­
the assumption of isotropy represented by vertical P- and meridge shales, which respectively display νdyn values
S-wave velocities. It is important to mention that in gen- smaller than 0.15 and greater than 0.25; and (3) the gen-
eral, νdyn ≠ ν31 ≠ ν13 ≠ ν12 in these TI rocks. Nonetheless, eral elongation of the cloud of data points (the trend) is
the pseudo-isotropic dynamic Poisson’s ratio lines are use- not quite parallel to the gas-sandstone and conventional-
ful references in VP-VS and AI-SI templates. shale lines, which are subparallel to each other. That
Also included in Figure 43 are the Sandstone Diagenesis clockwise deviation of the unconventional mudrock trend
Model for gas-saturated arenites at 20 MPa and the conven- in velocity space could be related to the impact of the
tional-shale-model line (both described in Chapter 3). These higher-velocity marls (from the Eagle Ford Formation
additions are to emphasize the unconventional properties of and the Niobrara Formation) — that is, it could be an ex-
organic mudrocks in a way similar to that of the VP–VS tem- pression of the effects of lithologic variation rather than
plate chosen by Vernik and Milovac (2011) and Khadeeva of compaction and diagenesis.
and Vernik (2014). As was mentioned in those studies, the Another important observation gleaned from Figure 43
organic-mudrock data cloud gravitates toward the gas-satu- is that immature organic-rich mudrocks that belong to
rated-sandstone line rather than toward the conventional ­maturity stage II (Ro < 0.5% and HI > 500 mg/gOC) plot
shales, for which the TOC content generally does not ex- in accord with the general trend, meaning that the overall
ceed 1.5%. data scatter is not related to organic maturity and/or the

06-Vernik_Ch06.indd 162 12-08-2016 20:00:28


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 163

water-wet, so that hydrocarbon fluids would have to ex-


hibit enormous capillary pressure in order to penetrate
those pores.
Even so, Khadeeva and Vernik (2014) have chosen
to ignore such complications and repeat the fluid-substi-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tution exercise of Lucier et al. (2011), starting with the


model in equations 28 for four representative vertical
wells drilled in the Bakken, Haynessville, Woodford,
and Eagle Ford plays. Khadeeva and Vernik conclude
that the fluid effect is secondary to the influence of solid
kerogen and mineral contents, which turn out to be the
main controlling factors for the position of log data in
the VP(0°)-versus-VS(0°) crossplot.
By excluding data from the Bazhenov and Kim­
meridge shale formations on the basis of their distinct
composition, it is possible to markedly reduce the data-
cloud scatter in Figure 43 and run a linear regression
through it to obtain a statistically meaningful trend line.
That line is shown in Figure 44 and has a slope b of
Figure 43.  Relationship between the P-wave velocity and 0.58 — slightly smaller than that of either the gas-sand-
the S-wave velocity for bedding-normal core measurements stone model or the conventional-shale model. This slope,
at 20 MPa (red) and high stresses of 50 to 70 MPa (green). A however, turns out to be quite similar to the slope of the
gas-saturated arenite line (at 20 MPa) and a conventional linear fit to the sonic-log-data cloud for the Austin Chalk
shale line are added for reference. The pseudo-isotropic and the Buda Limestone (containing minor conventional
Poisson’s ratios νdyn are computed using the bedding-normal shales), which overlie and underlie the organic marls and
velocities, assuming isotropy.
limestones of the Eagle Ford Formation (Texas, USA).
The median TOC value for the subset displayed in
presence of hydrocarbon fluid in kerogen pores. This find- Figure 44 is approximately 7%, which is also an upper
ing unambiguously indicates that (1) the abundance of bound for the Eagle Ford mudrocks. With that in mind,
kerogen and (2) its low Poisson’s ratio ν of 0.107 (based on Hu et al. (2015) formulated an empirical VS(0°) prediction
its effective moduli given in Table 6) are largely responsi- model for the Eagle Ford Formation that relies solely on
ble for the tendency of organic mudrocks to migrate away the information on VP(0°) and computed TOC logs, as
from the conventional-shale line and toward the gas-sand- follows:
stone line in either VP-VS or AI-SI space. The low Poisson’s
ratio inverted for porous kerogen is not unusual — compa- TOC
V S (0°) = bV P (0°) + (aref − a0 ) + a0 , (30)
rably low and, occasionally, even negative values are re- TOC ref

ported for polymer foams with a similar microstructure
(Lakes, 1987). where a0 and aref are the intercepts of the linear models, re-
An alternative explanation for the curious position spectively, for the background limestones (with TOC < 1%
of organic mudrocks in bedding-normal velocity and im- and a0 = −0.22) and for the organic marls (with a reference
pedance plots was suggested by Lucier et al. (2011) on TOCref = 7% and aref = a7 = 0.1, representing the maximum
the basis of fluid-substitution modeling, which led them organic richness in the area, i.e., TOCref ≥ TOC). This simple
to think that the presence of gaseous hydrocarbons is the model consists of a linear interpolation between the two lin-
cause of reduced vertical VP/VS and pseudo-isotropic ear trends, and it can serve as an empirical alternative to the
Poisson’s ratio values in mature organic shales. With re- more micromechanics-based Vernik-Kachanov model of
spect to fluid substitution, one should keep in mind that equations 28. Figure 45 illustrates the impressive accuracy
Gassman’s equation, which Lucier et al. used, is barely of the model predictions compared with log data for ten
applicable to mature organic shales because of their du- Eagle Ford vertical wells. The reduced-major-axis (RMA)
al-porosity system. The dual-porosity nature of these regression slope is close to unity.
rocks leads to a duality of wetting parameters that pre- It is important to emphasize, however, that to opti-
cludes water from entering the oil-wet kerogen pores, at mize the overall fit in other organic shale formations, both
least in maturation stages III through V. Moreover, as intercepts and the reference TOC value in this empirical
was discussed earlier in this chapter, the inorganic po- model will have to be adjusted on the basis of the mu-
rosity seems to be almost fully water-saturated and drock composition, effective stress, and available VS logs.

06-Vernik_Ch06.indd 163 12-08-2016 20:00:28


164 Seismic Petrophysics in Quantitative Interpretation

Certainly, formations such as Haynesville, Marcellus,


Bazhenov, and Kimmeridge Shale may all require differ-
ent input parameters for this empirical model of bedding-
normal S-wave prediction.
To gain full insight into the seismic amplitudes and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

especially the AVO attributes observed in typically aniso-


tropic unconventional reservoir plays, it is advisable to
use bedding-normal impedance crossplots and templates
rather than velocity ones. As was discussed in Chapters 3
and 4 for conventional siliciclastic sedimentary rocks, the
AI-SI template can be used not only for quality checking
the log data but also for generating an immediate predic-
tion of the AVO class and even of the strength of the AVO
gradient.
Figure 46 compares bedding-normal impedances
computed by Khadeeva and Vernik (2014) from density
and sonic logs from four vertical wells drilled in four dif-
ferent unconventional fields; as was previously m ­ entioned,
they used those wells to represent the organic mudrocks
of the Upper/Lower Bakken, Haynesville, Woodford, and
Figure 44.  A bedding-normal VP-versus-VS template, showing
Eagle Ford Formations. Again, the conventional-shale
a subset of the core data included in Figure 43 (here without
and gas-­saturated-sandstone lines are included for refer-
data from the Bazhenov and Kimmeridge shales), superposed
ence. Superposed on this plot are the core data from dif-
by the linear-regression line (blue) and a parallel line that
matches the limestones (with subordinate conventional shales)
ferent wells for the same formations, measured in room-
from the formations encasing the Eagle Ford Formation. dry conditions at 20 MPa — that is, at a stress close to the
vertical effective stress in situ. Overall, we can conclude
that there is good agreement between core data and log
3.5 data, although a minor mismatch occurs for the Eagle
Prediction: Ford rocks, which show slightly greater shear-impedance
VS = bVP + (a7–a0)TOC/7.0 + a0 (and VS) values measured in the laboratory. That mis-
3.0
match might be explained by the elevated density of natu-
ral subvertical fractures in situ that affect only the S-wave
velocity measured on the log scale, but, obviously, not on
VS predicition (km/s)

the core-plug scale.


2.5
Khadeeva and Vernik (2014) listed the following aver-
age kerogen contents and vertical effective stresses for the
unconventional reservoir in the four wells plotted in Figure
2.0 46: (1) Bakken: K = 0.27, σ  = 20 MPa; (2) Haynesville:
K = 0.1, σ  = 14 MPa; (3) Woodford: K = 0.13, σ  = 25 MPa;
Data points from and (4) Eagle Ford: K = 0.1, σ  = 20 MPa. The organic
1.5 10 wells shales of the Upper and Lower Bakken Formation are oil-
saturated; the Woodford Shale and Haynesville Formation
wells sampled reside in the wet- and dry-gas windows, re-
1.0
spectively, whereas the Eagle Ford Formation well pene-
1.0 1.5 2.0 2.5 3.0 3.5 trates the light-oil- to condensate-saturated marls and
VS log data (km/s) limestones.
1 25 Khadeeva and Vernik argue that although the high kero-
Color: Frequency gen content in the organic shales of the Bakken Formation
Figure 45.  Shear-wave-velocity prediction, on the basis of adequately explains the extremely low impedances, pushing
equation 30, in ten vertical wells penetrating the Eagle Ford them to the lower-left wing of the distribution cloud in the
organic mudrocks and encasing limestones (VS > 2.5 km/s), AI-SI space of Figure 46, the marked difference between the
with minor conventional shales. Here, the reference TOC Eagle Ford and Haynesville mudrocks in this plot clearly
value is 7%, b = 0.58, a0 = −0.22, and a7 = 0.1. Adapted from is not related to organic richness, which is approximately
Figure 2 of Hu et al. (2015). Used by permission. the same in these two formations. To the contrary, this

06-Vernik_Ch06.indd 164 12-08-2016 20:00:30


Chapter 6: Seismic Petrophysics of Unconventional Reservoirs 165

7.0

Gas sand
6.5
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

6.0

SI (km/s × g/cm3)
5.5

5.0

4.5

4.0

3.5 Shale
(TOC < 1%)
3.0
5 6 7 8 9 10 11 12
AI (km/s × g/cm3)
Figure 46.  A bedding-normal AI-SI crossplot with log data
Figure 47. An AI-SI crossplot of log data from four
from four wells — one each from the Bakken, Haynesville,
wells — one each from the Bakken, Haynesville, Woodford,
Woodford, and Eagle Ford Formations — compared with the
and Eagle Ford Formations — compared with the modeling
respective subsets of core data at 20 MPa. Adapted from
results from equations 28 for a range of total porosities
Figure 1 of Khadeeva and Vernik (2014). Used by permission.
from 4 to 12% (each represented by a small dot). Adapted
from Figure 5 of Khadeeva and Vernik (2014). Used by
difference in seismic rock properties can be fully accounted permission.
for by the higher carbonate content and greater effective
stress in the Eagle Ford Formation and the much higher clay The AI-SI data clouds and models presented in
content in the Haynesville Formation shale. Figures 46 and 47 for a variety of unconventional reser-
Interestingly, Khadeeva and Vernik (2014) were able voirs are instructive by themselves, but for them to realize
to use the model in equations 28 to accurately predict the their full potential the overlying and underlying forma-
loci of unconventional shales from all four wells in the tion impedances also must be added onto AI-SI displays
AI-SI space, with no need to adjust the model to the native with templates. Only then can one begin to envision the
fluid-saturation conditions. Their predictions were real- possible AVO signatures of the top and base reflections
ized for a range of total porosities, from 4 to 12%, and are associated with unconventional reservoirs. Such displays
included in Figure 47. will be used effectively in Chapter 8.

06-Vernik_Ch06.indd 165 12-08-2016 20:00:31


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 7: Geomechanics of Organic Shales

The anomalous elastic properties of unconventional oil- Poisson’s ratios: ν12, ν13, and ν31. Note that it is mislead-
and gas-shale reservoirs, described in Chapter 6, give rise to ing to refer to the Poisson’s ratios measured in vertical TI
controversial geomechanical concepts that persist in the in- (VTI) rocks as being vertical and/or horizontal, because
dustry despite significant discussion around them. Those there is only one that is purely horizontal: ν12. In terms of
concepts relate to brittleness and to estimation of least hori- the TI stiffness-tensor components Cij, a set of inverse for-
zontal stress from sonic and density logs in vertical bore- mulas expressing mechanical parameters, such as the
holes. To properly approach these leveraging geomechanics- Young’s moduli and Poisson’s ratios, is given as
related issues, it is necessary to consider the dynamic elastic 2
properties and, even more importantly, the static ones, be- C11 C33 − C33 C66 − C13
E3 = ,
cause the latter have a more straightforward bearing on a C11 − C66
rock’s ­affinity to fracture (its brittleness) and on vertical-to-­ 2
4C66 (C11 C33 − C33 C66 − C13 )
horizontal stress coupling in sedimentary basins. E1 = 2
,
C11 C33 − C13
In this chapter we (1) revisit, in TI shales, some of the
well-established relationships between the anisotropic 2C33 C66
ν 12 = 1 − 2
, (1)
tensors and the geomechanical elastic parameters, such as C11 C33 − C13
Young’s moduli and Poisson’s ratios; (2) critically review C13
the brittleness concept, especially in relation to shale ν 31 = ,
elasticity; (3) discuss estimation of the least horizontal 2(C11 − C66 )
principal stress (or fracture-closure gradient); and, final- 2C13 C66
ν 13 = .
ly, (4) explain, from a geomechanics perspective, the 2
C11 C33 − C13
­origin of the bedding-controlled, maturation-induced mi-
crocracks in organic shales.
3

Elastic-property relationships
E3 ν31
in TI shales
Because geomechanics, including the triaxial stress-
strain tests on rock samples, operates with mechanical
properties typically referred to as engineering parameters,
ν13
rather than with the elastic moduli Cij, let us take advantage
of the relationships among the anisotropic parameters, as 1
provided by Lekhnitsky (1963) and Fedorov (1968). Some
E1
of these relationships were mentioned earlier, with regard
to the model of a TI matrix with a distribution of cracks ν12
having a preferred orientation that is perpendicular to the 2
symmetry axis of the matrix (equations 25 of Chapter 6).
A schematic depiction of a TI sample is given in
Figure 1, in which axis 3 is the symmetry axis and the Figure 1.  Schematic description of the TI core sample, with
Poisson’s ratios are expressed using Lekhnitsky’s (1963) symmetry axis 3 and the anisotropic engineering parameters
notations. There are two independent Young’s moduli: E1 that can be measured in principal directions (Young’s moduli
and E3, because E2 = E1, and there are three different E1 and E3, and Poisson’s ratios ν12, ν13, and ν31).
167

07-Vernik_Ch07.indd 167 12-08-2016 19:59:19


168 Seismic Petrophysics in Quantitative Interpretation

The proportionality of the Poisson’s ratios ν13 and ν31 when also characterized by a lower-magnitude least hori-
to E1 and E3 is immediately established, which makes zontal stress, are deemed to lend themselves to easier and
only two out of the three Poisson’s ratios independent be- more-dispersed fracture generation and, hence, to more-ef-
cause ν31 = ν13(E3/E1) (see also equations 25 of Chapter fective stimulation.
6). In triaxial stress-strain testing programs this relation- Respectively, specialists in formation evaluation and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

ship necessitates acquisition of the Young’s modulus or rock mechanics have been striving to establish brittleness
the shear modulus in at least one nonprincipal direction criteria, with the goal of estimating brittleness from stan-
(e.g., at 45° to the symmetry axis, which is normal to the dard wireline logging data and determining landing inter-
bedding plane), in addition to the moduli at 0° and 90°, to vals for the horizontal laterals. For instance, Rickman
obtain five independent elastic constants for the full et al. (2008), among others, suggest that it is appropriate
­identification of the TI elastic tensor. Although the mea- to derive a rock-brittleness index from its elastic proper-
surement of E(45°) is not commonly accomplished in ties — that is, from the Young’s modulus and the Poisson’s
rock-mechanics laboratories, the expression for this ratio. Rickman et al. acknowledge that these properties
Young’s modulus is included here for completeness: would be best measured by using static loading condi-
tions and can be extended to log-based computations only
4C44 E1 E 3 upon proper calibration via triaxial stress tests. However,
E ( 45 ) = . (2)
E1 E 3 + C44 ( E1 + E 3 + 2ν 31 E1 ) the strong elastic anisotropy that is typical of organic

shales is neglected in log-based derivations of the Young’s
This expression also can be recast to solve for the softer modulus and the Poisson’s ratio, so it is unclear how cali-
of the two rigidities defining TI shale — that is, C44 = G13. bration should proceed without specifying which of the
When equation 2 is combined with equations 26 of directional Young’s moduli and Poisson’s ratios are taken
Chapter 6, the result completes the Cij tensor description into account. Moreover, a bigger question remains: Can
in terms of the static engineering parameters that can be brittleness, which by definition refers to the failure mode
derived from a series of triaxial tests on plugs drilled at of a material, be at all related to that material’s elastic
0°, 45°, and 90° to the symmetry axis. properties?
It is curious that the elusive, dynamic C13 stiffness, A material is said to be brittle if, when subjected to
which requires an accurate phase-velocity measurement on static triaxial stress loading, it ultimately fractures instead
the 45° plug, can be calculated readily from the static tri- of developing a more-distributed, ductile deformation
axial test results conducted only in principal directions (e.g., Jaeger and Cook, 1979). A ductile response results
(refer to the last f­ ormula of equations 26 of Chapter 6). On in a large strain and a small or negligible axial-stress in-
the other hand, the dynamic C44 stiffness is easily computed crement beyond the yield stress. Yield stress is defined as
from the bedding-normal S-wave velocity and does not re- the maximum axial stress that a rock can support before
quire testing of the plug drilled at 45° to the bedding plane its stress-strain curve starts to deviate from linearity
during static moduli acquisition implied by equation 2. (see below). As a consequence, all elastic moduli derived
Equations 26 of Chapter 6 and 1 and 2 here facilitate from the stress-strain curves obtained during static tests
a comparative analysis of the anisotropic parameters that must rely on the segments of those curves recorded before
are acquired by using P- and S-wave velocity measure- the yield point is reached.
ments (dynamic) and those derived from the directional One of the first attempts to introduce the definition of
strain data recorded in triaxial tests (static). The compari- brittleness in material science was made by Hucka and Das
son can be conducted in either the Cij stiffnesses domain (1974), who suggested several parametric ratios to be ex-
or the anisotropic engineering-­parameters domain. tracted from the static stress-strain curves. Perhaps the
most instrumental of those ratios are (1) the ratio of the
elastic strain to the total strain, as measured prior to failure,
Can brittleness be estimated from and (2) the ratio of the elastic energy stored in the loaded
elastic properties? sample to the total energy during failure. Arguably the
most sophisticated brittleness index, taking both pre- and
Unconventional shale reservoirs require multiple hy- postfailure portions of the stress-strain curve into account,
draulic-fracture treatments along the horizontal lateral wells is offered by Tarasov (2011). However, calculation of that
in order to enhance effective reservoir permeability, because index, ­although quite simple and intuitive, requires a spe-
the matrix permeability of those shales under in situ stress cial data-acquisition program carried out with extremely
conditions does not exceed 100 nd (e.g., Barree et al., 2015). stiff testing machines that store a minimal amount of elas-
The notion exists that tight organic mudrock formations tic energy during triaxial testing and allow stress-strain
often incorporate lithologic intervals characterized by a curves to be stably recorded postfailure. Unfortunately,
relatively greater brittleness. Such intervals, especially such setups are not commonly available.

07-Vernik_Ch07.indd 168 12-08-2016 19:59:19


Chapter 7: Geomechanics of Organic Shales 169

Rickman et al. (2008), the ratio E/ν should correlate posi-


f
200 tively with brittleness. However, ignoring the fact that
Rickman et al. neglect anisotropy, this ratio is much greater
for the limestone (134) than it is for the organic marl (69);
Differential stress (MPa)

ν31st = 0.31 y E3st = 41.6 GPa hence, if brittleness is estimated based on elastic properties
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

150
B = 0.58
of the two lithologies in question, then Bmarl < Blime, which
f
is in total contradiction to the more-direct brittleness crite-
100 ria employing strain and energy ratios.
y
The triaxial tests included in Figure 2 are quite repre-
sentative of the Eagle Ford Formation’s organic marls and
50 ν31st = 0.24 of low-porosity limestones in general. Thus, the elastic-
E3st = 41.6 GPa parameter-based brittleness estimation appears to be
B = 0.72 grossly misleading, at least in this unconventional shale
formation. Therefore, application of elasticity-based-brit-
0
–0.4 0.0 0.4 0.8 1.2 tleness profiling can erroneously lead an operator to land
Radial strain (%) Axial strain (%) horizontal boreholes in a zone with low kerogen, and,
hence, as follows from Chapter 6, low porosity and poor
Figure 2. Triaxial stress tests on low-kerogen limestone (red)
productivity. Let us revisit this issue after the following
and organic marl (blue) from the Eagle Ford Formation, with
discussion on static-dynamic-moduli relationships.
K = 0.104 (TOC = 4.0%), showing static elasticity parameters
E and brittleness B computed as the ratio of axial strains at
failure points f and yield points y.
Static and dynamic elastic
Figure 2 shows standard triaxial test results for two parameters in anisotropic mudrocks
main lithologies from the Eagle Ford Formation: (1) an or-
Because modern triaxial tests provide information
ganic marl and (2) a low-TOC limestone. Both tests were
about the static and dynamic Young’s moduli and
conducted in the bedding-normal direction (i.e., on subver-
Poisson’s ratios, it is tempting to conduct a comparative
tical plugs), and under the same lateral stress of 20 MPa to
analysis of static and dynamic elastic properties, espe-
simulate an in situ stress state. The kerogen content in the
cially in rocks with strong polar anisotropy. Mavko et al.
limestone shown is very low (K < 3%), whereas it is more
(2009) provide a comprehensive review of this subject by
than triple that amount in the organic marl (K = 10.4% that
considering isotropic moduli for quasi-isotropic sedimen-
corresponds to the TOC of approximately 4.0%).
tary rocks, such as sandstones and limestones. They con-
Brittleness B is computed by using the straightfor-
clude that the static Young’s modulus in room-dry rocks is
ward, albeit not entirely rigorous, strain-ratio approach of
almost always lower than the dynamically derived one,
Hucka and Das (1974):
especially at low stress levels when inter- and intragranu-
ey lar microcracks are still open and are contributing to non-
B = (3)
, linearly elastic and inelastic effects (e.g., sliding). Those
ef
effects are enhanced during the static tests, which are
where ey and ef are the axial strains at the yield point and characterized by strain amplitudes (10−3) that are orders
the failure point, respectively. As is shown in Figure 2, B is of magnitude greater than those induced by elastic wave
0.58 for the limestone and 0.72 for the organic marl. The propagation (10−7) used in dynamic measurements.
areas under the curve, up to the yield strain and the total The two data sets selected for the analysis and discus-
strain at failure, are respectively equal to (1) the elastic sion of the static versus dynamic elastic parameters in
energy stored in the rock and (2) the total energy expended anisotropic organic shales came from triaxial stress-strain
to break it. As mentioned above, according to Hucka and tests with simultaneous ultrasonic velocity acquisition, con-
Das (1974), the ratio of those two energies is another per- ducted by Corelab, Inc. and Metarock Laboratories. These
tinent brittleness index, and for the case illustrated in data sets include static and dynamic anisotropic moduli of
Figure 2 it is equal to 0.39 for the limestone and 0.53 for the organic mudrocks from the Eagle Ford and Haynesville
the marl. Those values are consistent with the strain-based Formations. Unfortunately, the Haynesville data set lacks
calculation: both methods show that Bmarl > Blime. the full elastic-tensor description from both the static tests
On the other hand, the limestone’s bedding-normal and the dynamic tests (no 45° plugs were available).
static Young’s modulus E3st is 41.6 GPa and its Poisson’s Consequently, for comparative analysis, an assumption is
ratio ν31 is 0.31, whereas the organic marl’s values are made that the relationship δ = 0.5ε, which approximates the
14.4 GPa and 0.24, respectively. Now, according to core measurements in Eagle Ford marls, is applicable to

07-Vernik_Ch07.indd 169 12-08-2016 19:59:19


170 Seismic Petrophysics in Quantitative Interpretation

both unconventional shale formations. This assumption is a)


consistent with the general database of velocity anisotropy
in organic mudrocks presented in Chapter 6.
As was previously indicated, the mean value of δ may
increase to δ = 0.6ε or even δ = 0.7ε under the lower stress
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

levels that are typical of in situ conditions conducive to


shales with a significant population of texture-controlled
open microcracks. However, that complication is ignored
for our purpose of comparing static versus dynamic mod-
uli, as is discussed below. The dynamic C13 stiffness was
simply computed on the basis of substituting δ = 0.5ε into
equation 35 of Chapter 1. In the next step, equations 1
were applied to convert the Cij dynamic stiffnesses to the
anisotropic dynamic Young’s moduli and Poisson’s ratios.
Figure 3a and 3b compare static E1st and E3st moduli
to dynamic moduli for the two data sets, using pseudo-
isotropic Edyn and anisotropic E1dyn and E3dyn computations
of Young’s moduli. The pseudo-isotropic pseudo-Young’s
modulus is computed from bedding-normal velocities,
with the assumption that the rocks are isotropic (ε = δ = 0), b)
even though that obviously is incorrect.
Several observations can be made on the basis of these
plots: (1) the static moduli are much softer than the dy-
namic ones in both directions, but E3st/E3dyn < E1st/E1dyn —
that is, the effect is stronger in the bedding-normal direc-
tion most significantly influenced by the bedding-con-
trolled microcracks in these mature organic shales (e.g.,
Figure 3a); (2) the relative difference between the static
and dynamic moduli is approximately the same for the
Haynesville shale and Eagle Ford marls, both of which can
be described by the following regression-based transforms:
E3st = 0.53E3dyn and E3st = 0.53Edyn; interestingly, these mu-
drocks also have approximately the same kerogen content
and stress sensitivity, albeit vastly different carbonate and
clay volume fractions; (3) the limestones from the Eagle
Ford Formation display approximately the same static-dy-
namic relationship as that for the bedding-parallel moduli
obtained on Eagle Ford marls and Haynesville shales; and,
finally, (4) the mismatch between the static moduli and the Figure 3. Static versus dynamic anisotropic (red) and
dynamic moduli is likely to be slightly less dramatic, but pseudo-isotropic (green) Young’s moduli for the
still significant, should the unloading stress-strain curves (a) Haynesville Formation, and the (b) Eagle Ford
be used in the static-modulus derivation (Sone, 2012). Formation. Regression lines to the pseudo-isotropic
The analytic results for the static Poisson’s ratio ν31st modulus Edyn (derived from the bedding-normal P- and
versus the dynamic Poisson’s ratio ν31dyn and the pseudo- S-wave velocities) are forced through the origin. The Eagle
isotropic Poisson’s ratio νdyn are displayed in Figure 4. Ford regression is focused on the organic marls.
Recall that ν31st is computed as the bedding-normal to
bedding-parallel strain ratio in TI rocks. Although the val- to ν31dyn. An approximate equality like ν31st ≈ ν31dyn is not
ues of this parameter are greater for the Eagle Ford mu- generally the case for the other two Poisson’s ratios (not
drocks, the static-to-dynamic relationship is the same: shown), for which ν13st > ν13dyn and ν12st < ν12dyn, as com-
ν31st ≈ ν31dyn, where ν31dyn was computed from equation 1 puted from the data sets in Tables 3, 4, and 5 of Chapter 6.
with inputs from all three plug orientations in velocity- The static and dynamic Poisson’s ratios obtained in
anisotropy measurements (C11, C66, and C13). Moreover, isotropic sandstones and carbonate rocks are often report-
the pseudo-isotropic Poisson’s ratio νdyn, obtained from ed also to be nearly the same (e.g., Mavko et al., 2009).
the bedding-normal VP/VS value, is statistically equivalent Interestingly, the isotropic-equivalent Poisson’s ratio,

07-Vernik_Ch07.indd 170 12-08-2016 19:59:21


Chapter 7: Geomechanics of Organic Shales 171

a) elasticity-based brittleness, as Rickman et al. (2008) sug-


gested. The log-derived values of the pseudo-isotropic
Young’s modulus and pseudo-isotropic Poisson’s ratio
from four vertical pilot wells, representative of the Bakken,
Woodford, Haynesville, and Eagle Ford Formations, must
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

be converted to static parameters that are at least partially


associated with the vertical direction in highly anisotropic
shales — that is, to E3st and ν31st.
From Figures 3 and 4, we learned that the static
Young’s modulus E3st = 0.53Edyn (except for limestones),
whereas ν31st ≈ νdyn. For the Eagle Ford data set as a
whole, which includes dual lithologies (organic marls and
limestones) plotted in Figure 3b, the following conversion
transform can be derived: E3st = 0.83Edyn – 10.4.
Figure 5 is a crossplot of the Young’s moduli (E3st and
Edyn) versus the Poisson’s ratios (ν31st and νdyn), with the
four wells each representing one of the four unconven-
tional shales mentioned above. The dynamic pseudo-­
isotropic parameters Edyn and νdyn, as computed from
b) sonic logs, are complemented in the crossplot by their
static counterparts, as derived from the dynamic-to-static
conversion transforms. Brittleness-index isolines, as sug-
gested by Rickman et al. (2008), are added as a template.
The Eagle Ford limestones are omitted in order to focus
on the potentially more comparable lithologies from the
four organic mudrock formations included in the analysis.
For reference, two low-porosity conventional siliciclastic
rocks (the Travis Peak sandstone and the Cotton Valley
shale), measured at approximately the same effective
stress, are also added in Figure 5.
As a result of dynamic-to-static conversion, Rickman
et al.’s elasticity-based brittleness index, which is com-
puted as an average of the Young’s modulus-based and
Poisson’s-ratio-based brittleness components, drops by
approximately 25% — from 0.6 ± 0.1 to between 0.35
and 0.55 for the four data sets included in Figure 5.
Unfortunately, other observations pertaining to this
Figure 4. Static Poisson’s ratio ν31st versus dynamic Poisson’s crossplot are more confusing than helpful — a statement
ratio ν31dyn and pseudo-isotropic Poisson’s ratio νdyn for the (a) that will be validated in the next several paragraphs.
Haynesville Formation and the (b) Eagle Ford Formation. First, the average brittleness index computed from
static parameters is noticeably lower for the Bakken
which can be approximated by the average of the three shales and Eagle Ford marls (0.41 ± 0.06) than for the
different values obtained in different principal directions Woodford and Haynesville shales (0.52 ± 0.07). There is
strongly TI shales (Chapter 6), yields a similar conclusion no logical explanation for this observation, because on the
— it is largely indifferent to the testing method, whether basis of its mineralogy the Haynesville shale is character-
that is static or dynamic. ized by a more significant average clay content and hence
would be expected to be less brittle. Similarly, the Bakken
shales and Eagle Ford marls have relatively low clay con-
Elasticity-based brittleness of organic tents that rarely exceed a vcl value of 20% — which theo-
mudrocks: A controversial notion retically would make them more brittle.
Second, the Eagle Ford limestone, with a vcl value
On the basis of dynamic-to-static calibration, it is now below 5%, is shown in Figure 2 to have an E3st value of
possible to employ a crossplot of the static Young’s modu- 41.6 GPa and a ν31st value of 0.31. If those values were
lus versus the Poisson’s ratio to provide an indicator of plotted in Figure 5, they would correspond to an “elastic

07-Vernik_Ch07.indd 171 12-08-2016 19:59:22


172 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 5. Brittleness-index crossplot, as suggested by Rickman et al. (2008), using the Poisson’s ratio and Young’s modulus.
Pseudo-isotropic elastic parameters Edyn and νdyn are derived from log data obtained in vertical wells using the assumption of
isotropy, whereas E3st and ν31st are derived from static-to-dynamic calibration results. Eagle Ford limestones are not included.
Dashed lines are apparent-brittleness-index isolines.

brittleness index” of approximately 0.6. Such a brittleness the apparent, elasticity-based brittleness estimated in
value for the limestone is higher than that for most of the each pilot well. That is an expected outcome, because al-
associated organic marls characterized by a higher clay though the high kerogen content makes the mudrock soft-
volume and shown in Figure 5. Curiously, that greater rel- er by lowering its E3 value, it also significantly reduces its
ative brittleness of the limestones would be quite logical bedding-normal VP / VS and ν31 values.
from a mineralogic perspective, if it were not totally con- Finally, within each specific formation the positive
tradictory to the results displayed in Figure 2. Those re- correlation of the elastic brittleness index values with the
sults unambiguously suggest that on the basis of the more TOC is supported by the more appropriate, strain-ratio-
relevant, rock-mechanics-based brittleness evaluation, the derived brittleness given by equation 3 and shown in
organic marl from the Eagle Ford Formation is, in fact, Figure 2. However, when different organic mudrock for-
significantly more brittle than the associated limestone. mations or lithologies are compared, the positive correla-
Third, if we choose to work with the dynamic tion breaks down. That breakdown is well exhibited by the
­properties derived from logs in vertical wells and ignore Bakken shales. Although they have a much greater average
anisotropy, then the crossplot in Figure 5 of the pseudo- organic richness than the other three formations consid-
isotropic Young’s modulus versus the pseudo-isotropic ered here, they also have an elasticity-based brittleness
Poisson’s ratio suggests that three out of the four uncon- index that turns out to be the lowest of all (Figure 5).
ventional formations considered plot along the same brit- As was previously mentioned, two conventional, low-
tleness index of approximately 0.6 ± 0.1, despite marked porosity siliciclastic rocks, measured at 20 MPa (the Cotton
differences in their mineral compositions. One observa- Valley shale and the Travis Peak sandstone, tested under
tion stands firm, however, whether one employs the dy- fully water-saturated and dry conditions, respectively) are
namic properties or the static ones — the greater the TOC included in Figure 5 for comparison with various organic
and hence the kerogen contents are within any specific mudrocks. Their elasticity-derived brittleness indexes differ
organic mudrock lithology considered here, the greater is dramatically, from 0.67 in this conventional shale to slightly

07-Vernik_Ch07.indd 172 12-08-2016 19:59:23


Chapter 7: Geomechanics of Organic Shales 173

more than 1.0 in the sandstone. Such a variation is at least Equation 4 assumes that the anisotropic Biot’s pa-
qualitatively consistent with the clay-content-based brittle- rameters are close to unity, because values derived from
ness estimates in conventional tight-sand/shale sequences. equation 22 of Chapter 2, or from its anisotropic versions,
Consequently, the E-ν-crossplot-based apparent brit- typically are much too low for proper calibration. The ne-
tleness can be appreciated intuitively in consolidated in- cessity for radical calibration immediately turns the start-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

organic shales and sandstones but not in unconventional ing theoretical model into an empirical exercise.
shale reservoirs. There, as the aforementioned detailed Equation 4 is often appended by the tectonic-stress
analysis suggests, the elastic brittleness index turns out to component, which is assumed to equal the hypothetical
be rather confusing and hardly adds any value to the in- tectonic strain eh multiplied by the horizontal Young’s
formation gleaned from other elastic-­property crossplots modulus E1 (e.g., Sayers, 2010). However, no solid evi-
such as VP-VS, AI-SI, or λρ-μρ. It also is quite obvious dence-based estimates of that tectonic strain exist in any
that we should avoid the term “brittleness” when referring specific area in question, so it becomes yet another purely
to an E-ν crossplot, unless we are referring to conven- empirical fitting term that is no more appropriate than is
tional tight siliciclastic rock formations for which an em- direct calibration of the uniaxial strain equation above to
pirical correlation with the rock’s true brittleness can be match the fracture-closure stress from the available DFITs.
established and the mechanical stratigraphy profiles using For the majority of conventional clastic sequences, it
this parameter can be meaningful. is totally inadequate to compute the stress-coupling-­factor
K0 log by using equation 20 of Chapter 2 with the true
isotropic Poisson’s ratio for sands and sandstones and the
Shmin stress-gradient estimation pseudo-isotropic Poisson’s ratio νdyn for shales. Instead,
some kind of pseudo-Poisson’s ratio v* has to replace
Uniaxial-strain-based approach them (e.g., M. McLean, personal communication, 2001) to
match the leak-off tests (see Chapter 2). The situation is
With few exceptions, unconventional shale reservoirs different for unconventional shale reservoirs, especially
occur in sedimentary basins that are subjected to normal- those with a highly elevated pore pressure caused by
and strike-slip-faulting stress regimes, meaning that the least ­kerogen-to-liquid hydrocarbon transformation. Clearly, an
horizontal principal stress Shmin in them is always smaller abnormally high pore pressure is a dominant variable that
than the vertical (overburden) stress. Hydraulic fractures cre- tends to mask the errors in K0 calculations, because the
ated during reservoir stimulation must overcome that resis- stress window for Shmin narrows markedly (PP < Shmin < Sv).
tance and will propagate in a perpendicular direction. Nonetheless, the vexing questions are: What does
Therefore, knowledge of the magnitude and orientation of control the stress coupling factor K0, and how can we esti-
Shmin is important in designing reservoir stimulation and esti- mate it in highly anisotropic organic shales? Vernik and
mating the stimulated reservoir volume. The basic formal- Milovac (2011) demonstrate that the controlling element
ism and the assumptions used in the subsurface-stress-state in estimating the anisotropic stress-coupling factor is
evaluation are provided in Chapter 2. Here we shall revisit Thomsen’s parameter δ.
some of the concepts that underlie application of the Shmin Indeed, by starting with that parameter’s definition, as
gradient to unconventional shale reservoirs. given in equation 35 of Chapter 1, we observe that C13 = C33
The most accurate information about Shmin comes – 2C44 when δ  = 0. Then, again following Thomsen (1986)
from diagnostic fracture-injection tests (DFITs), which in computing the stress-coupling factor K0 in TI rocks,
employ minihydraulic fractures to estimate both the pore
pressure and the fracture-closure stress. The latter is as- C13 C
K0 = = 1 − 2 44 , (5)
sumed to be a good measure of the Shmin. Typically, no C33 C33

more than two DFITs are acquired in a vertical pilot well,
so they can be used as calibration points for parameter- meaning that, if δ = 0, K0 is simply a function of the bed-
izing the uniaxial-strain-based equation 23 of Chapter 2 ding-normal moduli (or the sonic velocities measured in
for the Shmin gradient. That equation is repeated here in a vertical wells), regardless of how large the anisotropy pa-
more general form (e.g., Katahara, 2009): rameter ε is.
For the lack of a better choice, this special case for K0 is
Shmin G = K0 ( SvG − β 3 PpG ) + β1 PpG , (4)
referred to as being pseudo-anisotropic. By varying δ from
zero to a maximum value of δ = 0.8ε, which is suggested by
where K0 is the stress-coupling factor, Sv is the vertical stress, the database of organic shales, one can compute C13 by
PP is the pore pressure, β1 and β1 are the Biot’s poroelastic using equation 35 of Chapter 1, and then K0 = C13/C33. From
coefficients normal to bedding and parallel to bedding, re- equations 26 of Chapter 6, it is also possible to express the
spectively, and G in the subscripts refers to gradient. stress-coupling factor in the following anisotropic

07-Vernik_Ch07.indd 173 12-08-2016 19:59:23


174 Seismic Petrophysics in Quantitative Interpretation

functional form (which is similar to its more familiar isotro- a)


pic form written in terms of Poisson’s ratio):

C13 ν 13
K0 = = , (6)
C33 1 − ν 12
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/


where, somewhat counterintuitively, only the Poisson’s
ratios computed from the active horizontal strain and the
reactive vertical and horizontal strains in VTI are includ-
ed. It follows that the K0 value in general VTI settings can
be computed from the two Poisson’s ratios obtained on
the 90° core plug oriented parallel to the bedding plane,
which incidentally is the easiest to obtain. Then equation
6 can be conveniently employed for calculating both the
dynamic and static stress coupling in anisotropic
mudrocks.
It is straightforward to verify that increasing values of δ
result in increasing values for K0 and, consequently, for the
Shmin gradient (Vernik and Milovac, 2011). The Eagle Ford b)
and Haynesville core data for the dynamic and static modu-
li, discussed in the previous section, can be resorted to once
more to (1) test the static-versus-dynamic relationship for
K0 and (2) investigate simultaneously the effects of δ on K0
and, hence, on the Shmin gradient. Equation 6 and its special
pseudo-anisotropic case given by equation 5 are used in
Figure 6a and 6b, which compare the static stress-coupling
factor with the two realizations of the dynamic one: (1) the
pseudo-anisotropic case with δ = 0, independently of ε, and
(2) the anisotropic case in which 0 < δ  ≤ 0.7ε.
As is seen in Figure 6a (Haynesville core data), the
pseudo-anisotropic assumption would require a consider-
able amount of dynamic-to-static correction for the proper
computation of the stress coupling factor. However, when δ
is increased to 0.5ε, which is likely to be more typical of
the highly anisotropic Haynesville shale, virtually no dy-
namic-to-static correction may be necessary.
The same observation applies to the Eagle Ford data
Figure 6. Static versus dynamic stress-coupling factors,
set in Figure 6b, which shows that increasing δ from zero
using pseudo-anisotropic (δ  = 0) and anisotropic calculations
to 0.5ε and, in the best scenario, to 0.7ε, results in almost
(δ  = 0.5ε to δ  = 0.7ε) for the (a) Haynesville and (b) Eagle
perfect e­ quivalence between the static and dynamic cou- Ford core data sets. Green data points that plot close to the
pling f­actors. Note that in the latter case of δ = 0.7ε the K0st = K0dyn line are quasi-isotropic Eagle Ford limestones.
data cluster is almost perfectly bisected by the K0st = K0dyn
line. It is noteworthy that, the 0.5ε to 0.7ε range for the
parameter δ is actually observed with the Eagle Ford marls encountered at the in situ stress magnitudes found in
that were tested in the laboratory. Quasi-isotropic Eagle unconventional shale reservoirs (0.5ε < δ  < 0.7ε), the
Ford limestones tend to plot close to the K0st = K0dyn line, sonic velocities VP(0°) and VS(0°), or their approxima-
independently of the assumption regarding the value of δ. tions, in a vertical well can be sufficient to construct a
1D profile of the Shmin gradient. However, there are two
Anisotropy prediction in organic additional conditions that need to be satisfied: (1) the
and conventional shales pore-pressure gradient is well constrained, and (2) an
approximate anisotropy parameter ε is estimated from
Because it is evident that in estimates of the stress- VP(0°). Vernik and Milovac (2011) suggest the follow-
coupling factor no dynamic-to-static conversion is nec- ing simple empirical relationship, which is exponen-
essary for the most plausible range of δ values tially regressed from the data set of tight organic shales

07-Vernik_Ch07.indd 174 12-08-2016 19:59:24


Chapter 7: Geomechanics of Organic Shales 175

at the effective stress of 20 MPa presented in Table 5 of compaction and clay content, for anisotropic-velocity
Chapter 6: model-building in prestack migration of 3D seismic data.
The following empirical model seeks to capture both of
ε = b exp  −V P (0 )  , (7) those controlling factors:

Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

where b ≈ 12 ± 1 for 20 ± 5 MPa, but it can be slightly dif- ε = ε m − b exp[−1.5V P (0 )], (8)
ferent otherwise. When applied to the mentioned data set,
this model predicts ε values that are in good agreement where ε m = 0.052 vcl2 + 0.462 vcl − 0.093 , and b = 0.50v cl2 +
with measured anisotropy data (the coefficient of correla- 4.36v cl − 0.883. In seismic petrophysics applications,
tion R is 0.88), even though the model contains no explicit equation 8 should be applied to conventional shale only,
accounting for the known anisotropy agents, such as clay that is for intervals when Vsh = 1 (see Chapter 1). Massive
and kerogen content, in organic mudrocks. Actually, the sandstones and carbonates can be assumed quasi-­isotropic,
high predictive power of equation 7 is not surprising given whereas for thin-bed intervals including shale components
the fact that the same attributes of mudrock composition the Backus (1962) model can be used.
also determine their bedding-normal velocity (Chapter 6). The information on compaction in this model for con-
To illustrate this predictive power and simultane- ventional shales is implicit in the bedding-normal velocity
ously distinguish between tight organic mudrocks and term, whereas the first term in the model determines the
compacting conventional shales, the data set presented intrinsic anisotropy of the maximally compacted and re-
in Table 5 of Chapter 6 is complemented by a much crystallized shales with varying clay contents.
scarcer subset of quality-checked published core, log, An alternative empirical model, published by Li
and check-shot data on inorganic mudstones (Tosaya, (2006), is also designed to estimate the anisotropy param-
1982; Thomsen, 1986; Vernik et al., 2002; Dewhurst and eter ε as a function of the bedding-­normal velocity and the
Siggins, 2006; Shumaker and Vernik, 2009). This com- clay content. For a specified clay content, Li’s model pre-
pilation is plotted in Figure 7. dicts linearly increasing anisotropy with increasing veloc-
It should be emphasized that the empirical model of ity, which defies the common view that the clay-platelet
equation 7 assumes that organic mudrocks have under- orientation-distribution function (ODF) flattens exponen-
gone significant compaction and mineral recrystalliza- tially with compaction. It is logical that this flattening
tion. On the other hand, it is of separate interest in explo- should take place at different levels of polar anisotropy
ration geophysics to be able to predict the elastic anisot- predetermined by the clay content, as is shown in Figure
ropy of conventional shales, as a function of both 7. It also is important to note that the model line given by
equation 7 caps the family of lines from equation 8 on the
high velocity end.
More diverse and well characterized database includ-
ing both conventional and organic shales would have to
be built and analyzed to further explore interactions be-
tween the level of compaction, clay, and kerogen contents
in controlling vertical velocities and polar anisotropy pa-
rameters of mudrocks in general.

Stress profiling from log data


Returning to our geomechanical model, Figure 8 is an
example of a sonic-log-based Shmin gradient computation,
compared with the only DFIT available in this vertical
well penetrating the Eagle Ford Formation. As reflected
in the log-based limestone volume computation (track 4
of Figure 8), the Eagle Ford interval (3710 to 3763 m) is
encased between the Austin Chalk and Buda Formations.
The mean sonic velocity in the Eagle Ford interval domi-
Figure 7.  Anisotropy parameter ε versus the bedding-normal nated by organic marls is 4.00 ± 0.10 km/s, so ε = 0.22 is
P-wave velocity for organic and conventional shales. Models estimated from equation 7.
shown are from equation 7 for tight organic mudrocks (red) The pore-pressure-gradient profile is prescribed by
and from equation 8 for conventional mudstones with varying the heuristic model given by equation 9 (see the next sec-
levels of compaction and clay content (blue). tion) and calibrated with a DFIT interpretation that

07-Vernik_Ch07.indd 175 12-08-2016 19:59:25


176 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 8.  A pseudo-anisotropic (blue) and an anisotropic (red) fracture-closure stress-gradient profile for an Eagle Ford well,
compared with an interpretation from a single DFIT.

PPG = 17.6 kPa/m (0.78 psi/ft). Simple modeling using Due to extremely low permeability of tight shales the
equation 4 immediately suggests that the horizontal stress drained experiments necessary to measure those coeffi-
estimate is overwhelmed by the uncertainty with the pore cients are extremely difficult to realize. Note, however,
pressure gradient, which must be known precisely and the that no additional parameter, which sometimes is intro-
best way to constrain it is to run a DFIT. The bedding- duced to quantify the effect of δ on K0 (Sayers et al., 2015),
normal Biot’s coefficient β3 = 1.0, which is a reasonable is required to model the influence of TI mudrocks on Shmin.
assumption because of the elevated density of bedding- It is also unclear whether this exercise represents a
parallel microcracks in mature organic mudrocks. The genuine calibration, because the sensitivity to pore pres-
bedding-parallel Biot’s coefficient should be β1 < β3, but sure overwhelms the anisotropic effects in the uniaxial-
not by much, so β1 values in the 0.95 – 1.0 range are cho- strain-based fracture-gradient equation 4. More calibra-
sen to match the Shmin (or closure-stress) gradient of tion work definitely needs to be done, and additional
19.2 kPa/m (0.85 psi/ft), which was obtained from the DFITs must be acquired and carefully analyzed to make
DFIT. The estimates of the Shmin gradient, included in sure that the closure-stress-gradient profiles computed are
Figure 8, are given by equations 4 and 6: (1) for the pseu- accurate enough for modeling hydraulic fracture propaga-
do-anisotropic case, δ = 0; and (2) for the anisotropic tion. Nonetheless, the following is obvious: Fracture gra-
case, δ = 0.09, that is 0.41ε. dients may be more sensitive to pore-pressure distribution
As is seen in Figure 8, the anisotropic fracture-gradi- than they are to anisotropy parameterization.
ent profile (a proxy for the least horizontal stress gradi-
ent) computed by using δ = 0.09 and β1 = 0.95 matches
the DFIT-derived closure-stress gradient, but it should be Geomechanics of maturation-induced
noted that an alternative selection of β1 = β3 = 1.0 results microcracking
in a better match by the pseudo-anisotropic calculation.
Vernik and Milovac (2011) emphasize a similar ambi- As is demonstrated in Chapter 6, one of the most sig-
guity regarding an Shmin estimate that is based on wireline nificant outcomes of seismic petrophysics analysis of or-
log data from a Woodford well. The reasons for this ambi- ganic mudrocks is identification of high-density bedding-
guity are the uncertainty with δ = f(ε) input and the even subparallel open microcracks in cores that are subjected
greater uncertainty with the anisotropic Biot coefficients: to thermal-maturity stage IVa (Ro values of 0.75 to 0.9%)

07-Vernik_Ch07.indd 176 12-08-2016 19:59:26


Chapter 7: Geomechanics of Organic Shales 177

and, especially, to stage IVb (Ro values of 0.9 to 1.3%). Explanation 4 relates to a stress concentration that oc-
However, do the ultrasonic velocity- and anisotropy- curs at the bottomhole during drilling and that can be sig-
based interpretations lead us to believe that those open nificant at depths beyond approximately 3 km, in regions
subhorizontal microcracks are present in situ? characterized by high horizontal compressive stresses in
In general, there are five possible explanations for the the crust (Vernik et al., 1994). Cores retrieved from such
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

presence of texture-controlled microcracks in the core depths in the Williston Basin, USA are completely free of
samples of organic mudrocks: any saddle-shaped disking cracks, which are a very com-
mon manifestation of this effect. Moreover, stage IV
1. A differential elastic rebound of minerals occurs,
Bakken shales recovered from horizontal wells in our data-
caused by overburden stress relief;
base also contain interpreted bedding-parallel microcracks
2. Cooling-induced thermal cracking takes place;
that occur at a relatively high density and are oriented sub-
3. Overburden stress removal occurs, causing microhy-
parallel to the core axis. These horizontal cores display a
draulic fracturing (decrepitation) of highly overpres-
weak bedding-parallel SH-wave-velocity sensitivity to
sured fluid inclusions;
stress, contrary to the hypothetical bottomhole-stress-con-
4. A bottomhole concentration of stress develops that
centration effect, which would have introduced subvertical
may result in microcrack development and, eventual-
cracks parallel to the bottom of the horizontal well.
ly, in core disking; and/or
Consequently, the detection of open, bedding-subpar-
5. Microcracks, indeed, exist in situ.
allel microcracks in core samples of organic shales that
Note that the first four factors are basically technogenic are subjected to a certain stage of thermal maturity sug-
ones — that is, they are related to drilling and core-­ gests that these microcracks indeed should be present in
recovery operations. situ. This interpretation is also supported by the differ-
The stress-relief and cooling-mechanism explana- ence in wireline-sonic-logging velocities acquired in
tions (1 and 2 above) are the most plausible in accounting areas of different maturities and overpressures in the
for foliation-controlled microcracks in crystalline rocks Eagle Ford Formation. (Those velocity differences will be
(gneisses and amphibolites) recovered from great depths discussed further in Chapter 8.) The mechanism of that
(e.g., Vernik et al., 1994). In laboratory experiments, with microcrack formation is apparently related to overpres-
increasing stress these rocks also show an initially high- suring of the rock matrix in response to thermal matura-
gradient increase in bedding-normal velocity and an tion stages IVa and IVb — the main and final stages of oil
equally high-gradient reduction in anisotropy. generation. It should be noted that the horizontal micro-
However, these factors fail to account for the micro- cracks can be initiated a little earlier in the thermal-matu-
cracks inferred just in the stage IV upper/lower Bakken rity scale in certain organic shales not represented in our
shales, because maturity stages III and IV both represent the database, such as the Upper Devonian shale of the
oil window, and the cores tested were recovered from ap- Appalachian Basin, USA (Lash and Engelder, 2005).
proximately the same narrow depth interval of 3.0 to 3.5 km. Fracturing of source rocks has been discussed fre-
This means that the Bakken cores with stage III and IV ma- quently as a mechanism that enhances primary migration
turity characteristics have been subjected to roughly the of hydrocarbons via increased permeability (e.g.,
same stress relief and cooling. Nonetheless, they display Momper, 1978; Talukdar et al., 1987; Lehner, 1991). The
quite a dramatic difference in their sensitivity to stress — a best production from the Bakken Formation’s unconven-
sensitivity that may be caused just by microcrack ­closure. tional reservoir is often claimed to be related to open
In addition, it should be noted that cooling may be an bedding-plane fractures with residual oil staining that
­
important factor only when quartz, having an anomalously typically is associated with intervals containing higher
high thermal-expansion coefficient, is a major component amounts of clay and kerogen (Carlisle et al., 1992).
of the rock frame, as is the case in the metamorphic gneiss- McAuliffe (1979) proposed that oil-phase diffusion
es penetrated by the Kola superdeep well in Russia (Vernik through a 3D oil-wet kerogen network is a mechanism for
et al., 1994). To the contrary, in organic mudrocks, quartz primary oil migration. Also, predominantly horizontal
silt grains are suspended in the clay-kerogen or calcite- bitumen-filled microcracks or partings were found by
clay-­kerogen matrix, according to the SEM images pre- Lewan (1987) in artificially matured black shales of the
sented in Chapter 6. Woodford Formation, by Talukdar et al. (1987) in natu-
The importance of decrepitating isolated pores (ex- rally matured kerogen-rich limestones of the La Luna
planation 3) should also be questioned, because even Formation, Venezuela, by Lash and Engelder (2005) in
though significant overpressure can be attained during the Upper Devonian shales of the Appalachian Basin, and,
stage IV maturation, the formation-pressure gradient is finally, by Alan et al. (2014) in artificially pyrolized
also elevated within the neighboring, slightly less mature Bakken shales with a high kerogen concentration. Perhaps
areas of the Bakken Formation (Meissner, 1984). the most impressive petrographic evidence of fully or

07-Vernik_Ch07.indd 177 12-08-2016 19:59:26


178 Seismic Petrophysics in Quantitative Interpretation

partially bitumen-filled bedding-plane microcracks is


documented by Lash and Engelder (2005).
Consistently with earlier petrographic findings of
Lewan (1987) and Talukdar et al. (1987), Vernik (1994)
presents geomechanical arguments in favor of initiation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

and propagation sites for these microcracks within, or at


least adjacent to, kerogen laminae. The following discus-
sion leans on Vernik’s paper to emphasize the interplay
among the stress state, rock strength, and overpressure-
induced microcrack generation in mature organic shales.
Let us start by building hypothetical, but data-­
constrained, 1D stress and pore-pressure profiles of a ver-
tical cross section through the Eagle Ford Formation,
Texas, USA, where the far-field principal stress state can
be described as Sv > SHmax > Shmin. The vertical (overbur-
den) stress can be approximated by a straight line with the
overburden gradient (OBG) equal to 22.6 kPa/m (1 psi/ft).
The onset of overpressure is assumed to occur at approxi-
mately 3400 m.
Pore-pressure buildup below 3400 m is dramatic and
culminates at the hydrocarbon-generating Eagle Ford unit
at 3700 m. The following sigmoidal function can be used
Figure 9.  A total-stress profile for the subsurface, containing
to model the pore-pressure transition that occurs between
a mature Eagle Ford Formation interval.
3400 and 3850 m:

PPG min − PPG max


PPG = PPG max + , (9)
1 + ( z /zm )100

where PPGmax = 18.5 kPa/m, PPGmin = 10.0 kPa/m, and the


mean depth of the transition zone is zm = 3550 m. This
pressure profile has been adjusted to match the diagnostic
fracture-injection tests (DFITs) in the Eagle Ford and im-
mediately above it, in the Austin Chalk. At 3700 m — the
top of the Eagle Ford Formation at this location — the
overburden stress Sv is 84 MPa, whereas pore pressure PP is
68 MPa (0.81 psi/ft). Figure 9 shows the pore-pressure pro-
Figure 10.  The total-stress state above the onset of
file, in addition to the vertical stress and the least horizontal
overpressure generated by the Eagle Ford mudrocks (at a
principal stress, the second of which is considered next.
3000-m depth, in blue) and at the top of the Eagle Ford
The lower bound on the horizontal stress can be de- Formation (at 3700 m, in red), illustrated with Mohr circles.
termined from the condition of the frictional equilibrium Thin-lined circles correspond to the lower bounds, whereas
on optimally oriented faults (equations 26 and 27 of thick-lined circles represent the more likely stress state.
Chapter 2), assuming a coefficient of sliding friction of
ϕ = 0.6 on normal faults in the basin (Zoback and Healy,
1984). Evaluation of these conditions at 3000 m yields an This yields the Shmin gradient of 19.9 kPa/m and an Shmin
Shmin value of 42 MPa (0.63 psi/ft). It is quite obvious that value of 60 MPa at 3000 m. Together with the Sv value of
this value is the lower bound, whereas the actual horizon- 68 MPa, this least-principal-stress value forms the total-
tal stress should be greater in order to reduce the stress stress semicircle for the depth of 3000 m, which is plotted
deviator and move the Mohr circle away from the condi- as a Mohr diagram in Figure 10.
tion of permanent frictional sliding. The evaluation using the frictional equilibrium for the
A more realistic stress value, which is constrained by top of the Eagle Ford interval at 3700 m yields an Shmin
the LOTs, is given by equation 4 with the stress-coupling value of 73 MPa, whereas the more realistic stress value
factor computed on the basis of the pseudo-Poisson’s computed in the same fashion, as mentioned above, is
ratio (McLean’s formula in equation 25 of Chapter 2). more likely to be no less than 78 MPa (21.0 kPa/m). Note

07-Vernik_Ch07.indd 178 12-08-2016 19:59:28


Chapter 7: Geomechanics of Organic Shales 179

that although both of these estimates are represented by bedding-subparallel cracks should be equal to the rock’s
the respective Mohr semicircles in Figure 10, only the tensile strength T0:
more likely stress profile is included in Figure 9, which
is also better supported by DFIT tests in the area. ∆P = T0 = PPc − Sn , (10)
Interestingly, a similar stress value can be derived from
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

equation 4 with K0st = 0.6, which is only slightly outside where PPc is the critical pore pressure needed to generate
of the range that is typical of Eagle Ford marls (Figure 6). a microcrack. The tensile strength is a function of the
The question remains regarding whether the pore- size, geometry, and position of the original flaw (e.g., a
pressure gradient that is close to 18.9 kPa/m (0.82 psi/ft) micropore with an inflated penny-shaped geometry) as
is the maximum gradient attainable in the Eagle Ford and, well as of the fracture toughness of the host material.
furthermore, whether the deviatoric stress (Sv – Shmin) of LEFM states that for crack propagation to occur, the
6 MPa at the reference depth of 3700 m, and the corre- stress intensity at a flaw tip must exceed the critical stress-
sponding stress anisotropy of 7%, are the lower limits. An intensity factor (i.e., the fracture toughness). The expres-
argument can be made that the DFITs cannot measure the sion for the stress-intensity factors of a penny-shaped
virgin in situ stress because the formation could have un- crack between two dissimilar materials is (Kassir and
dergone a minor pore-pressure depletion because of on- Bregman, 1972):
going production.
If this assumption is valid, the stress anisotropy in the 2  Γ( 2 + iω ) 
organic marls at the late stages of oil generation could be K1 + iK 2 = ∆P c , (11)
π  Γ(0.5 + iω ) 
no greater than a few percent. Moreover, according to
Rice (1992), a weak layer (weak because of its mechani-
cal strength and extremely low effective stress) under a 1  G + G2 ( 3 − 4ν 1 ) 
where ω = ln 1 .
slightly anisotropic stress state with a structural dip angle 2π  G2 + G1 ( 3 − 4ν 2 ) 
of approximately 10° can cause a stress reorientation of as
much as 20°, compared with the far-field overburden. Here, K1 and K2 are the mode I (tensile) and mode II
This means that the vertical stress may no longer be the (shear) stress-intensity factors; G1, ν1, and G2, ν2 are the
maximum, even in the normal-faulting stress regime that shear moduli and Poisson’s ratios, respectively, of the two
prevails in the basin (Vernik, 1994). Under such circum- host materials (clay aggregate and kerogen, both taken
stances, the minimum energy that is needed to generate a from Table 6 of Chapter 6); and Γ is the gamma function
microhydraulic fracture can be dictated by the mechanical of a complex argument. Solving equation 11 for the ex-
strength anisotropy of the rock rather than by the stress tensional mode I stress intensity gives
anisotropy and its orientation.
The concept of mechanical strength is closely related
K1 ≈ 0.75∆ P c . (12)
to the concept of fracture toughness. Mode I fracture
toughness is a measure of a material’s resistance to tensile Thus, the tensile strength can be estimated if the critical
fracture propagation (Lawn, 1993). Defects such as pre- stress-intensity factor is known — that is, if the fracture
existing microcracks and pores in a rock can induce high toughness K1c of the weaker material and the dominant
concentrations of local stress, thereby initiating fracture pore semiaxis c are known.
propagation and material failure under substantially lower Given the highly anisotropic nature of the majority
stresses than the rock’s intrinsic strength. Under the con- of organic mudrocks, it is appropriate to consider two
ditions of a quasi-isotropic (low-deviatoric) state of stress fracture toughnesses and, respectively, two distinct ten-
operating in an active ­hydrocarbon-generating shale with sile strengths for the same rock: one that is normal to
nanodarcy permeability, the associated question is how bedding and another that is parallel to bedding.
the pervasive subhorizontal microhydraulic fracturing Anisotropic fracture toughness in shales has been
takes place in the originally crack-free, low-porosity rock widely reported (e.g., Costin, 1981), with K1c values dif-
matrix. To answer this, we need to consider mechanical- fering by a factor of 1.8 to 2.0 between bedding-parallel
strength anisotropy and some elements of linear elastic and bedding-normal orientations. Fracture toughness of
fracture mechanics (LEFM). the bedding-parallel cracks has been shown to be at least
From SEM observations, the size (length) of the long 60% lower than that of the bedding-perpendicular ones,
axis of large, fabric-parallel pores and microcracks (e.g., even in the relatively kerogen-poor and less-anisotropic
Figure 9 of Chapter 6) is in the range of 2c values from siltstones of the Ithaca Formation in the Appalachian
0.5 to 2.5 µm, where c is the major semiaxis of the origi- Basin, USA (Scott et al., 1992).
nal pore. The excess pore pressure needed to overcome The fracture toughness of kerogen can tentatively be
the total bedding-normal stress Sn and to initiate assumed to be equal to that of a brown coal that has

07-Vernik_Ch07.indd 179 12-08-2016 19:59:28


180 Seismic Petrophysics in Quantitative Interpretation

similar vitrinite reflectance Ro values, in the range of et al. (2007). Moreover, in a laboratory experiment by
0.4 to 0.6%, and is reported by Atkinson and Meredith Tissot and Pelet (1971), an excess pore pressure as great
(1987) to be in the following range: K1c is 0.006 to as 10 MPa was recorded in a shale sample under confin-
0.015 MPa m1/2. ing stress, until microfracturing caused a release of that
Tensile-strength anisotropy in organic shales is stress.
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

strongly manifested during sample handling and plug Estimates of the excess pore pressure that can be gen-
preparation. Even if no preferred orientation of the initial erated during kerogen-hydrocarbon transformation reac-
pores existed in a shale prior to microcracking, it is rea- tions in pores with an aspect ratio α of 0.01 to 0.05 also
sonable to assume the tensile-strength anisotropy to be of yield a similar range of values (Jizba, 1991; Carcione,
the same order as that observed for fracture toughness — 2000), even though these estimates should be regarded
that is, approximately a factor of two — because of the with caution because of the uncertainties in material-­
kerogen latticework permeating the rock. balance calculations. Those uncertainties are related to
Equation 12 is plotted in Figure 11 for K1c equal to primary migration of the reaction products, which are
0.005MPa m1/2 for the vertical tensile stress K1vc and hard to trace and account for. For instance, du Rouchet
0.01 MPa m1/2 for the horizontal tensile stress K1hc . The (1981) considered that generated oil is expelled from ker-
excess pore pressure for the original elongated pore with ogen by compaction and that the oil migrates through
a half-length of 1.5 to 2.5 µm should be in the range from laminae partings and microcracks.
4 MPa for the bedding-parallel crack-propagation direc- Figure 12 is another Mohr diagram, this time con-
tion to 10 MPa for the bedding-normal crack-propagation structed in terms of the effective stresses, by using the
direction. This means that the magnitude of the tensile- principal stresses and pore-pressure values estimated
strength anisotropy is almost twice the difference between above for the depths of 3000 and 3700 m (refer to Figure
the vertical and horizontal stresses estimated above for 10). Note that, as discussed above, for the depth of 3700 m
the final stages of oil generation at 3700 m, even if no as- centered at the hydrocarbon-generating Eagle Ford inter-
sumption is made regarding stress rotation. val the principal-stress directions may have rotated from
It is interesting to note that similar values for the ver- the vertical and horizontal, so they are simply identified
tical tensile strength T0v and horizontal tensile strength T0h as σ1 and σ3. Note also that Figure 12 illustrates the effect
were reported for the Woodford shale by Abousleiman of pore pressure cycling, which results in extremely low
stress anisotropy for the smallest red Mohr circle on the
left corresponding to the peak of pore pressure.
The lower strength envelope shown in Figure 12
is ­obtained by fitting the modified Griffith criterion
τ 2 = 6T0 (σ + T0 ) (Jaeger and Cook, 1979) to a single tri-
axial compressive strength test (not shown) for a weak lami-
nation orientation of approximately 45° on a typical Eagle
Ford organic marl sample. The T0 value for this envelope is
the uniaxial tensile strength perpendicular to bedding —
that is, T0v, which is the weakest tensile strength direction.
According to equation 10, as soon as the excess pore
pressure reaches the value of T0v, the stress circle touches
the weaker-strength envelope at a negative normal-stress
value, thereby causing a microcrack to initiate and propa-
gate in the perpendicular ­direction — which is the bed-
ding plane.
Laubie (2013) recently reached a similar conclusion,
stating that in VTI space there exists a critical original
crack length below which the favored crack orientation
Figure 11.  The excess pore pressure (differential stress) should be in the direction of the minimal resistance felt by
necessary for initiation of microhydraulic fractures in a the crack — that is, in the horizontal direction.
mature black shale, versus the major semiaxis of the original To reiterate, the implied effective bedding-normal
pore. Assuming a 1.5 to 2.5 μm major semiaxis of the stress σn, cycling from +16 to –4 MPa in Figure 12 for a
original pore and the fracture-toughness anisotropy shown, depth of 3700 m, is the result of the pore-pressure cycling
the tensile-strength anisotropy of the organic shale can be that should be characteristic of any source rock in the
estimated (T0v is 4 to 5 MPa and T0h is 8 to 10 MPa). Adapted lower oil window. Given the effective closure stress σclosure
from Figure 5 of Vernik (1994). Used by permission. and elastic constants of the crack-encasing material, the

07-Vernik_Ch07.indd 180 12-08-2016 19:59:29


Chapter 7: Geomechanics of Organic Shales 181
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 12.  The effective stress state in the far field (blue), and stress cycling, in a hydrocarbon-generating organic shale (red).
The strength envelopes (green dashed lines) pertain to the vertical and horizontal tensile strengths in the shale. A crack-
extension condition is attained when the deviatoric stress is reduced to a minimum (smallest Mohr circles on the left) and the
strength envelope for the horizontal crack extension, working against the weaker (vertical) direction, is reached.

maximum aspect ratio for the cracks that are closed under
this stress can be estimated (Nordgren, 1972) by

2σ closure (1 − ν 2) (13)
α = .
E
The evaluation results for kerogen (isotropic values
of elastic properties: Est = 0.5Edyn = 5.0 MPa, ν = 0.11)
and organic marl (mixed anisotropic/isotropic values:
Young’s modulus across bedding E3st = 15 MPa, isotro-
pic-equivalent Poisson’s ratio νiso = 0.24) are shown in
Figure 13 as a plot of the closure stress σclosure versus the
microcrack aspect ratio α = a/c. On the basis of the
elastic-moduli-­versus-stress curves that are typical of
Eagle Ford organic marls (e.g., Figure 35 of Chapter 6),
50 MPa can certainly be taken as a proxy for closure
stress for the vast majority of low-aspect-ratio inclusions
in this lithology.
The very concept of closure stress closure stress sug- Figure 13. The effective crack-closure stress versus the crack
aspect ratio for organic marl and kerogen as host materials.
gests that all inclusions with an aspect ratio α smaller
The σclosure range of 30 to 50 MPa, taken from ultrasonic
than 0.02 are essentially closed if they are encased in or
velocity measurements in room-dry organic marls from the
adjacent to kerogen sites (Figure 13). However, if the en-
Eagle Ford Formation, implies that all the microcracks with
casing material is approximated by the average properties an aspect ratio a smaller than 0.01 to 0.02 are closed at this
of the organic marl, even the microcracks with aspect ra- effective-stress range if hosted by kerogen. Adapted from
tios as low as 0.005 can survive an effective stress as high Figure 7 of Vernik (1994). Used by permission.
as 50 MPa. Obviously, all microcracks with aspect ratios
greater than 0.001 to 0.003 are open at an effective stress
level of 14 to 20 MPa, which is typical of the Eagle Ford To better identify the morphology and distribution of
Formation in situ. These microcracks, visible on some the open bedding-plane microcracks in mature organic
very-high-resolution SEM images (such as the one in shales, it is useful to employ mechanical constraints such
Figure 9 of Chapter 6) may markedly enhance the in situ as equation 13 for the aspect ratio, the crack-density-­
horizontal permeability of these rocks. parameter inversion from equations 25 of Chapter 6, and

07-Vernik_Ch07.indd 181 12-08-2016 19:59:30


182 Seismic Petrophysics in Quantitative Interpretation

FIB-SEM images of bitumen-filled and partially open mi- direction for future research. Such investigative work will
crocracks, such as those documented by Lash and be critical because to estimate crack permeability, one
Engelder (2005) and K. Milliken (Figure 9 of Chapter 6). needs to know both the length of the typical subhorizontal
In addition to studies of natural fractures of tectonic crack (that can be constrained by using the crack-density
origin (Sonnenberg et al., 2011), bedding-plane micro- parameter) and the crack aperture from the effective nor-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

cracks in unconventional shales should be targeted as a mal stress acting upon it.

07-Vernik_Ch07.indd 182 12-08-2016 19:59:30


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Chapter 8: Seismic Analysis in Unconventional Shales

Traditionally, seismic data processing for the de- The chapter discusses exactly how core measurements,
scription of unconventional shale reservoirs has been log data, and prestack 3D seismic data are used for de-
geared toward structural interpretation, fault mapping, tailed analysis and impedance inversion to produce organ-
and geometric-attribute computations such as for reflec- ic-richness volumes and to map the least horizontal stress
tion curvature, which is thought to be related to faults in situ. The Eagle Ford case may serve as a guide for what
and fracture intensity. These techniques of structural seismic analysis can deliver in the form of useful geologic
geophysics have only an indirect linkage to rock proper- and engineering information.
ties and, hence, to seismic petrophysics in general.
Additionally, microseismic data acquisition and process-
ing have made significant progress over the last several Reservoir-scale anisotropy estimation
years by employing the VTI (and its tilted variety) in ve-
locity model building for the purpose of improved micro- The upscaling of bedding-normal P- and S-wave ve-
seismic event location around hydraulically stimulated locities and polar anisotropy parameters is a challenging
horizontal wells (e.g., Van Dok, 2011). However, that but necessary step for proper calibration of surface-seis-
subject is also beyond the scope of this book. mic (and microseismic) data acquired from unconven-
On the other hand, attempts to obtain seismic rock tional shale reservoirs. The primary data sources for the
properties from surface seismic amplitudes and imped- most accurate velocity-anisotropy upscaling have been
ance inversions in oil and gas shales are still rare (e.g., and continue to be laboratory measurements on cores, and
Goodway et al., 2010; Gray et al., 2012; Hu et al., 2015). petrophysical-log-model computations that address reser-
In the majority of cases, those attempts are focused on voir heterogeneity. Both types of sources are available for
geomechanical parameters, such as Young’s modulus and the Eagle Ford shale reservoir, especially in the central
the Poisson’s ratio, or on λρ-μρ attributes, some of which area of its development.
may have an indirect bearing on rock brittleness, “fraca- The Eagle Ford core data set is an important part of
bility,” and stress state. For example, least-horizontal- the overall database presented in Tables 3, 4, and 5 of
stress-gradient estimates have been made from those prop- Chapter 6 and is characterized lithologically by thinly in-
erties in combination with pore-pressure data (Chapter 7). terbedded organic marls and limestones, the latter of which
In this chapter, we will use the compiled database of comprise 15 to 30% of the Eagle Ford’s gross thickness.
core measurements, log data, models, and interpretations Organic marls in the central area of the field, where the
presented in Chapter 6 and will extend the discussion into velocity and anisotropy measurements have been carried
the domain of surface seismic application and analysis. In out, are the predominant mudrock lithologies that are rela-
mature unconventional shales, organic richness, unload- tively low in clay content, with vcl values ranging from 14
ing-related microcrack development, and least-horizon- to 19% of the solid-rock matrix (including kerogen) and a
tal-stress profiles are all established to be among the most relatively stable TOC content of approximately 4.0  ± 2.2%.
important parameters that bear on hydrocarbon-in-place Despite the limited clay content in organic marls,
and hydraulic-fracturing-stimulation efficiencies, so we however, the intrinsic elastic anisotropy parameters ε and
will make them the focus of our investigation. δ (Tables 3 and 4 of Chapter 6) and their in situ values at
We use an Eagle Ford shale reservoir case history 20 MPa (Table 5 of Chapter 6) in these specific marls
throughout this chapter for two main reasons: (1) it is generally are quite high and average 0.23 and 0.15, re-
stratigraphically less complex than many other unconven- spectively. On the other hand, the anisotropy of lime-
tional resource plays and is resolvable by surface seismic stones is very weak, averaging an ε value of 0.06 at in situ
data, in contrast, for instance, to the Bakken play; and (2) stress conditions. It should be emphasized that, as was
it is a data-rich area of interest to many operators. previously mentioned, in other areas of the Eagle Ford
183

08-Vernik_Ch08.indd 183 12-08-2016 19:58:48


184 Seismic Petrophysics in Quantitative Interpretation

Formation the clay content in organic marls can be great- parameter ε from the bedding-normal sonic velocity, be-
er, thus potentially resulting in a stronger velocity cause both of them are controlled by the same set of vari-
anisotropy. ables (TOC, clay content, and effective stress) in organic
To upscale the available core data to the gross reser- mudrocks in general. Similarly, equation 8 of Chapter 7
voir scale that is relevant for seismic analysis, accurate can be applied to estimate ε in conventional shales in the
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

petrophysical-model-based laminar limestone-fraction log overburden. Limestones and dolomites that may encase
curves must be computed, as shown in Figure 8 of Chapter organic mudrocks can be safely considered to be quasi-
7, and calibrated against geologic core descriptions. Note, isotropic lithologies. In combination with the most plau-
that accurate computation of the laminar limestone log sible δ = 0.5ε relationship gleaned from our database,
curve is not a trivial task that requires solid petrophysics those estimates of ε also can be employed as necessary in
skills. anisotropic depth migration of seismic data.
The best choice for the upscaling is the anisotropic In principle, it is possible to further complicate the
Backus model (equations 34 of Chapter 1) because the anisotropic model by incorporating the azimuthal veloci-
dominant thickness of the limestones in question does ty variations that would render an orthorhombic symme-
not exceed 1.5 to 2 m, making them more than an order try. One way of doing that is to use the crack-density ten-
of magnitude smaller than the average Eagle Ford gross sor to introduce open vertical fractures with a strongly
thickness. Equations 34 of Chapter 1 help us evaluate the preferred orientation. However, such an exercise lies out-
effective polar anisotropy parameters of the gross Eagle side of the scope of this book.
Ford interval, and the model’s computed values for ε and
δ on the seismic scale are presented in Figure 1.
Not at all surprisingly, the upscaled TI parameters in Seismic ties
this binary geologic system are primarily determined by
the much-more-anisotropic component — that is, the or- The inaccuracies with either kinematic or dynamic
ganic marls. It can be gleaned from Figure 1 that for an (AVO) ties in seismic characterization are even less for-
average limestone fraction of 22% in the study area, the giving in unconventional shale reservoirs than they are in
upscaled anisotropy parameters are ε = 0.20 and δ = 0.13. siliciclastic and carbonate sequences. Surprisingly, the
Again, as was discussed in Chapter 6, greater values of primary reason for this challenging situation relates to
anisotropy can be predicted in areas that are character- the low gradients that are typical of the lateral variations
ized by greater kerogen or clay concentrations and lower in the key parameters controlling the elasticity of organic
effective-stress levels than those in the study area. mudrocks in the play area. As we have already learned
In the absence of reliable core measurements, equa- from Chapters 6 and 7, those parameters include the in-
tion 7 of Chapter 7 is useful in estimating the anisotropy trinsic parameters (TOC, mineralogy) but also the extrin-
sic ones (pore pressure, effective stress, stress anisotro-
py). Therefore, we may need to be able to detect small
variations in AVO attributes and AI-SI contrasts in order
to gain meaningful insight for their interpretation in
terms of the reservoir parameters of interest. In this re-
spect, the role of synthetic modeling and its comparative
analysis with seismic gathers becomes even more critical
than it is in conventional exploration and reservoir
characterization.
The necessary prerequisites for good ties between the
seismic and log data are threefold: (1) the prestack seis-
mic time- or depth-migrated gathers must be well-condi-
tioned and AVO-ready, (2) the log velocities and density
should be quality-checked and edited using VP-VS
crossplots with TOC or vcl values as a third variable, and
(3) the seismic wavelet ought to be adequately extracted
at the modeling location. The most important steps to fol-
Figure 1.  VTI anisotropy parameters ε and δ as functions low in seismic-gather conditioning are noise/multiple at-
of the limestone fraction in the Eagle Ford Formation. The tenuation, NMO correction and initial flattening, higher-
end-member anisotropy parameters are taken from Table 5 order residual correction, and spectral balancing. The lat-
of Chapter 6 by averaging separately over organic marls ter strives to unify the frequency spectrum and strength of
and limestones. the wavelet across the range of offsets.

08-Vernik_Ch08.indd 184 12-08-2016 19:58:49


Chapter 8: Seismic Analysis in Unconventional Shales 185

Synthetic forward modeling of the well data is the among other parameters, by unconventional-reservoir
most efficient process for marrying seismic gathers to ge- properties and stress states of interest.
ology when log data sets are checked for internal consis- Perhaps one of the most commonly overlooked ef-
tency. Such modeling could include depth shifting, fects in forward modeling of unconventional plays is the
crossplotting, model-based restoration of missing log effect imparted by the TI-parameter contrasts between
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

data, and blocking to minimize spurious log responses. As organic mudrocks and encasing rocks. It is mainly for
was mentioned above, the generation of synthetics relies this reason the anisotropy upscaling was considered first
heavily on the adequacy of the wavelet that is to be con- in this chapter. For the Eagle Ford mudrocks, the encas-
volved with log-based reflection coefficients. The most ing lithologies are predominantly quasi-isotropic carbon-
available option at the initial stages of modeling is statis- ates of the Austin Chalk and the Buda Limestone. The
tical wavelet extraction from the seismic minivolume over P-wave reflectivity as a function of the incident angle is
the target window centered at the well location. given by the approximate anisotropic AVO equations 9
Figure 2 shows a wavelet extracted from an Eagle and 10 of Chapter 4, which incorporate the VTI-parameter
Ford near-incident-angle minivolume centered at a verti- contrasts. Because these anisotropy contrasts are often
cal pilot well. The narrow-band character of this wavelet is comparable to the vertical (or at least near vertical) VP,
obvious and may hamper any reliance on raw seismic am- VS, and density contrasts across the top and base organic-­
plitudes for the quantitative interpretation. The amplitude mudrock interfaces (and are clearly so in the Eagle Ford
mapping can hardly deliver information necessary for ac- case), the VTI synthetics should be the preferred option
curate prediction of reservoir parameters. A more quanti- in AVO modeling.
tative approach must be sought in unconventional plays – As Figure 3 shows, both the isotropic and anisotropic
an approach that necessarily involves simultaneous im- synthetic AVO models tie the seismic CDP gather fairly
pedance inversion of seismic data. For the same reason accurately in a kinematic sense. However, the anisotropic
related to the wavelet uncertainty, band-limited (that is, synthetic also provides a better dynamic tie for the seis-
relative) AI and SI volumes are also unlikely to deliver sig- mic amplitudes, at least in the incident angle range from
nificant value in organic mudrock formations. 0 to 40°. Consequently, all other modeling examples in
Therefore, one of the most important tasks of the syn- this chapter will rely primarily on anisotropic AVO syn-
thetic forward modeling, as an ingredient of well/seismic thetics, with the VTI-parameter contrasts being numeri-
ties, is not only to get a good handle on the wavelet but cally equal to the upscaled ε and δ values — that is, with
also to constrain the vertical velocity and to refine the ini- an assumption of zero values for encasing carbonates.
tial impedance model (or low-frequency model). An ac- Those upscaled anisotropy parameters were calculated
curate low-frequency model is an integral part of the si- for the Eagle Ford Formation in the previous section.
multaneous inversion because it is employed for comput- An obvious feature of the synthetics shown in Figure 3
ing the absolute AI and SI volumes, which are controlled, is the relatively symmetric response of the top and base

Figure 2.  A statistical wavelet extracted from seismic data, for use in generating synthetics.

08-Vernik_Ch08.indd 185 12-08-2016 19:58:50


186 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 3.  Synthetic angle gathers with and without the effects of VTI-parameter contrasts, across the tight-limestone and
organic-marl interfaces at the top (trough) and base (peak) of the Eagle Ford interval. The VTI-parameter contrasts used in
anisotropic synthetics are Δε = 0.20 and Δδ = 0.13. A seismic CDP gather is shown on the right for comparison.

interfaces, irrespective of the asymmetric TOC distribu-


tion, which is characterized by vertical skewness and a
gradual reduction toward the overlying limestones of the
Austin Chalk. The main reason for the symmetric re-
sponse of the top/base interfaces in this particular well is
related to the Eagle Ford’s gross thickness of 40 m at the
well location, which is estimated to be slightly below the
tuning thickness for the dominant frequency of 25 Hz.
Unfor­tunately, however, greater gross thicknesses in the
central locations do not necessarily guarantee a reduction
in tuning effects, especially for the top interface, which
may be subjected to either destructive or constructive
­interference from the reflections inside the vertically het-
erogeneous overlying formation – the Austin Chalk.
A P-wave reflection coefficient’s dependence on the
angle of incidence is plotted as the tuning-free half-space
model in Figure 4. The models in the figure that corre-
spond to 2% TOC and 4% TOC averages for the Eagle
Ford interval show markedly different AVO intercepts and Figure 4.  P-wave reflection coefficients versus the incident
gradients. Again for comparison, both the isotropic and angle, for Eagle Ford organic marls with varying organic-
anisotropic Fatti-equation predictions are plotted. The richness values based on isotropic Fatti equations (dotted
lines) and anisotropic (solid lines) Fatti equations. The VTI
AVO curves clearly indicate that, for a range of incident
parameter contrasts are assumed to be the same as those in
angles smaller than approximately 40°, the VTI-based re-
Figure 3: Δε = 0.20 and Δδ = 0.13.
flection coefficients are much stabler and more consistent
with seismic CDP gathers; both show relatively strong shear-­impedance estimations from prestack inversion, the
positive AVO gradients that are typical of Class IV ­primary goal of which is to obtain accurate AI and SI vol-
behavior. umes for unconventional reservoir characterization.
Further discussion of the variation in organic richness
and its impact on seismic amplitudes and AVO behavior is
presented later in this chapter. One of the most important TOC estimation from an AI-SI template
lessons from this section is that the anisotropic versions
of the AVO equations should always be used, both in As was discussed in Chapter 6, the TOC content may
­simple half-space models and in synthetic gathers gener- be the most influential parameter for determining the qual-
ated from log data. Choosing an isotropic-synthetics ity of mature unconventional shale reservoirs, because it
­option may detrimentally affect our AVO analysis and determines their total and, notably, hydrocarbon-filled

08-Vernik_Ch08.indd 186 12-08-2016 19:58:51


Chapter 8: Seismic Analysis in Unconventional Shales 187

porosity. This is especially true in areas of any play in The results of density-model and velocity-model
which other controlling parameters, such as thermal matu- computations for varying TOC contents can be translated
rity, mineral composition, and overpressure, are found to readily into the impedance domain while keeping all other
be approximately constant. There­fore, mapping the TOC controlling factors the same. Depending on one’s prefer-
content or volumetric kerogen concentration from 3D seis- ences, the model computation of TOC content could also
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

mic data is a critical task in characterizing an unconven- be accomplished in AI-VP/VS or λρ-μρ domains, albeit
tional reservoir. Fortunately, the use of seismic amplitudes such computations will be much less straightforward.
and impedance inversions for remote TOC predictions has Figure 6 shows the AI-SI plot for 11 vertical pilot wells in
strong prerequisites that are based on core and log data ac- the central Eagle Ford area, color-coded for TOC logs
quired from vertical pilot holes (Chapter 6). Thus, it is in- (Hu et al., 2015). The log data displayed span a broad
structive to continue our focus on the specific Eagle Ford vertical interval, from the overlying Austin Chalk
Formation case history investigated by Hu et al. (2015).
Let us start with core data acquired in well 1, at the
reservoir stress conditions of 20 MPa, as presented in
Table 5 of Chapter 6. The clay content in this subset does
not exceed 19% for organic marls and 8% for limestones.
There are two subgroups of limestones in the Eagle Ford
interval: (1) those with relatively low TOC < 2.5% are char-
acterized by greater bed thickness, and (2) thinly laminated
limestones with elevated TOC of 5 to 7%.
Figure 5a plots the bulk density of the core samples
against the TOC values separately for these two main li-
thologies, using the same template as the one in Figure 14
of Chapter 6, again with the same kerogen-porosity model
lines superposed. Those lines are given by equation 10 of
Chapter 6, whereas the model parameters are included in
Figure 5a.
An interesting observation is that the organic lime-
stones tend to plot in Figure 5a to the right of the organic
marls. If the model input parameters used for the petro-
physical model realization are equally applicable to both
lithologies, that observation immediately implies that the
kerogen porosity is much lower in the limestones than it
is in the associated organic marls. In fact, many organic
limestones in the subset should be interpreted as having a
kerogen porosity close to zero, so their productivity is
likely to be very limited, even for the subgroup with high-
er TOC concentration. On the other hand, the position of
the organic marls on this template indicates that their ker-
ogen porosity averages 30 to 35%. This lithology is defi-
nitely the Eagle Ford pay that should draw special atten-
tion in any detailed rock-physics modeling.
Figure 5b shows the bedding-normal P- and S-wave
velocities from the same subset, plotted against the kero-
gen content derived from the TOC values. The data can be
visually compared with the rock-physics model given by
equation 28 of Chapter 6, with parameters similar to those
chosen for wells 1 and 2 in Figure 41 of Chapter 6. As is Figure 5. (a) TOC versus bulk density ρb, and (b) ultrasonic
evident from Figure 5b, the model turns out to be highly velocities at 20 MPa versus the kerogen volume fraction,
predictive and very useful in estimating the organic rich- measured for the organic marls and limestones of the Eagle
ness from seismic velocities and/or impedances, if other Ford Formation, from cores in the central area of the play.
controlling factors, such as the clay and calcite contents The data are superposed in (a) by the petrophysical model
and the effective stress, are not subjected to noticeable predictions for kerogen porosities φk, ranging from 10 to
lateral variations in the study area. 40%, and in (b) by the rock-physics model for organic shales.

08-Vernik_Ch08.indd 187 12-08-2016 19:58:52


188 Seismic Petrophysics in Quantitative Interpretation

limestones, through the Eagle Ford Formation, to the car- definition, the zero-TOC line of the AI-SI template can be
bonate rocks of the Buda Limestone at the base. viewed as the local background trend that corresponds
The two parallel lines superposed on Figure 6 as a to the formations that are encasing the Eagle Ford
template are zero-TOC and 7%-TOC reference models, mudrocks.
consistent with those presented in Figure 44 of Chapter 6. As is the case with impedance templates for conven-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Their slopes are b = 0.577, and their intercepts are tional rocks with a superposition of the AVO class bound-
a0 = −0.54 and a7 = 0.30, respectively. Note that the zero- aries (see Chapter 4), it is also instructive here, in the
TOC line is not the same as the conventional shale line AI-SI space of Figure 6, to superpose the ΔSI = 0.25ΔAI
shown in Figures 46 and 47 of Chapter 6. Instead, this is line passing through the middle of the Austin Chalk lime-
a generalized linear regression designed to pass through stone AI-SI cloud. The Austin Chalk is chosen as the ref-
the limestones and conventional shales of the Austin erence because it is an impinging medium for seismic
Chalk and to share the same slope with the 7% TOC refer- waves reflected from the top Eagle Ford interface.
ence line, which is more rigorously derived (see Chapter Even though the average VP/VS value of the forma-
6). The 7%-TOC line can be defined loosely as the bound- tions of interest here is lower than 2.0, this AVO Class-III-
ing line in AI-SI space for the study area, with Eagle Ford to-IV transition line approximately identifies the AI-SI
marls rarely exceeding this level of organic richness. By loci of the reflecting medium — the Eagle Ford — that

9.0
11 Eagle Ford wells

8.0 TOC = 7%

Buda

7.0 Austin
Chalk
Sl (g/cm3 × km/s)

ΔSI = 0.25ΔAl

6.0

5.0

4.0
8.0 10.0 12.0 14.0 16.0

Al (g/cm3 × km/s)
0 7
Color: TOC
Figure 6. An AI-SI template, showing log-derived impedances and TOC values (in wt. %) from 11 vertical pilot wells in the
central Eagle Ford area, with low clay content vcl values of 14 to 19%. Superposed are: (1) zero-TOC background-trend (red)
and 7%-TOC (dark blue) model lines; (2) the approximate boundary between Class III and Class IV AVO behaviors (dotted),
and (3) a vector connecting Eagle Ford organic marls with the overlying Austin Chalk limestones. Adapted from Figure 3 of Hu
et al. (2015). Used by permission.

08-Vernik_Ch08.indd 188 12-08-2016 19:58:53


Chapter 8: Seismic Analysis in Unconventional Shales 189

would produce a zero-AVO gradient for the top Eagle expected to be sensitive only to the average (or upscaled)
Ford seismic reflector. The blue line regressed through TOC content at the specific location. For that reason, it is
the Austin Chalk and Eagle Ford data clouds in Figure 6 informative to construct a set of pseudo-logs with per-
is appreciably rotated counterclockwise from this AVO turbed organic richness, starting from the original TOC,
class boundary. From the seismic rock-property perspec- density, VP, VS, AI, and SI logs. These pseudo-logs, of
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

tive, this observation implies that the top Eagle Ford in- course, are based on existing well data in combination
terface is expected to produce a strong Class IV response. with the rock-physics models presented in Chapter 6. By
The same behavior can be expected for the base reflection analogy with the fluid-substitution process in sands and
and is actually observed in both the synthetic and CDP carbonates, this type of controlled log perturbation can be
angle gathers (e.g., Figure 3). referred to as kerogen substitution.
A remarkable feature of the AI-SI plot in Figure 6 is Figure 7 demonstrates the kerogen-substitution lay-
the gradual migration of the Eagle Ford data, with in- out for Eagle Ford well no. 1. The TOC curve is computed
creasing organic richness, along the blue line passing using the petrophysical model discussed in Chapter 6 and
through the Austin Chalk limestones and directed away is shown to be well calibrated by the core TOC data in
from the background trend. Notice that the highest-­quality track 3, which average 4.0% for the Eagle Ford interval.
organic marls, with higher TOC values, approach the 7% The log-model-derived laminar limestone fraction is in
line and are characterized not only by lower impedances track 4. Tracks 5 through 8 show seismic rock-property
but also by lower vertical VP/VS values, which again is curves and modeled TI-parameter curves for the anisot-
consistent with the core database presented in Chapter 6. ropy parameters ε and δ.
That tendency for a reduction in VP/VS values in high- The original density, velocity, and impedance logs,
TOC mudrocks in general was first noticed by Vernik and which correspond to the TOC average value of 4.0%, are
Milovac (2011). presented in red and magenta in Figure 7, and those com-
High vertical resolution of the well-log data permits puted on the basis of reducing the TOC curve by 2 absolute
identification of vertically varying TOC values in indi- percentage points for the Eagle Ford interval are in green.
vidual wells. The much lower resolution of seismic data is Equation 10 of Chapter 6 is used to alter the bulk density

Vs_TOC 2 SI_TOC 2
0.5 KM/S 6.5 3 KM/S × G/CC 17
NPHI Vp_TOC2 AI_TOC 2
0.45 v/v –0.15 0.5 KM/S 6.5 3 KM/S × G/CC 17
Depth TOC_CORE RHOB_TOC 2 Vs SI Epsilon
feet 0 10 1.95 G/C3 2.95 0.5 KM/S 6.5 KM/S × G/CC 17 0 0.4
GR TOC_Log Lime RHOB Vp AI Delta
0 GAPI 150 0 % 10 0 v/v 1 1.95 G/C3 2.95 0.5 KM/S 6.5 KM/S × G/CC 17 0 0.4

12xxx

12xxx

12xxx

Figure 7.  Eagle Ford well data showing seismic rock-property logs perturbed by reducing the TOC content — from the average
of 4.0% actually observed in the Eagle Ford interval, to 2%. The kerogen-substituted density and velocity curves are shown in
green, whereas the original curves are in red and magenta.

08-Vernik_Ch08.indd 189 12-08-2016 19:58:54


190 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 8.  Synthetic angle gathers for the original set of logs (red) and those corresponding to a TOC content reduced to 2% (blue).
A strong, positive AVO gradient is seen in both cases, but the amplitude anomaly is markedly diminished in the lower-TOC case.

on the basis of a constant kerogen porosity of 30%.


Equation 28 of Chapter 6 is applied for bedding-normal
moduli and, finally, for bedding-normal VP and VS compu-
tations that correspond to the new average TOC value of
2%. In doing so, no attempt is made to alter the anisotropy-­
parameter curves, although some reduction in ε and δ val-
ues can be anticipated. Note that a mere 2% reduction in
TOC produces a predictable and noticeable effect on both
acoustic and shear impedances.
Figure 8 shows synthetic angle gathers generated
using (1) the original set of logs and (2) the set corre-
sponding to a TOC content reduced to 2%. Both sets have
been translated into a reflection-coefficient series and
then convolved with the same wavelet shown in Figure 2,
to produce the original and the reduced-TOC synthetic
angle gathers.
Both synthetics displayed in Figure 8 show an equal-
ly strong-gradient Class IV AVO response, as would be
predicted from the log-data positions in the AI-SI space.
However, the reduction in organic richness results in Figure 9.  Reflection coefficients RP and normalized
­substantial weakening of the originally prominent ampli- synthetic amplitudes versus incident angle for the top Eagle
tude anomaly. Another interesting observation can be Ford reflector in well 1, before and after kerogen-reduction
made here: The gross thickness of the Eagle Ford in the modeling. Solid lines are based on VTI approximations of the
model well is 53 m, which is only slightly above tuning. linearized three-term equations; dotted lines are based on
Significant constructive interference related to the nar- isotropic AVO equations.
row-band wavelet with prominent sidelobes, shown in
Figure 2, is definitely expected at the well location. That
interference, in turn, gives rise to a mismatch between the constant the other variables that control seismic rock
normalized amplitudes at the top Eagle Ford interface on properties — strongly suggest that it is feasible to use si-
the synthetic gather and the reflection coefficients, which multaneous inversion of prestack 3D seismic data for
are free of tuning effects (Figure 9). acoustic and shear impedances to predict and map the
The seismic petrophysics analysis in AI-SI space and TOC content over limited study areas for which well-­
the modeling for varied organic richness — while holding conditioned gathers are available.

08-Vernik_Ch08.indd 190 12-08-2016 19:58:55


Chapter 8: Seismic Analysis in Unconventional Shales 191

AI-SI Inversion of Prestack Seismic Data logs. Additionally, the empirical model given by equation
30 of Chapter 6 may provide an adequate estimation of VS
The process of simultaneous AI-SI inversion of in the Eagle Ford organic mudrocks and their encasing
prestack 3D seismic data is discussed in principle in carbonates, and that could be a straightforward alternative
Chapter 5. However, certain specifics of the seismic pet- to the rock-physics modeling (Figure 10). The multiwell,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

rophysics of organic shales render it helpful to reiterate vertical AI-SI crossplot of Figure 6 clearly suggests that
the main steps of the inversion workflow here: adding S-wave-velocity or shear-impedance information
helps us to better identify organic marls that have limited
1. Petrophysical log modeling to calculate the key reser- clay content, because they exhibit a reduced vertical
voir properties targeted by the inversion P-wave to S-wave velocity ratio relative to that of the en-
2. Rock-physics modeling to check quality or estimate casing limestones.
shear sonic logs in all the wells used in the inversion As an important preparatory step to simultaneous
3. Construction of a low-frequency model inversion, special attention should be paid to prestack
4. CDP gather conditioning and seismic/well tying with seismic gather conditioning and quality checking against
synthetic angle gathers; wavelet estimation the synthetic angle gathers constructed with edited and/
5. Simultaneous prestack inversion for AI and SI volumes or computed log data, including the upscaled VTI pa-
6. Construction of an AI-SI-TOC model, on the basis of rameter curves. Deficient dynamic ties, notably at mid-
log data and TOC volume computation from seismic to far-angle ranges, inevitably introduce significantly
AI and SI volumes unstable SI inversions. An operator should always be
For our case study, the bulk density, VP, and VS logs aware that although the synthetics may represent better-
from 11 vertical pilot wells from the area are first quality constrained AVO behaviors because of well-log editing
checked by using appropriate crossplots color-coded for and a­ ccounting for the polar anisotropy effects, ampli-
TOC and clay content. Then the TOC content is calculat- tude preservation in CDP g­ athers is critical and should
ed and core-calibrated for each of the key wells (Figure be adjusted only when hard-to-account-for discrepan-
10). The rock-physics model designed for the bedding- cies with well synthetics are observed.
normal direction in highly anisotropic unconventional Angle of incidence coverage out to 40° and good-
shales aids in quality checking and/or estimation of the VS quality, conditioned gathers turned out to be sufficient to
support a robust simultaneous inversion for AI and SI over
the area of interest in the Eagle Ford play. That was un-
dertaken by Hu et al. (2015), who (1) chose approximate-
ly evenly distributed wells (9 out of 11 included in Figure
6) to construct a low-frequency model, (2) generated AVO
synthetics to tie CDP gathers, and (3) ran a sparse-spike
simultaneous inversion with a merge frequency of 8 Hz.
Figure 11 shows the resulting inverted AI and SI vol-
umes, quality checked with well-log impedances from two
blind wells that were not used in construction of the low-
frequency model. As is illustrated by the close correlation
with AI and SI logs in these two wells, the overall inversion
can be qualified as superior for AI and reasonable for SI.
The relative instability of the SI inversion, notably at
one of the two blind wells located updip in Figure 11, can
be attributed to difficulties with far-angle amplitude pres-
ervation during gather conditioning and lateral wavelet
variations. Those pitfalls are aggravated by the fact that
current commercial prestack inversion software does not
adequately account for anisotropy, which in the general
case of organic shales clearly is not negligible.

Figure 10.  A log plot showing the petrophysical-model and TOC mapping
VS-log estimates (green) based on equation 30 of Chapter 6,
compared with the dipole shear log (blue). Adapted from As was mentioned earlier, the multiwell AI-SI log plot
Figure 1 of Hu et al. (2015). Used by permission. of Figure 6 reveals a simple empirical means for computing

08-Vernik_Ch08.indd 191 12-08-2016 19:58:56


192 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 11. Inverted AI and SI sections in the downdip direction, with blind-well impedances inserted as a quality check.
Top/base Eagle Ford horizons are shown in black. Adapted from Figure 4 of Hu et al. (2015). Used by permission.

the average TOC content of the Eagle Ford Formation at show virtually the same positions as do the log-­derived im-
each x-y location over the study area. Figure 12 shows the pedances, relative to the AI-SI template.
same AI-SI plot, populated now with AI and SI values ex- This observation immediately implies that an approx-
tracted from inverted volumes along the trajectories of the imate, upscaled TOC content can be estimated by using
wells used for inversion. The background-trend line and the following equation, which is similar in form to the
the 7%-TOC reference line forming the template are the net-to-gross calculation proposed by Vernik et al. (2002)
same as those shown in Figure 6. It is quite remarkable that for conventional sandstone reservoirs and described in
the inverted AI-SI data, also color-coded for the TOC logs, Chapter 3:

TOC (SI − bAI ) − a0


(w. %) TOC = TOCref , (1)
Invertion results aref − a0
6

7

5
where TOCref is the upper bound for the TOC value and is
Sl (g/cm3 × km/s)

7%-TOC marl
empirically selected for the study area. It follows from
6 4
Figure 6 that the slopes b of the model lines in this AI-SI
3
template are both 0.577 and that the intercepts a0 = −0.54
5
and aref = a7 = 0.30 correspond to the TOC value of zero
0%-TOC encasing rocks 2 and the reference value of 7%, respectively. The 7% refer-
ence likely represents the highest reservoir quality found
4 1 in the Eagle Ford organic marls in the central locations of
8 9 10 11 12 13 14 the study area, but in general it can be chosen arbitrarily
Al (g/cm3 × km/s) 0
during the template construction/calibration process. It is
Figure 12. An AI-SI template, showing inverted impedances important to emphasize the empirical nature of equation 1
extracted at the nine wells used in the inversion process, and its total dependence on the selection of input param-
color-coded for the TOC log data. Adapted from Figure 6 of eters, which can be numerically different for other geo-
Hu et al. (2015). Used by permission. graphic locations mainly because of the effects of mineral

08-Vernik_Ch08.indd 192 12-08-2016 19:58:58


Chapter 8: Seismic Analysis in Unconventional Shales 193

composition — notably, variations in clay and carbonate Finally, it should be pointed out that the mean and cumu-
contents, which will be discussed later. Therefore, the ref- lative TOC maps presented in Figures 13 and 14 both cor-
erence value for TOCref and, hence, the aref intercept can relate positively with the cumulative-production map for
be quite different in other organic-mudrock formations of the area (not shown). That is expected because of the
the world. strong dependence of organic porosity on organic rich-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Because SI inversion typically is less stable than AI ness, as is discussed in Chapter 6.


inversion is, Hu et al. (2015) designed a two-pronged ap- Note that although the method described above for
proach to honor both the equation above and the nonlin- computing the TOC volume from seismic data is quite
ear, Vernik-Kachanov model-based TOC effect on VP(0°) straightforward, care should be exercised when applying
(equation 28 of Chapter 6, and Figure 41 of Chapter 6), it in other areas of the Eagle Ford Formation or in other
now cast in terms of AI. The final TOC computation is unconventional reservoirs. Proper calibration using local
stabilized by using partial weights between the two ap- well control will be required. Importantly, the AI-SI
proaches. Those partial weights are derived from optimi- crossplot parameters that enter equation 1 should be ad-
zation of the match between the predicted TOC curve and justed to reflect local mineral composition and kerogen
the upscaled well-log-derived curve. distribution. Alternatively, a TOC volume can be derived
The mean and cumulative TOC maps of the Eagle solely from the nonlinear model relating it to AI, while
Ford interval are computed from the inverted TOC vol- using the effective-stress and mineralogic variables as
ume and are presented in Figures 13 and 14, respectively. input parameters.
Of course, the cumulative maps are affected by the thick- Figure 15 is an AI-SI template of log-derived Eagle
ness of the interval and, hence, by the seismic two-way- Ford impedances from the two wells also discussed in
traveltime thickness of the Eagle Ford Formation, which Chapter 6 — well 1, which is representative of the inversion
increases toward the southeast. study area, and well 4, which is located outside the TOC
Note that geologic features, such as some northwest- inversion maps (see also Table 7 and Figure 41 of Chapter
trending faults and areal trends of organic richness and 6). Figure 15 provides a good accounting of the factors,
variations in thickness, are clearly visible on both maps. other than TOC content, that affect the position of the data
Again, the validity of the inversion-based TOC estimates cloud. The organic marls from well 4 have an elevated clay-
was verified by comparing the predicted and upscaled- volume fraction along with a much larger overpressure gra-
well-log-derived TOC curves at the blind-well locations. dient that reduces the effective stress σ from 20 MPa to

Figure 13.  A mean-TOC map of the Eagle Ford Formation, computed from AI and SI inversion volumes in the study area.
Adapted from Figure 8 of Hu et al. (2015). Used by permission.

08-Vernik_Ch08.indd 193 12-08-2016 19:59:00


194 Seismic Petrophysics in Quantitative Interpretation
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 14.  A cumulative-TOC map of the Eagle Ford interval in the study area. Adapted from Figure 7 of Hu et al. (2015).
Used by permission.

8.0 13 MPa. Those factors together clearly shift the AI-SI val-
Two Eagle Ford wells
ues of the marls in well 4 toward the background trend line
TOC > 1.5% but also away from the Austin Chalk cloud. The straight
7.0 Austin
Chalk
cyan line connecting the Eagle Ford marls with the Austin
Chalk in this different area is rotated counterclockwise from
the blue line observed in the inversion study area, thereby
predicting an increase in a positive AVO gradient. That is
Sl (g/cm3 × km/s)

6.0
exactly what is seen in Figure 16, which shows the synthetic
angle gather generated using the well 4 logs.
Well 1
5.0 If the inversion study had incorporated a wider area of
the Eagle Ford (e.g., if it had included the well 4 location),
Well 4
application of equation 1 with the same parameters might
4.0 have compromised the quality of the TOC inversion by
markedly underestimating organic richness in locations
with elevated clay content. On the other hand, if the TOC
3.0 had been computed at these locations solely based on the
7.0 8.0 9.0 10.0 11.0 12.0 13.0 14.0 15.0 AI volume, organic richness would have been markedly
Al (g/cm3 × km/s)
overestimated. Thus, to guard against potential pitfalls, it is
0 0.5
advisable for the operator to integrate alternative approach-
Color: vcl es until the best-fit TOC volume inversion is obtained.
Figure 15. An AI-SI template, showing two Eagle Ford wells
(1 and 4) with markedly different clay and calcite contents. Shmin stress-gradient mapping
Only Eagle Ford interval is included for both wells, unlike
Figure 6. The relatively high clay content and greater Fracture gradients and fracture-closure gradients
overpressure found in well 4, which is located outside of the (assumed to be equal to the least-horizontal-stress Shmin
TOC inversion maps, combine to shift the cloud farther away gradients), when calibrated by DFITs, are, perhaps,
from the Austin Chalk and closer to the background trend, among the most leveraging parameters that pertain di-
with additional counterclockwise rotation of the cyan vector rectly to the fracability of any unconventional shale res-
relative to the blue one. ervoir. The subject of using seismic data in computing

08-Vernik_Ch08.indd 194 12-08-2016 19:59:01


Chapter 8: Seismic Analysis in Unconventional Shales 195
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 16.  A synthetic angle gather for well 4, characterized by the same TOC but a higher clay content and a lower effective stress
(of 13 MPa) than are found in well 1, which is representative of the inversion study area with its lower clay content and higher stress.

the least-horizontal-principal-stress gradient has been of the Eagle Ford study area is shown in Figure 7 of
discussed by Goodway et al. (2010) and Hu et al. (2015). Chapter 7. Because of the vertical-resolution limitation, it
Hu et al.’s more recent work demonstrates that inverted is not feasible to detect thin, localized intervals of elevat-
AI and SI volumes may be sufficient for estimating Shmin ed closure stress from the inversion volumes. That is un-
on the basis of the isotropic zero-lateral-strain equation fortunate, because such intervals potentially can be barri-
20 of Chapter 2. Still, some knowledge of the upscaled ers or arrest zones to vertical propagation of hydraulic
polar anisotropy parameters of organic mudrock forma- fractures. However, the gross differences in Shmin between
tions should be available to provide anisotropy-consis- the Eagle Ford Formation and the encasing carbonate
tent and, hence, more-accurate calculations. rocks can be detected, if accurate pore-pressure-gradient
In our case study, the anisotropic coupling factor be- profiles and maps are available.
tween the overburden stress and the horizontal stress Admittedly, information on the pore-pressure gradi-
(K0 = C13/C33) is defined by the four upscaled critical ele- ent in mature organic shales and in encasing rocks in
ments: VP(0°), VS(0°), ρb, and δ (e.g., Vernik and Milovac, general is quite sparse and is not readily available even
2011). Because it is virtually impossible to properly esti- from well data, unless several DFITs have been run in
mate density from the simultaneous inversion, the density different vertical depth points in horizontal lateral wells.
information should be derived from log upscaling with a However, even in this most optimistic scenario, it still is
subsequent 3D volume calculation (i.e., the low-frequen- not quite clear what the PP gradient profile should look
cy model for bulk density), using either AI or TOC like and how quickly the overpressure related to kerogen-
volumes as arguments. Then, bedding-normal-velocity
­ to-hydrocarbon transformation dissipates in encasing
volumes can be extracted from the respective inverted im- formations. When no accurate information on pore pres-
pedance volumes, whereas the upscaled polar anisotropy sure is available, the heuristic equation 9 of Chapter 7
parameter δ is obtained from the local TI parameters of can be helpful in constraining the pore-pressure transi-
organic marl and limestone cores in combination with the tion zone, especially when augmented by a DFIT and
laminar limestone fraction in the gross interval (e.g., drilling observables. Finally, an attempt can also be made
Figure 1). at estimating pore pressure and effective stress by invert-
The Shmin-stress-gradient calculation is based on equa- ing equation 28 of Chapter 6, if both mineral composi-
tion 4 of Chapter 7, with anisotropic Biot’s coefficients tion and kerogen content are known. The following exer-
β3 = 1 and β1 = 0.95 to match the local DFITs. The 1D cise sheds more light on this interesting but challenging
stress-gradient profiles are generated in key vertical pilot problem.
holes, as was illustrated in Chapter 7. For instance, the Revisiting Table 7 of Chapter 6 and focusing again on
log-derived Shmin profile for well 1 in the central locations wells 1 and 4, we notice marked differences not only in

08-Vernik_Ch08.indd 195 12-08-2016 19:59:03


196 Seismic Petrophysics in Quantitative Interpretation

average sonic velocities, but also in the clay contents and compute C33 and C44 values for organic marls only. Based
in the effective-stress gradients caused by varying over- on the laminar limestone fractions, the Eagle Ford marls
pressure gradients. It is important to note that the mean comprise 75% and 92% of the gross interval in wells 1
organic richness of the Eagle Ford in both well locations and 4, respectively. Next, the Vernik-Kachanov model
is approximately the same. There is another important pa- (Vernik and Kachanov, 2010), which was adapted to or-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

rameter that must be accounted for — the laminar lime- ganic mudrocks by Khadeeva and Vernik (2014), can be
stone fraction — that is log-derived and verified by a geo- applied. This is performed via equation 28 of Chapter 6,
logic description of cores extracted from key pilot wells. with the two-step calculation of C33m and C44m and the
Most of the thin limestone beds that are typical of the same model parameters used in constructing Figure 41 of
Eagle Ford are below log resolution. In well 1, the mean Chapter 6. However, this time we use the vertical effec-
laminar limestone fraction is 25% and the vertical effec- tive stress as an argument.
tive stress is 20 MPa, whereas in well 4 the mean lime- The results of this exercise are illustrated in Figure 17.
stone fraction is estimated to be 8% and the vertical effec- First, the error bars included for each well are designed to
tive stress is lower (13 MP) due to greater overpressure. indicate the uncertainties in both the velocity-extraction
The question then becomes: Is it at all possible to distin- and stress estimates from DFITs. Second, given the un-
guish pore-pressure effects from compositional effects? If certainties, the rock-physics model provides an excellent
the answer is yes, then it can be stated, in principle, that fit to the data, if the pore-shape factors are forced to be
seismic amplitudes, seismic velocities, or inversion prod- strongly clay-content-dependent and if p is smaller than q
ucts are capable of sensing the areas of dramatically dif- (refer to Chapter 6). Third, the differences between the
ferent vertical-effective-stress gradients in unconvention- organic marls from these two Eagle Ford locations in
al plays. terms of composition, such as the clay content, although
By using the average bedding-normal elastic con- quite significant, fail to fully account for the velocity dif-
stants C33 and C44 for the Eagle Ford limestone cores list- ference observed. The effective-stress factor, which in
ed in Table 5 of Chapter 6, and the sonic/density-log aver- this case is related to the difference in the pore-pressure
ages for the gross interval, it is straightforward to apply gradients, rather than to depth, adds noticeably (approxi-
Backus inversion to downscale the gross interval and mately 30%) to this velocity gap.
The only explanation we can provide for this additive
effect is the presence of subhorizontal microcracks in the
Eagle Ford marls, which are quite sensitive to the magni-
tude of vertical stress. That is probably one of the most
interesting conclusions that can be drawn from this mod-
eling exercise.
As was mentioned previously, well 4 is located far
outside of the inversion study area, so the final Shmin gradi-
ent volume is free of the potential complication that could
have been imparted by the dramatic lateral variations in
Eagle Ford composition. The pore-pressure-gradient map,
which focuses on the inversion area, is constrained by the
well data (Figure 18). The pore-pressure gradient is as-
sumed to be vertically constant for the Eagle Ford interval.
Together with the impedance and density volumes, which
were translated into elastic constants and then into the
VTI-based stress-coupling factor K0, this PP gradient is
used to finally arrive at the least-horizontal-stress-gradient
­volume via equation 4 of Chapter 7. The overburden-stress
Figure 17.  Velocity-versus-effective-stress lines given by the gradient is assumed to be laterally invariant and is kept
Vernik-Kachanov model (equation 28 of Chapter 6) realized constant at 23.1 kPa/m across the limited area of interest.
for organic marls of the Eagle Ford interval, compared with the Figure 19 presents a map of the Shmin stress gradient
sonic-log averages for organic marls in wells 1 and 4, which for the Eagle Ford Formation inside the study area. Quite
have similar kerogen values but different clay contents and predictably, it correlates strongly with the pore-pressure
vertical-effective-stress gradients. Organic marl velocities are gradient employed in the calculation. Nonetheless, in
inverted from the gross Eagle Ford interval’s bedding-normal some fault blocks where the pore-pressure gradient is ap-
elastic-stiffness averages, on the basis of downscaling with proximately the same, the Shmin map shows a negative cor-
laminar limestone content as the concentration parameter. relation with the average-TOC map. The negative

08-Vernik_Ch08.indd 196 12-08-2016 19:59:03


Chapter 8: Seismic Analysis in Unconventional Shales 197
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 18.  A pore-pressure-gradient map, constrained by well data in the study area. Adapted from Figure 10 of Hu et al.
(2015). Used by permission.

Figure 19. An Shmin gradient map for the Eagle Ford Formation inside the study area. Adapted from Figure 12 of Hu et al.
(2015). Used by permission.

08-Vernik_Ch08.indd 197 12-08-2016 19:59:06


198 Seismic Petrophysics in Quantitative Interpretation

correlation between the mean TOC and the least horizon- the best reservoir zones and, simultaneously, the most
tal stress for the Eagle Ford Formation indicates that fra- brittle and most fracable ones, because they are subjected
cability of this unconventional shale reservoir in areas of to lower horizontal stresses than are the encasing rocks.
limited clay content is affected positively by its organic As is stated in the introduction to this chapter, surface
richness. seismic applications to characterize unconventional shale
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This finding is consistent with the fracture-closure reservoirs remain predominantly restricted to attributes,
stress gradient in key calibration wells (e.g., Figure 7 of such as curvature and coherence, that are not directly
Chapter 7) and the earlier findings by Vernik and Milovac linked to seismic rock properties. However, it is likely to
(2011) for the Woodford shale. Consequently, another im- be just a matter of time until this status quo shifts, espe-
portant conclusion can be reached for the Eagle Ford that cially with further advances in surface seismic data qual-
may also apply to many other unconventional shale reser- ity and in microseismic and 4D seismic monitoring of the
voirs worldwide: The most organic-rich mudrocks render reservoir-stimulation process.

08-Vernik_Ch08.indd 198 12-08-2016 19:59:06


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

References

Abousleiman, Y. N., M. H. Tran, S. Hoang, C. P. Bobko, A. high-porosity siliciclastic sediments and rocks –– A
Ortega, and F.-J. Ulm, 2007, Geomechanics field and r­eview of selected models and suggested work flows:
laboratory characterization of the Woodford Shale: The Geophysics, 75, no. 5, 75A31–75A47.
next gas play: Annual Technical Conference and Bachrach, R., and P. Avseth, 2008, Rock physics modeling
Exhibition, Society of Petroleum Engineers, SPE- of unconsolidated sands: Accounting for nonuniform
110120-MS, http://dx.doi.org/10.2118/110120-MS. contacts and heterogeneous stress fields in the effective
Aki, K., and P. G. Richards, 1980, Quantitative seismology: media approximation with applications to hydrocarbon
Theory and methods, 1: W. H. Freeman and Co. exploration: Geophysics, 73, no. 6, E197–E209, http://
Alexandrov, K. S., and T. V. Ryzhova, 1961, Elastic proper- dx.doi.org/10.1190/1.2985821.
ties of rock-forming minerals II. Layered silicates [in Backus, G. E., 1962, Long-wave elastic anisotropy pro-
English]: Bulletin of the Academy of Sciences of the duced by horizontal layering: Journal of Geophysical
USSR, Geophysics Series 12, 1165–1168. Research, 67, no. 11, 4427–4440, http://dx.doi.org/10​.​
Alfred, D., and L. Vernik, 2012, A new petrophysical model JZ067i011p04427.
for organic shales: 53rd Annual Logging Symposium, Bandyopadhyay, K., 2009, Seismic anisotropy: Geological
Society of Petrophysicists and Well Log Analysts, causes and its implications to reservoir geophysics:
SPWLA conference paper 2012-217. Ph.D. dissertation, Stanford University, https://pangea.
Alkhalifah, T., and I. Tsvankin, 1995, Velocity analysis for stanford.edu/departments/geophysics/dropbox/SRB/
transversely isotropic media: Geophysics, 60, no. 5, public/docs/theses/SRB_118_AUG09_Bandyopadhyay.
1550–1566, http://dx.doi.org/10.1190/1.1443888. pdf, accessed 13 January 2016.
Allan, A. M., T. Vanorio, and J. E. Dahl, 2014, Pyrolysis- Barree, R. D., S. A. Cox, J. L. Miskimins, J. V. Gilbert, and
induced P-wave velocity anisotropy in organic-rich M. W. Conway, 2015, Economic optimization of hori-
shales: Geophysics, 79, no. 2, D41–D53, http://dx.doi. zontal-well completions in unconventional reservoirs:
org/10.1190/geo2013-0254.1. Hydraulic Fracturing Technology Conference, Society
Aplin, A. C., Y. Yang, and S. Hansen, 1995, Assessment of of Petroleum Engineers, SPE-168612-PA, http://​doi.
β, the compression coefficient of mudstones and its re- org/10.2118/168612-PA.
lationship with detailed lithology: Marine and Petro­ Bateman, R. M., 1990, Thin-bed analysis with conventional
leum Geology, 12, no. 8, 955–963, http://dx.doi.org/​ log suites: 31st Annual Logging Symposium, Society of
10.1016/0264-8172(95)98858-3. Professional Well Log Analysts, SPWLA Transactions,
Atkinson, B. K., and P. G. Meredith, 1987, Experimental paper II.
fracture mechanics data for rocks and minerals, in B. Batzle, M., and Z. Wang, 1992, Seismic properties of pore
K. Atkinson, ed., Fracture mechanics of rock: Academic fluids: Geophysics, 57, no. 11, 1396–1408, http://dx.doi.
Press, 477–522. org/10.1190/1.1443207.
Auld, B. A., 1973, Acoustic fields and waves in solids, I: Beard, D. C., and P. K. Weyl, 1973, Influence of texture on
John Wiley & Sons. porosity and permeability of unconsolidated sands:
Avseth, P., T. Mukerji, and G. Mavko, 2005, Quantitative AAPG Bulletin, 57, no. 2, 349–369.
seismic interpretation: Applying rock physics tools Benveniste, Y., 1987, A new approach to the application
to reduce interpretation risk: Cambridge Univer­ of Mori-Tanaka’s theory in composite materials: Mec­
sity Press, http://dx.doi.org/10.1017/CBO97805116​ hanics of Materials, 6, no. 2, 147–157, http://dx.doi.
00074. org/10.1016/0167-6636(87)90005-6.
Avseth, P., T. Mukerji, G. Mavko, and J. Dvorkin, 2010, Berg, R. R., 1970, Method for determining permeability
Rock-physics diagnostics of depositional texture, dia- from reservoir rock properties: Gulf Coast Association
genetic alterations, and reservoir heterogeneity in of Geological Societies, Transactions, 20, 303–317.
199

09-Vernik_References.indd 199 12-08-2016 19:58:30


200 Seismic Petrophysics in Quantitative Interpretation

Berge, P. A., B. P. Bonner, and J. G. Berryman, 1995, Carman, P. C., 1956, Flow of gases through porous media:
Ultrasonic velocity-porosity relationships for sandstone Academic Press.
analogs made from fused glass beads: Geophysics, 60, Carmichael, R. S., 1989, Practical handbook of physical
no. 1, 108–119, http://dx.doi.org/10.1190/1.1443738. properties of rocks and minerals: CRC Press.
Berryman, J. G., 1980, Long-wavelength propagation in Castagna, J. P., M. L. Batzle, and R. L. Eastwood, 1985,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

composite elastic media I: Spherical inclusions: Journal Relationships between compressional-wave and shear-
of the Acoustical Society of America, 68, 1809–1831, wave velocities in clastic silicate rocks: Geophysics, 50,
http://dx.doi.org/10.1121/1.385171. no. 4, 571–581, http://dx.doi.org/10.1190/1.1441933.
Biot, M. A., 1941, General theory of three-dimensional Castagna, J. P., M. L. Batzle, and T. K. Kan, 1993, Rock
consolidation: Journal of Applied Physics, 12, no. 2, physics: The link between rock properties and AVO
155–164, http://dx.doi.org/10.1063/1.1712886. ­response, in J. P. Castagna and M. M. Backus, eds.,
Biot, M. A., 1956, Theory of propagation of elastic waves Offset-dependent reflectivity –– Theory and practice of
in a fluid saturated porous solid. I. Low-frequency range: AVO analysis: SEG Investigations in Geophysics no. 8,
Journal of the Acoustical Society of America, 28, no. 2, 135–171, http://dx.doi.org/10.1190/1.9781560802624.
168–178, http://dx.doi.org/10.1121/1.1908239. ch2.
Blatt, H., 1982, Sedimentary petrology: W. H. Freeman and Castagna, J. P., H. W. Swan, and D. J. Foster, 1998,
Co. Framework for AVO gradient and intercept interpreta-
Bourbie, T., O. Coussy, and B. Zinsner, 1987, Acoustics of tion: Geophysics, 63, no. 3, 948–956, http://dx.doi.
porous media: Gulf Publishing Co. org/10.1190/1.1444406.
Bowers, G. L., 1995, Pore pressure estimation from veloc- Cheng, C. H., 2015, Can we ever trust the shear wave log?:
ity data: Accounting for overpressure mechanisms be- The Leading Edge, 34, no. 3, 278–284, http://dx.doi.
sides undercompaction: SPE Drilling & Completion, org/10.1190/tle34030278.1.
SPE-27488-PA, 10, no. 2, 89–95, http://dx.doi.org/10.​ Chopra, S., and J. P. Castagna, 2014, AVO: SEG Investi­
2118/27488-PA. gations in Geophysics No. 16.
Bowers, G. L., 2001, Determining an appropriate pore- Connolly, P., 2007, A simple, robust algorithm for seismic
pressure estimation strategy: Offshore Technology net pay estimation: The Leading Edge, 26, no. 10,
Conference, OTC paper 13042. 1278–1282, http://dx.doi.org/10.1190/1.2794386.
Bristow, J. R., 1960, Microcracks, and the static and dynamic Costin, L. S., 1981, Static and dynamic fracture behavior of
elastic constants of annealed heavily cold-worked oil shale, in S. W. Freiman and E. R. Fuller Jr., eds.,
­metals: British Journal of Applied Physics, 11, no. 2, Fracture mechanics methods for ceramics, rocks, and
81–85, http://dx.doi.org/10.1088/0508-3443/11/2/309. concrete: American Society for Testing and Materials,
Brown, A. R., R. M. Wright, K. D. Burkart, and W. L. ASTM STP 745, 169–184.
Abriel, 1984, Interactive seismic mapping of net pro- Couzens-Schultz, B. A., and K. Azbel, 2014, Predicting
ducible gas sand in the Gulf of Mexico: Geophysics, 49, pore pressure in active fold–thrust systems: An empiri-
no. 6, 686–714, http://dx.doi.org/10.1190/1.1441698. cal model for the deepwater Sabah foldbelt: Journal of
Brown, R. J. S., and J. Korringa, 1975, On the dependence Structural Geology, 69, Part B, 465–480, http://dx.doi.
of the elastic properties of a porous rock on the com- org/10.1016/j.jsg.2014.07.013.
pressibility of the pore fluid: Geophysics, 40, no. 4, Daley, P. F., and F. Hron, 1977, Reflection and transmis-
608–616, http://dx.doi.org/10.1190/1.1440551. sion coefficients for transversely isotropic media:
Budiansky, B., and R. J. O’Connell, 1976, Elastic moduli of Bulletin of the Seismological Society of America, 67,
a cracked solid: International Journal of Solids and 661–675.
Structures, 12, no. 2, 81–97, http://dx.doi.org/10.1016/​ Dellinger, J., and L. Vernik, 1994, Do traveltimes in pulse-
0020-7683(76)90044-5. transmission experiments yield anisotropic group or
Carcione, J. M., 2000, A model for seismic velocity and at- phase velocities?: Geophysics, 59, no. 11, 1774–1779,
tenuation in petroleum source rocks: Geophysics, 65, http://dx.doi.org/10.1190/1.1443564.
no. 4, 1080–1092, http://dx.doi.org/10.1190/1.1444801. Dewhurst, D. N., and A. F. Siggins, 2006, Impact of fabric
Carlisle, W. J., L. Druyff, M. S. Fryt, J. S. Artindale, and H. microcracks and stress field on shale anisotropy:
Von Der Dick, 1992, The Bakken Formation –– An inte- Geophysical Journal International, 165, no. 1, 135–
grated geologic approach to horizontal drilling, in J. W. 148, http://dx.doi.org/10.1111/j.1365-246X.2006.028​
Schmoker, E. B. Coalson, and C. A. Brown, eds., 34.x.
Geological studies relevant to horizontal drilling: Domenico, S. N., 1977, Elastic properties of unconsolidat-
Examples from western North America: Rocky Mountain ed porous sand reservoirs: Geophysics, 42, no. 7,
Association of Geologists, 215–226. 1339–1368, http://dx.doi.org/10.1190/1.1440797.

09-Vernik_References.indd 200 12-08-2016 19:58:30


References 201

Draege, A. M., M. Jakobsen, and T. A. Johansen, 2006, Faust, L. Y., 1951, Seismic velocity as function of depth
Rock physics modeling of shale diagenesis: Petroleum and geologic time: Geophysics, 16, no. 2, 192–206,
Geoscience, 12, no. 1, 49–57, http://dx.doi.org/10.1144/​ http://dx.doi.org/10.1190/1.1437658.
1354-079305-665. Fedorov, F. I., 1968, Theory of elastic waves in crystals:
Durand, B., and J. Espitalie, 1976, Geochemical studies on Plenum Press. http://dx.doi.org/10.1007/978-1-4757-
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

the organic matter from the Doula basin, Cameroon, 1275-9.


Evolution of kerogen: Geochimica et Cosmochimica Flemings, P. B., B. B. Stump, T. Finkbeiner, and M. Zoback,
Acta, 40, no. 7, 801–808, http://dx.doi.org/10.1016/​ 2002, Flow focusing in overpressured sandstones:
0016-7037(76)90032-6. Theory, observations, and applications: American
Durand, B., A. Y. Huc, and J. L. Oudin, 1987, Oil saturation Journal of Science, 302, no. 10, 827–855, http://dx.doi.
and primary migration: Observations in shales and org/10.2475/ajs.302.10.827.
coals from the Kerbau wells, Mahakam Delta, Füchtbauer, H., 1967, Influence of different types of dia-
Indonesia, in B. Doligez, ed., Migration of hydrocar- genesis on sandstone porosity: 7th World Petroleum
bons in sedimentary basins: 2nd IFP Exploration and Congress, Proceedings, 2, 353–369.
Production Research Conference, Proceedings, Gal, D., J. Dvorkin, and A. Nur, 1998, A physical model for
Editions Technip, 173–196. porosity reduction in sandstones: Geophysics, 63, no.
du Rouchet, J. D., 1981, Stress fields, a key to oil migra- 2, 454–459, http://dx.doi.org/10.1190/1.1444346.
tion: AAPG Bulletin, 65, no. 1, 74–85. Gardner, G. H., L. W. Gardner, and A. R. Gregory, 1974,
Dutta, N. C., 2002, Geopressure prediction using seismic Formation velocity and density – The diagnostic basis
data: Current status and the road ahead: Geophysics, for stratigraphic traps: Geophysics, 39, no. 6, 770–780,
67, no. 6, 2012–2041, http://dx.doi.org/10.1190/1.​ http://dx.doi.org/10.1190/1.1440465.
1527101. Gassmann, F., 1951, Über die Elastizität poröser Medien
Dvorkin, J., and A. Nur, 1996, Elasticity of high-porosity [On elasticity of porous media]: Vierteljahrsschrift der
sandstones: Theory for two North Sea data sets: Naturforschenden Gesellschaft in Zürich, 96, 1–23.
Geophysics, 61, no. 5, 1363–1370, http://dx.doi.org/​ Goodway, B., T. Chen, and J. Downton, 1997, Improved
10.1190/1.1444059. AVO fluid detection and lithology discrimination using
Eastwood, R. L., and J. P. Castagna, 1987, Interpretation Lamé petrophysical parameters “λρ”, “μρ”, & “λ/μ fluid
of VP/VS ratios from sonic logs, in S. H. Danbom stack”, from P and S inversions: 67th Annual Interna­
and S. N. Domenico, eds., Shear-wave exploration: tional Meeting, SEG, Expanded Abstracts, 183–186,
SEG Geophysical Developments Series No. 1, http://dx.doi.org/10.1190/1.1885795.
139–153. Goodway, B., M. Perez, J. Varsek, and C. Abaco, 2010,
Eaton, B. A., 1975, The equation for geopressure prediction Seismic petrophysics and isotropic-anisotropic AVO
from well logs: Society of Petroleum Engineers, SPE methods for unconventional gas exploration: The
5544-MS, http://dx.doi.org/10.2118/5544-MS. Leading Edge, 29, no. 12, 1500–1508, http://dx.doi.
Eaton, B. A., and T. L. Eaton, 1997, Fracture gradient pre- org/10.1190/1.3525367.
diction for the new generation: World Oil, 128, Gray, D., P. Anderson, J. Logel, F. Delbecq, D. Schmidt,
93–100. and R. Schmid, 2012, Estimation of stress and geome-
Eberhart-Phillips, D. M., 1989, Investigations of crustal chanical properties using 3D seismic data: First Break,
structure and active tectonic processes in the coast 30, no. 1821, 59–68, http://dx.doi.org/10.3997/1365-
ranges, central California: Ph.D. dissertation, Stanford 2397.​2011042.
University, https://pangea.stanford.edu/departments/ Gratwick, D., and C. Finn, 2005, What’s important in mak-
geophysics/dropbox/SRB/public/docs/theses/SRB_​ ing far-stack well-to-seismic ties in West Africa?: The
036_JAN89_EberhartPhillips.pdf, accessed 13 January Leading Edge, 24, no. 7, 739–745, http://dx.doi.
2016. org/10.1190/1.1993270.
Ehrenberg, S. N., 1990, Relationship between diagenesis Grechka, V., and M. Kachanov, 2006, Effective elasticity of
and reservoir quality in sandstones of the Garn fractured rocks: A snapshot of the work in progress:
Formation, Haltenbanken, mid-Norwegian continental Geophysics, 71, no. 6, W45–W58, http://dx.doi.org/​
shelf: AAPG Bulletin, 74, no. 10, 1538–1558. 10.1190/1.2360212.
Fatti, J. L., G. C. Smith, P. J. Vail, P. J. Strauss, and P. R. Greenberg, M. L., and J. P. Castagna, 1992, Shear-wave
Levitt, 1994, Detection of gas in sandstone reservoirs velocity estimation in porous rocks: Theoretical
­
using AVO analysis: A 3-D seismic case history using ­formulation, preliminary verification and applications:
the Geostack technique: Geophysics, 59, no. 9, 1362– Geophysical Prospecting, 40, no. 2, 195–209, http://
1376, http://dx.doi.org/10.1190/1.1443695. dx.doi.org/10.1111/j.1365-2478.1992.tb00371.x.

09-Vernik_References.indd 201 12-08-2016 19:58:30


202 Seismic Petrophysics in Quantitative Interpretation

Greensmith, J. T., 1989, Petrology of the sedimentary rocks, Hower, J., E. V. Eslinger, M. E. Hower, and E. A. Perry,
7th ed: Unwin Hyman. 1976, Mechanism of burial metamorphism of argilla-
Guéguen, Y., and V. Palciauskas, 1994, Introduction to the ceous sediment: 1. Mineralogical and chemical evi-
physics of rocks: Princeton University Press. dence: Geological Society of America Bulletin, 87, no.
Guéguen, Y., J. Sarout, J. Fortin, and A. Schubnel, 2009, 5, 725–737, http://dx.doi.org/10.1130/0016-7606(1976)​
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Cracks in porous rocks: Tiny defects, strong effects: 87<725:MOBMOA>2.0.CO;2.


The Leading Edge, 28, no. 1, 40–47, http://dx.doi. Hu, R., L. Vernik, L. Nayvelt, and A. Dicman, 2015, Seismic
org/10.1190/1.3064145. inversion for organic richness and fracture gradient
Han, D.-H., A. Nur, and D. Morgan, 1986, Effects of poros- in unconventional reservoirs: Eagle Ford Shale, Texas:
ity and clay content on wave velocities in sandstones: The Leading Edge, 34, no. 1, 80–84, http://dx.doi.org/​
Geophysics, 51, no. 11, 2093–2107, http://dx.doi.org/​ 10.1190/tle34010080.1.
10.1190/1.1442062. Hucka, V., and B. Das, 1974, Brittleness determination of
Hart, B. S., P. B. Flemings, and A. Deshpande, 1995, rocks by different methods: International Journal of
Porosity and pressure: Role of compaction disequilib- Rock Mechanics and Mining Sciences & Geomechanics
rium in the development of geopressures in a Gulf Abstracts, 11, no. 10, 389–392, http://dx.doi.org/10.​
Coast Pleistocene basin: Geology, 23, no. 1, 45–48, 1016/0148-9062(74)91109-7.
http://dx.doi.org/10.1130/0091-7613(1995)023​ Iijima, A., 1980, Geology of natural zeolites and zeolitic
<0045:PAPROC>2.3.CO;2. rocks: Pure and Applied Chemistry, 52, no. 9, 2115–
Hashin, Z., and S. Shtrikman, 1963, A variational approach 2130, http://dx.doi.org/10.1351/pac198052092115.
to the theory of the elastic behaviour of multiphase ma- Issler, D. R., 1992, A new approach to shale compaction and
terials: Journal of the Mechanics and Physics of Solids, stratigraphic restoration, Beaufort-Mackenzie Basin and
11, no. 2, 127–140, http://dx.doi.org/10.1016/0022-​ Mackenzie Corridor, northern Canada: AAPG Bulletin,
5096​(63)90060-7. 76, no. 8, 1170–1189.
Herron, M., 1987, Estimating the intrinsic permeability of Jaeger, J. C., and N. G. W. Cook, 1979, Fundamentals of
clastic sediments from geochemical data: 28th Annual rock mechanics, 3rd ed.: Chapman and Hall.
Logging Symposium, Society of Professional Well Log Jizba, D. L., 1991, Mechanical and acoustical properties of
Analysts, Transactions, SPWLA HH1-23. sandstones and shales: Ph.D. dissertation, Stanford
Herron, M., 1998, A robust permeability estimator for silici- University, https://pangea.stanford.edu/departments/
clastics: International Meeting on Petroleum Engineering, geophysics/dropbox/SRB/public/docs/theses/SRB_​
Society of Petroleum Engineers, SPE-49301. 045_MAR91_Jizba.pdf, accessed 13 January 2016.
Hill, R., 1963, Elastic properties of reinforced solids: Some Johnston, D. H., 1987, Physical properties of shale at tem-
theoretical principles: Journal of the Mechanics and perature and pressure: Geophysics, 52, no. 10, 1391–
Physics of Solids, 11, no. 5, 357–372, http://dx.doi. 1401, http://dx.doi.org/10.1190/1.1442251.
org/10.1016/0022-5096(63)90036-X. Johnston, J. E., and N. I. Christensen, 1995, Seismic anisot-
Holt, R. M., O.-M. Nes, and E. Fjaer, 2005, In-situ stress ropy of shales: Journal of Geophysical Research, 100,
dependence of wave velocities in reservoir and over- no. B4, 5991–6003, http://dx.doi.org/10.1029/​ 95JB​
burden rocks: The Leading Edge, 24, no. 12, 1268– 00031.
1274, http://dx.doi.org/10.1190/1.2149650. Kaarsberg, E. A., 1959, Introductory studies of natural and
Hood, A., C. C. M. Gutjahr, and R. L. Heacock, 1975, artificial argillaceous aggregates by sound-propaga-
Organic metamorphism and the generation of petro- tion and X-ray diffraction methods: The Journal of
leum: AAPG Bulletin, 59, no. 6, 986–996. Geology, 67, no. 4, 447–472, http://dx.doi.org/10.1086/​
Hornby, B. E., 1998, Experimental laboratory determina- 626597.
tion of the dynamic elastic properties of wet, drained Kachanov, M., 1992, Effective elastic properties of cracked
shales: Journal of Geophysical Research, 103, no. solids: Critical review of some basic concepts: Applied
B12, 29945–29964, http://dx.doi.org/10.1029/97JB0​ Mechanics Reviews, 45, no. 8, 304–335, http://dx.doi.
2380. org/10.1115/1.3119761.
Hornby, B. E., J. M. Howie, and D. W. Ince, 2003, Kachanov, M., I. Tsukrov, and B. Shafiro, 1994, Effective
Anisotropy correction for deviated-well sonic logs: moduli of solids with cavities of various shapes:
Application to seismic well tie: Geophysics, 68, no. 2, Applied Mechanics Reviews, 47, no. 1S, S151–S174,
464–471, http://dx.doi.org/10.1190/1.1567212. http://dx.doi.org/10.1115/1.3122810.
Hornby, B. E., L. M. Schwartz, and J. A. Hudson, 1994, Kassir, M. K., and A. M. Bregman, 1972, The stress-inten-
Anisotropic effective-medium modeling of the elastic sity factor for a penny-shaped crack between two dis-
properties of shales: Geophysics, 59, no. 10, 1570– similar materials: Journal of Applied Mechanics, 39,
1583, http://dx.doi.org/10.1190/1.1443546. no. 1, 308–310, http://dx.doi.org/10.1115/1.3422648.

09-Vernik_References.indd 202 12-08-2016 19:58:30


References 203

Kuster, G. T., and M. N. Toksöz, 1974, Velocity and attenu- Larionov, V. V., 1969, Radiometriya skvazhin [Radiometry
ation of seismic waves in two-phase media: Part I. of boreholes]: Nedra.
Theoretical formulations: Geophysics, 39, no. 5, 587– Lash, G. G., and T. Engelder, 2005, An analysis of horizon-
606, http://dx.doi.org/10.1190/1.1440450. tal microcracking during catagenesis: Example from
Katahara, K., 2003, Analysis of overpressure on the Gulf of the Catskill delta complex: AAPG Bulletin, 89, no. 11,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Mexico shelf: Offshore Technology Conference, OTC- 1433–1449, http://dx.doi.org/10.1306/05250504141.


15293-MS, http://dx.doi.org/10.4043/15293-MS. Laubie, H. H., 2013, Linear elastic fracture mechanics in
Katahara, K., 2009, Lateral earth stress and strain: 79th anisotropic solids: Application to fluid-driven crack
Annual International Meeting, SEG, Expanded propagation: M.S. Thesis, Massachusetts Institute of
Abstracts, 2165–2169, http://dx.doi.org/10.1190/1.32​ Technology.
55285. Lawn, B. R., 1993, Fracture of brittle solids: Cambridge
Katahara, K. W., 1996, Clay mineral elastic properties: 66th Solid State Science Series, 517, http://dx.doi.org/​
Annual International Meeting, SEG, Expanded Abstracts, 10.1017/CBO9780511623127
1691–1694, http://dx.doi.org/10.1190/1.1826​454. Lehner, F. K., 1991, Pore-pressure induced fracturing of
Kenter, J. A. M., H. Braaksma, K. Verwer, and X. M. T. van ­petroleum source rocks: Implications for primary mi-
Lanen, 2007, Acoustic behavior of sedimentary rocks: gration, in G. Imarisio, M. Frias, and J. M. Bemtgen,
Geologic properties versus Poisson’s ratios: The eds., European Oil and Gas Conference, proceedings:
Leading Edge, 26, no. 4, 436–444, http://dx.doi.org/​ Graham and Trotman, 131–141.
10.1190/1.2723206. Lekhnitsky, S. G., 1963, Theory of elasticity of an anisotro-
Keys, R. G., and S. Xu, 2002, An approximation for the pic elastic body: Holden-Day.
Xu-White velocity model: Geophysics, 67, no. 5,
­ Levin, V., M. Markov, and S. Kanaun, 2004, Effective field
1406–1414, http://dx.doi.org/10.1190/1.1512786. method for seismic properties of cracked rocks: Journal
Khadeeva, Y., and L. Vernik, 2014, Rock-physics model of Geophysical Research, 109, no. B8, B08202, http://
for unconventional shales: The Leading Edge, 33, no. dx.doi.org/10.1029/2003JB002795.
3, 318–322, http://dx.doi.org/10.1190/tle33030318.1. Lewan, M. D., 1987, Petrographic study of primary petro-
Kim, K. Y., K. H. Wrolstad, and F. Aminzadeh, 1993, leum migration in the Woodford Shale and related rock
Effects of transverse isotropy on P-wave AVO for gas units, in B. Doligez, ed., Migration of hydrocarbons in
sands: Geophysics, 58, no. 6, 883–888, http://dx.doi. sedimentary basins: Editions Technip, 113–130.
org/10.1190/1.1443472. Li, Y., 2006, An empirical method for estimation of aniso-
King, M. S., 1966, Wave velocities in rocks as a function of tropic parameters in clastic rocks: The Leading Edge, 25,
changes in overburden pressure and pore fluid satu- no. 6, 706–711, http://dx.doi.org/10.1190/1.2210052.
rants: Geophysics, 31, no. 1, 50–73, http://dx.doi. Lo, T. W., K. B. Coyner, and M. N. Toksöz, 1986,
org/10.1190/1.1439763. Experimental determination of elastic anisotropy of
Klimentos, T., and C. McCann, 1990, Relationships among Berea sandstone, Chicopee shale, and Chelmsford
compressional wave attenuation, porosity, clay con- granite: Geophysics, 51, no. 1, 164–171, http://dx.doi.
tent, and permeability in sandstones: Geophysics, 55, org/10.1190/1.1442029.
no. 8, 998–1014, http://dx.doi.org/10.1190/1.1442928. Løseth, H., L. Wensaas, M. Gading, K. Duffaut, and M.
Kowallis, B. J., L. E. A. Jones, and H. F. Wang, 1984, Springer, 2011, Can hydrocarbon source rocks be iden-
Velocity-porosity-clay content systematics of poorly tified on seismic data?: Geology, 39, no. 12, 1167–
consolidated sandstones: Journal of Geophysical 1170, http://dx.doi.org/10.1130/G32328.1.
Research, 89, no. B12, 10355–10364, http://dx.doi.org/​ Loucks, R. G., R. M. Reed, S. C. Ruppel, and D. M. Jarvie,
10.1029/JB089iB12p10355. 2009, Morphology, genesis, and distribution of nano-
Krief, M., J. Garat, J. Stellingwerff, and J. Ventre, 1990, A meter-scale pores in siliceous mudstones of the
petrophysical interpretation using the velocities of P Mississippian Barnett Shale: Journal of Sedimentary
and S waves (full-waveform sonic): The Log Analyst, Research, 79, no. 12, 848–861, http://dx.doi.org/​
31, 355–369. 10.2110/jsr.2009.092.
Kushch, V. I., I. Sevostianov, and L. Mishnaevsky, Jr., 2009, Loucks, R. G., and S. C. Ruppel, 2007, Mississippian
Effect of crack orientation statistics on effective stiff- Barnett Shale: Lithofacies and depositional setting of a
ness of microcracked solid: International Journal of deep-water shale-gas succession in the Fort Worth
Solids and Structures, 46, no. 6, 1574–1588, http:// Basin, Texas: AAPG Bulletin, 91, no. 4, 579–601,
dx.doi.org/10.1016/j.ijsolstr.2008.11.023. http://dx.doi.org/10.1306/11020606059.
Lakes, R., 1987, Foam structures with a negative Poisson’s Lucier, A. M., R. Hofmann, and L. T. Bryndzia, 2011,
ratio: Science, 235, no. 4792, 1038–1040, http://dx.doi. Evaluation of the variable gas saturation on acoustic
org/10.1126/science.235.4792.1038. PubMed. log data from the Haynesville Shale gas play, NW

09-Vernik_References.indd 203 12-08-2016 19:58:30


204 Seismic Petrophysics in Quantitative Interpretation

Louisiana, USA: The Leading Edge, 30, no. 3, 300– chemical constraints on petroleum migration: AAPG
311, http://dx.doi.org/10.1190/1.3567261. Continuing Education Course Notes Series No. 8,
Malik, M., J. M. Salazar, C. Torres-Verdin, G. L. Wang, B1–B60.
H. J. Lee, and K. Sepehrnoori, 2008, Effects of petro- Moran, J. H., and S. Gianzero, 1979, Effects of formation
physical properties on array-induction measurements anisotropy on resistivity-logging measurements: Geo­
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

acquired in the presence of oil-base mud-filtrate inva- physics, 44, no. 7, 1266–1286, http://dx.doi.org/10.​
sion: Petrophysics, 49, 1, 74–92. 1190/1.1441006.
Mallick, S., 2007, Amplitude-variation-with-offset, elastic- Mori, T., and K. Tanaka, 1973, Average stress in matrix and
impedance, and wave-equation synthetics –– A model- average elastic energy of materials with misfitting in-
ing study: Geophysics, 72, no. 1, C1–C7, http://dx.doi. clusions: Acta Metallurgica, 21, no. 5, 571–574, http://
org/10.1190/1.2387108. dx.doi.org/10.1016/0001-6160(73)90064-3.
Marion, D. P., 1990, Acoustical, mechanical, and transport Nelson, P. H., 1994, Permeability-porosity relationships in
properties of sediments and granular materials: Ph.D. sedimentary rocks: The Log Analyst, 35, no. 3, 38–62,
dissertation, Stanford University, Stanford Rock http://www.pe.tamu.edu/blasingame/data/z_zCourse_
Physics & Borehole Geophysics Project, 39, https:// Archive/P324_03A/Lecture_Refs_(pdf)/P324_Mod1_​
pangea.stanford.edu/departments/geophysics/dropbox/ 01_Nelson_(Log_Analyst_May_June_1994).pdf, ac-
SRB/public/docs/theses/SRB_039_FEB90_Marion. cessed 11 January 2016.
pdf, accessed 11 January 2016. Nordgren, R. P., 1972, Propagation of a vertical hydraulic
Mavko, G., and D. Jizba, 1991, Estimating grain-scale fluid fracture: Society of Petroleum Engineers Journal, 12,
effects on velocity dispersion in rocks: Geophysics, 56, no. 4, 306–314.
no. 12, 1940–1949, http://dx.doi.org/10.1190/1.1443005. Norris, A. N., 1985, A differential scheme for the effective
Mavko, G., T. Mukerji, and J. Dvorkin, 2009, The rock moduli of composites: Mechanics of Materials, 4, no.
physics handbook: Tools for seismic analysis of porous 1, 1–16, http://dx.doi.org/10.1016/0167-6636(85)900​
media, 2nd ed: Cambridge University Press, http:// 02-X.
dx.doi.org/10.1017/CBO9780511626753 Nur, A., 1992, The role of critical porosity in the physical
Mavko, G. M., and A. Nur, 1978, The effect of nonelliptical response of rocks: EOS, Transactions of the American
cracks on the compressibility of rocks: Journal of Geophysical Union, 73, no. 43, 66.
Geophysical Research, 83, no. B9, 4459–4468, http:// Nur, A., D. Marion, and H. Yin, 1991, Wave velocities
dx.doi.org/10.1029/JB083iB09p04459. in sediments, in J. M. Hovem, M. D. Richardson, and
McAuliffe, C. D., 1979, Oil and gas migration: Chemical R. D. Stoll, eds., Shear waves in marine sediments:
and physical constraints: AAPG Bulletin, 63, no. 5, Kluwer Academic Publishers, 131–140.
761–781. Nur, A., and G. Simmons, 1969, Stress-induced velocity
Meissner, F. F., 1984, Petroleum geology of the Bakken anisotropy in rocks: An experimental study: Journal of
Formation, Williston Basin, North Dakota and Geophysical Research, 74, no. 27, 6667–6674, http://
Montana, in G. Demaison and R. J. Murris, eds., dx.doi.org/10.1029/JB074i027p06667.
Petroleum geochemistry and basin evaluation, AAPG O’Connell, R. J., and B. Budiansky, 1974, Seismic veloci-
Memoir 35, 159–179. ties in dry and saturated cracked solids: Journal of
Melia, P. J., and R. L. Carlson, 1984, An experimental test of Geophysical Research, 79, no. 35, 5412–5426, http://
P-wave anisotropy in stratified media: Geophysics, 49, dx.doi.org/10.1029/JB079i035p05412.
no. 4, 374–378, http://dx.doi.org/10.1190/1.1441673. Okiongbo, K. S., A. C. Aplin, and S. R. Larter, 2005,
Militzer, B., H.-R. Wenk, S. Stackhouse, and L. Stixrude, Changes in type II kerogen density as a function of ma-
2011, First-principles calculation of the elastic moduli turity: Evidence from the Kimmeridge Clay Formation:
of sheet silicates and their application to shale anisot- Energy & Fuels, 19, no. 6, 2495–2499, http://dx.doi.
ropy: American Mineralogist, 96, no. 1, 125–137, org/10.1021/ef050194+.
http://dx.doi.org/10.2138/am.2011.3558. Ortega, J. A., 2010, Microporomechanical modeling of
Milliken, K., 2014, A compositional classification for grain shale: Ph.D. thesis, Massachusetts Institute of
assemblages in fine-grained sediments and sedimenta- Technology.
ry rocks: Journal of Sedimentary Research, 84, no. 12, Ostrander, W. J., 1984, Plane-wave reflection coefficients
1185–1199, http://dx.doi.org/10.2110/jsr.2014.92. for gas sands at nonnormal angles of incidence:
Mindlin, R. D., 1949, Compliance of elastic bodies in con- Geophysics, 49, no. 10, 1637–1648, http://dx.doi.
tact: Journal of Applied Mechanics, 16, 259–268. org/10.1190/1.1441571.
Momper, J. A., 1978, Oil migration limitations suggested Partyka, G., 2007, The birth of spectral decomposition: The
by geological and geochemical considerations, in Leading Edge, 26, no. 12, 1624–1624, http://dx.doi.
W. H. Roberts and R. J. Cordel, eds., Physical and org/10.1190/tle26121624.1.

09-Vernik_References.indd 204 12-08-2016 19:58:30


References 205

Passey, Q. R., S. Creaney, J. B. Kulla, F. J. Moretti, and Rickman, R., M. J. Mullen, J. E. Petre, W. V. Grieser, and
J. D. Stroud, 1990, A practical model for organic rich- D. Kundert, 2008, A practical use of shale petrophysics
ness from porosity and resistivity logs: AAPG Bulletin, for stimulation design optimization: All shale plays are
74, no. 12, 1777–1794. not clones of the Barnett Shale: Annual Technical
Paxton, S. T., J. O. Szabo, J. M. Ajdukiewicz, and R. E. Conference and Exhibition, Society of Petroleum
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Klimentidis, 2002, Construction of an intergranular Engineers, SPE 115258-MS, http://dx.doi.org/10.2118/​


volume compaction curve for evaluating and predicting 115258-MS.
compaction and porosity loss in rigid-grain sandstone Ross, C. P., 2016, Unbiased AVO crossplotting?: The
reservoirs: AAPG Bulletin, 86, no. 12, 2047–2067. Leading Edge, 35, no. 4, 338–344, http://dx.doi.org/​
Pendrel, J., H. Debeye, R. Pedersen-Tatalovic, B. Goodway, 10.1190/tle35040338.1.
J. Dufour, M. Bogaards, and R. R. Stewart, 2000, Rubey, W. W., and M. K. Hubbert, 1959, Role of fluid pres-
Estimation and interpretation of P and S impedance sure in mechanics of overthrust faulting II: Overthrust
volumes from simultaneous inversion of P-wave offset belt in geosynclinal area of western Wyoming in light
seismic data: 70th Annual International Meeting, SEG, of fluid-pressure hypothesis: Geological Society of
Expanded Abstracts, 146–149, http://dx.doi.org/​ 10.​ America Bulletin, 70, no. 2, 167–206, http://dx.doi.
1190/1.1815683. org/10.1130/0016-7606(1959)70[167:ROFPIM]2.0.
Pennington, W. D., 1997, Seismic petrophysics: An applied CO;2.
science for reservoir geophysics: The Leading Edge, 16, Rudnicki, M. D., 2015, Variation of organic matter density
no. 3, 241–246, http://dx.doi.org/10.1190/1.1437608. with thermal maturity: AAPG Bulletin, published on-
Prasad, M., K. C. Mba, T. Sadler, and M. L. Batzle, 2011, line 09 November 2015.
Maturity and impedance analysis of organic-rich shales: Rüger, A., 1998, Variation of P-wave reflectivity with
SPE Reservoir Evaluation & Engineering, 14, no. 5, ­offset and azimuth in anisotropic media: Geophysics,
533–543, http://dx.doi.org/10.2118/123531-​PA. 63, no. 3, 935–947, http://dx.doi.org/10.1190/1.​
Price, L. C., 1983, Geologic time as a parameter in organic 1444405.
metamorphism and vitrinite reflectance as an absolute Ruiz, F., and A. Cheng, 2010, A rock physics model for
paleogeothermometer: Journal of Petroleum Geology, 6, tight gas sand: The Leading Edge, 29, no. 12, 1484–
no. 1, 5–37, http://dx.doi.org/10.1111/j.1747-5457.1983. 1489, http://dx.doi.org/10.1190/1.3525364.
tb00260.x. Rutherford, S. R., and R. H. Williams, 1989, Amplitude-
Price, L. C., 1994, Basin richness and source rock disrup- versus-offset variations in gas sands: Geophysics, 54,
tion — A fundamental relationship?: Journal of no. 6, 680–688, http://dx.doi.org/10.1190/1.1442696.
Petroleum Geology, 17, no. 1, 5–38, http://dx.doi. Saenger, E. H., O. S. Krüger, and S. A. Shapiro, 2004,
org/10.1111/j.1747-5457.1994.tb00112.x. Effective elastic properties of randomly fractured soils:
Price, L. C., T. Ging, T. Daws, A. Love, M. Pawlewicz, and 3D numerical experiments: Geophysical Prospecting,
D. Anders, 1984, Organic metamorphism in the 52, no. 3, 183–195, http://dx.doi.org/10.1111/j.1365-​
Mississippian–Devonian Bakken shale, North Dakota 2478​.2004.00407.x.
portion of the Williston basin, in J. Woodward, F. F. Sayers, C., 2008, The elastic properties of carbonates: The
Meissner, and J. L. Clayton, eds., Hydrocarbon source Leading Edge, 27, no. 8, 1020–1024, http://dx.doi.
rocks of the greater Rocky Mountain region: Rocky org/10.1190/1.2967555.
Mountain Association of Geologists, 83–133. Sayers, C., 2010, Geophysics under stress: Geomechanical
Pride, S. R., J. G. Berryman, and J. M. Harris, 2004, Seismic applications of seismic and borehole acoustic waves:
attenuation due to wave-induced flow: Journal of SEG Distinguished Instructor Short Course, Distin­
Geophysical Research, 109, no. B1, B01201.1–​B012​ guished Instructor Series No. 13, http://dx.doi.
01.19. org/10.1190/1.9781560802129.
Quakenbush, M., B. Shang, and C. Tuttle, 2006, Poisson Sayers, C., 2013, The effect of kerogen on the elastic an-
impedance: The Leading Edge, 25, no. 2, 128–138, isotropy of organic-rich shales: Geophysics, 78, no. 2,
http://dx.doi.org/10.1190/1.2172301. D65–D74, http://dx.doi.org/10.1190/geo2012-0309.1.
Raymer, L. L., E. R. Hunt, and J. S. Gardner, 1980, An im- Sayers, C. M., K. Fisher, and J. J. Walsh, 2015, Sensitivity
proved sonic transit time-to-porosity transform: 21st of P- and S-impedance to the presence of kerogen in
Annual Logging Symposium, Society of Professional the Eagle Ford Shale: The Leading Edge, 34, no. 12,
Well Log Analysts, Transactions, 1–12. 1482–1486, http://dx.doi.org/10.1190/tle34121482.1.
Rice, J. R., 1992, Fault stress states, pore pressure distribu- Schlumberger, 2000, Log interpretation charts.
tions, and the weakness of the San Andreas fault, in B. Schmoker, J. W., 1979, Determination of organic content of
Evans and T.-F. Wong, eds., Fault mechanics and trans- Appalachian Devonian shales from formation-density
port properties of rock: Academic Press, 475–503. log: AAPG Bulletin, 63, no. 9, 1504–1537.

09-Vernik_References.indd 205 12-08-2016 19:58:30


206 Seismic Petrophysics in Quantitative Interpretation

Scott, P. A., T. Engelder, and J. J. Mecholsky, Jr., 1992, The Skempton, A. W., 1969, The consolidation of clays by
correlation between fracture-toughness anisotropy and gravitational compaction: Quarterly Journal of The
­
crack-surface morphology of siltstones in the Ithaca Geological Society of London, 125, no. 1–4, 373–411,
Formation, Appalachian Basin, in B. Evans and T.-F. http://dx.doi.org/10.1144/gsjgs.125.1.0373.
Wong, eds., Fault mechanics and transport properties Slatt, R. M., E. V. Eslinger, and S. K. Van Dyke, 2009,
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

of rocks: Academic Press, 341–370. Acoustic and petrophysical properties of a clastic deep-
Sevostianov, I., and M. Kachanov, 2001, Authors’ response: water depositional system from lithofacies to architec-
Journal of Biomechanics, 34, no. 5, 709–710, http:// tural elements’ scales: Geophysics, 74, no. 2, WA35–
dx.doi.org/10.1016/S0021-9290(00)00208-6. WA50, http://dx.doi.org/10.1190/1.3073760.
Sevostianov, I., and M. Kachanov, 2008, On approximate Smith, G. C., and P. M. Gidlow, 1987, Weighted stacking
symmetries of the elastic properties and elliptic orthot- for rock property estimation and detection of gas:
ropy: International Journal of Engineering Science, Geophysical Prospecting, 35, no. 9, 993–1014, http://
46, no. 3, 211–223, http://dx.doi.org/10.1016/j.ijengsci​ dx.doi.org/10.1111/j.1365-2478.1987.tb00856.x.
.2007.11.003. Smith, G. C., and R. A. Sutherland, 1996, The fluid factor
Sevostianov, I., and M. Kachanov, 2011, Elastic fields gen- as an AVO indicator: Geophysics, 61, no. 5, 1425–
erated by inhomogeneities: Far-field asymptotics, its 1428, http://dx.doi.org/10.1190/1.1444067.
shape dependence and relation to the effective elastic Smith, T. M., 2011, Practical seismic petrophysics: The
properties: International Journal of Solids and effective use of log data for seismic analysis: The
­
Structures, 48, no. 16–17, 2340–2348, http://dx.doi. Leading Edge, 30, no. 10, 1128–1141, http://dx.doi.
org/10.1016/j.ijsolstr.2011.04.014. org/10​.1190/1.3657071.
Sevostianov, I., M. Kachanov, and T. Zohdi, 2008, On com- Smith, T. M., C. M. Sayers, and C. H. Sondergeld, 2009,
putation of the compliance and stiffness contribution Rock properties in low-porosity/low permeability
tensors of non-ellipsoidal inhomogeneities: Interna­ sandstones: The Leading Edge, 28, no. 1, 48–59, http://
tional Journal of Solids and Structures, 45, no. 16, dx.doi.org/10.1190/1.3064146.
4375–4383, http://dx.doi.org/10.1016/j.ijsolstr​.2008.​ Sondergeld, C. H., R. J. Ambrose, C. S. Rai, and J.
03.020. Moncrieff, 2010, Micro-structural studies of gas shales:
Sevostianov, I., N. Yilmaz, V. Kushch, and V. Levin, 2005, Unconventional Gas Conference, Society of Petroleum
Effective elastic properties of matrix composites with Engineers, SPE 131771, http://dx.doi.org/10.2118/​
transversely-isotropic phases: International Journal of 131771-MS.
Solids and Structures, 42, no. 2, 455–476, http://dx.doi. Sondergeld, C. H., C. S. Rai, R. W. Margesson, and K. J.
org/10.1016/j.ijsolstr.2004.06.047. Whidden, 2000, Ultrasonic measurement of anisotropy
Shapiro, S., 2003, Elastic piezosensitivity of porous and on the Kimmeridge Shale: 70th Annual International
fractured rocks: Geophysics, 68, no. 2, 482–486, http:// Meeting, SEG, Expanded Abstracts, 1858–1861, http://
dx.doi.org/10.1190/1.1567215. dx.doi.org/10.1190/1.1815791.
Sharp, J. M. Jr., and P. A. Domenico, 1976, Energy trans- Sone, H., 2012, Mechanical properties of shale gas reser-
port in thick sequences of compacting sediment: voir rocks and its relation to the in-situ variation ob-
Geological Society of America Bulletin, 87, no. 3, served in shale gas reservoirs: Ph.D. dissertation,
390–400, http://dx.doi.org/10.1130/0016-7606(1976)​ Stanford University, https://pangea.stanford.edu/de-
87<390:ETITSO>2.0.CO;2. partments/geophysics/dropbox/SRB/public/docs/the-
Shuey, R. T., 1985, A simplification of the Zoeppritz equa- ses/SRB_128_MAR12_Sone.pdf, accessed 12 January
tions: Geophysics, 50, no. 4, 609–614, http://dx.doi. 2016.
org/10.1190/1.1441936. Sonnenberg, S. A., J. A. LeFever, and R. J. Hill, 2011,
Shumaker, N., and L. Vernik, 2009, The use of “vertical- Fracturing in the Bakken petroleum system, Williston
ized” stacking velocities to constrain shale properties Basin, in J. W. Robinson, J. A. LeFever, and S. B.
in West Africa: The Leading Edge, 28, no. 2, 184–188, Gaswirth, eds., The Bakken-Three Forks Petroleum
http://dx.doi.org/10.1190/1.3086053. System in the Williston Basin: Rocky Mountain
Sinha, B. K., E. Simsek, and Q. H. Liu, 2006, Elastic-wave Association of Geologists, 393–417.
propagation in deviated wells in anisotropic forma- Strandenes, S., and J. P. Blangy, 1991, Ultrasonic velocity
tions: Geophysics, 71, no. 6, D191–D202, http://dx.doi. measurements in Troll sandstones: Stanford Rock
org/10.1190/1.2358402. Physics Project, 47 A1–A54.
Skelt, C., 2004, Fluid substitution in laminated sands: The Talukdar, S., O. Gallango, C. Vallejos, and A. Ruggiero,
Leading Edge, 23, no. 5, 485–493, http://dx.doi. 1987, Observations on the primary migration of oil in
org/10.1190/1.1756839. the La Luna source rocks of the Maracaibo Basin,

09-Vernik_References.indd 206 12-08-2016 19:58:30


References 207

Venezuela, in B. Doligez, ed., Migration of hydrocar- Van Dok, R., B. Fuller, L. Engelbrecht, and M. Sterling,
bons in sedimentary basins: Editions Technip, 59–78. 2011, Seismic anisotropy in microseismic event loca-
Tarasov, B. G., 2011, Universal scale of brittleness for rocks tion analysis: The Leading Edge, 30, no. 7, 766–770,
failed at compression: 13th International Conference, http://dx.doi.org/10.1190/1.3609091.
International Association for Computer Methods and Vanorio, T., T. Mukerji, and G. Mavko, 2008, Emerging
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Advances in Geomechanics, Proceedings, 669–673. methodologies to characterize the rock physics proper-
Terzaghi, K., and R. B. Peck, 1948, Soil mechanics in engi- ties of organic-rich shales: The Leading Edge, 27, no.
neering practice: John Wiley & Sons, Inc. 6, 780–787, http://dx.doi.org/10.1190/1.2944163.
Thomas, E. C., and S. J. Stieber, 1975, The distribution of Vasin, R. N., H.-R. Wenk, W. Kanitpanyacharoen, S.
shale in sandstones and its effect upon porosity: 16th Matthies, and R. Wirth, 2013, Elastic anisotropy mod-
Annual Logging Symposium, Society of Professional eling of Kimmeridge Shale: Journal of Geophysical
Well Log Analysts, SPWLA-1975-T. Research, 118, no. 8, 3931–3956.
Thomsen, L., 1986, Weak elastic anisotropy: Geophysics, Vavakin, A. S., and R. L. Salganik, 1975, Effective charac-
51, no. 10, 1954–1966, http://dx.doi.org/10.1190/​ 1.​ teristics of nonhomogeneous media with isolated inho-
1442051. mogeneities: Mechanics of Solids, 10, 58–66 [English
Thomsen, L., 1990, Poisson was not a geophysicist: The translation from Izvestia AN SSSR: Mekhanika
Leading Edge, 9, no. 12, 27–29, http://dx.doi. Tverdogo Tela].
org/10.1190/1.1439706. Verm, R., and F. Hilterman, 1995, Lithology color-coded
Timur, A., 1968, An investigation of permeability, porosity, seismic sections: The calibration of AVO crossplotting
and residual water saturation relationships for sand- to rock properties: The Leading Edge, 14, no. 8, 847–
stone reservoirs: The Log Analyst, 9, no. 4, 8–17. 853, http://dx.doi.org/10.1190/1.1437170.
Tissot, B. P., and R. Pelet, 1971, Nouvelles donées sur les Vernik, L., 1994, Hydrocarbon generation-induced micro-
mécanismes de genèse et de migration du pétrole, simu- cracking of source rocks: Geophysics, 59, no. 4, 555–
lation mathématique et application a la prospection: 8th 563, http://dx.doi.org/10.1190/1.1443616.
World Petroleum Congress, Proceedings, 2, 35–46. Vernik, L., 1998, Acoustic velocity and porosity systemat-
Tissot, B. P., and D. H. Welte, 1978, Petroleum formation ics in siliciclastics: The Log Analyst, 39, 27–35.
and occurrence: Springer, http://dx.doi.org/10.1007/​ Vernik, L., 2000, Permeability prediction in poorly consoli-
978-​3-642-96446-6 dated siliciclastics based on porosity and clay volume
Tosaya, C., and A. Nur, 1982, Effects of diagenesis and logs: Petrophysics, 41, no. 2, 138–147.
clays on compressional velocities in rocks: Geophysical Vernik, L., 2008, Anisotropic correction of sonic logs in
Research Letters, 9, no. 1, 5–8, http://dx.doi.org/10​.​ wells with large relative dip: Geophysics, 73, no. 1,
1029/GL009i001p00005. E1–E5, http://dx.doi.org/10.1190/1.2789776.
Tosaya, C. A., 1982, Acoustical properties of clay-bearing Vernik, L., 2011, Unified model for continuous pore pres-
rocks: Ph.D. dissertation, Stanford University, https:// sure prediction in shale: 45th US Rock Mechanics/
pangea.stanford.edu/departments/geophysics/dropbox/ Geomechanics Symposium, American Rock Mechanics
SRB/public/docs/theses/SRB_015_JUN82_Tosaya. Association, conference paper ARMA-11-444.
pdf, accessed 12 January 2016. Vernik, L., D. Fisher, and S. Bahret, 2001, Estimation of
Traugott, M., 1997, Pore pressure and fracture pressure de- net-to-gross from P and S impedance: Part II: 3D seis-
terminations in deepwater: World Oil: Deepwater mic inversion: 71st Annual International Meeting,
Technology Special Supplement, 218, 68–70, http:// SEG, Expanded Abstracts, 211–214, http://dx.doi.org/​
www.hxrdrillingservices.com/wp-content/uploads/ 10.1190/1.1816568.
WorldOil.pdf, accessed 12 January 2016. Vernik, L., D. Fisher, and S. Bahret, 2002, Estimation of
Tsvankin, I., 2001, Seismic signatures and analysis of re- net-to-gross from P and S impedance in deepwater tur-
flection data in anisotropic media: Handbook of bidites: The Leading Edge, 21, no. 4, 380–387, http://
Geophysical Exploration: Seismic Exploration, 29, dx.doi.org/10.1190/1.1471602.
Pergamon, Elsevier. Vernik, L., and J. Hamman, 2009, Stress sensitivity of sand-
Tsvankin, I., and L. Thomsen, 1994, Nonhyperbolic reflec- stones and 4D applications: The Leading Edge, 28, no.
tion moveout in anisotropic media: Geophysics, 59, no. 1, 90–93, http://dx.doi.org/10.1190/1.3064152.
8, 1290–1304, http://dx.doi.org/10.1190/1.1443686. Vernik, L., S. Hickman, D. Lockner, and M. Rusanov, 1994,
Van Baaren, J. P., 1979, Quick-look permeability estimates Ultrasonic velocities in cores from the Kola superdeep
using sidewall samples and porosity logs: 6th Annual well and the nature of subhorizontal seismic reflec-
European Logging Symposium, Society of Professional tions: Journal of Geophysical Research, 99, no. B12,
Well Log Analysts, Transactions, 19–25. 24209–24219, http://dx.doi.org/10.1029/94JB01236.

09-Vernik_References.indd 207 12-08-2016 19:58:30


208 Seismic Petrophysics in Quantitative Interpretation

Vernik, L., and M. Kachanov, 2010, Modeling elastic Widess, M. B., 1973, How thin is a thin bed?: Geophysics,
­properties of siliciclastic rocks: Geophysics, 75, no. 6, 38, no. 6, 1176–1180, http://dx.doi.org/10.1190/​ 1.​
E171–E182, http://dx.doi.org/10.1190/1.3494031. 1440403.
Vernik, L., and M. Kachanov, 2012, On some controversial Winkler, K. W., 1983, Frequency dependent ultrasonic
issues in rock physics: The Leading Edge, 31, no. 6, properties of high-porosity sandstones: Journal of
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

636–642, http://dx.doi.org/10.1190/tle31060636.1. Geophysical Research, 88, no. B11, 9493–9499, http://


Vernik, L., and C. Landis, 1996, Elastic anisotropy of dx.doi.org/10.1029/JB088iB11p09493.
source rocks: Implications for hydrocarbon generation Wood, D. M., 1990, Soil behaviour and critical state soil
and primary migration: AAPG Bulletin, 80, no. 4, mechanics: Cambridge University Press.
531–544. Wyllie, M. R. J., A. R. Gregory, and L. W. Gardner, 1956,
Vernik, L., and X. Liu, 1997, Velocity anisotropy in shales: Elastic wave velocities in heterogeneous and porous
A petrophysical study: Geophysics, 62, no. 2, 521–532, media: Geophysics, 21, no. 1, 41–70, http://dx.doi.org/​
http://dx.doi.org/10.1190/1.1444162. 10.1190/1.1438217.
Vernik, L., and J. Milovac, 2011, Rock physics of organic Wyllie, M. R. J., A. R. Gregory, and L. W. Gardner, 1958,
shales: The Leading Edge, 30, no. 3, 318–323, http:// An experimental investigation of factors affecting elas-
dx.doi.org/10.1190/1.3567263. tic-wave velocities in porous media: Geophysics, 23,
Vernik, L., and A. Nur, 1992a, Ultrasonic velocity and an- no. 3, 459–493, http://dx.doi.org/10.1190/1.1438493.
isotropy of hydrocarbon source rocks: Geophysics, 57, Xu, S., and R. E. White, 1995, A new velocity model for
no. 5, 727–735, http://dx.doi.org/10.1190/1.1443286. clay-sand mixtures: Geophysical Prospecting, 43, no.
Vernik, L., and A. Nur, 1992b, Petrophysical classification 1, 91–118, http://dx.doi.org/10.1111/j.1365-2478.1995.
of siliciclastics for lithology and porosity prediction tb00126.x.
from seismic velocities: AAPG Bulletin, 76, no. 9, Yale, L. B., 1984, Rock catalog: Stanford rock physics:
1295–1309. Stanford University.
Vernik, L., M. D. Zoback, and M. Brudy, 1992, Methodology Zimmerman, R. W., 1986, Compressibility of two-dimen-
and application of the wellbore breakout analysis in es- sional cavities of various shapes: Journal of Applied
timating the maximum horizontal stress magnitude in Mechanics, 53, no. 3, 500–504, http://dx.doi.org/10​.​
the KTB pilot hole: Scientific Drilling, 3, 161–169. 1115/1.3171802.
Walls, J. D., and S. W. Sinclair, 2011, Eagle Ford Shale Zimmerman, R. W., 1991, Compressibility of sandstones:
reservoir properties from digital rock physics: First Elsevier.
Break, 29, no. 6, 97–101. Zoback, M. D., 2007, Reservoir geomechanics: Cambridge
Walsh, J. B., 1965, The effect of cracks on the compress- University Press. http://dx.doi.org/10.1017/CBO97​
ibility of rock: Journal of Geophysical Research, 70, 80511586477
no. 2, 381–389, http://dx.doi.org/10.1029/JZ070i002​ Zoback, M. D., R. Apel, J. Baumgärtner, M. Brudy, R.
p00381. Emmermann, B. Engeser, K. Fuchs, W. Kessels, H.
Wang, Z., 2002a, Seismic anisotropy in sedimentary rocks, Rischmüller, F. Rummel, and L. Vernik, 1993, Upper-
part 1: A single plug laboratory method: Geophysics, crustal strength inferred from stress measurements to
67, no. 5, 1415–1422, http://dx.doi.org/10.1190/​ 1.​ 6 km depth in the KTB borehole: Nature, 365, no.
1512787. 6447, 633–635, http://dx.doi.org/10.1038/365633a0.
Wang, Z., 2002b, Seismic anisotropy in sedimentary rocks, Zoback, M. D., and J. H. Healy, 1984, Friction, faulting, and
part 2: Laboratory data: Geophysics, 67, no. 5, 1423– “in-situ” stresses: Annales Geophysicae, 2, 689–698.
1440, http://dx.doi.org/10.1190/1.1512743. Zoeppritz, K., 1919, Erdbebenwellen VII, VII B, Über
Wenk, H. R., I. Lonardelli, H. Franz, K. Nihei, and S. Reflexion und Durchgang seismischer Wellen durch
Nakagawa, 2007, Preferred orientation and elastic an- Unstetigkeitsflächen [On the reflection and transmis-
isotropy of illite-rich shale: Geophysics, 72, no. 2, sion of seismic waves at surfaces of discontinuity]:
E69–E75, http://dx.doi.org/10.1190/1.2432263. Nachrichten von der Königlichen Gesellschaft der
White, J. E., 1965, Seismic waves: Radiation, transmission, Wissenschaften zu Göttingen, Mathematisch-
and attenuation: McGraw-Hill. physikalische Klasse, 66–84.

09-Vernik_References.indd 208 12-08-2016 19:58:31


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Index

A anisotropic Backus model, 106, 184 C


anisotropic correction of sonic logs, 23.
accurate low-frequency model, 185 see also density log centroid effect, 31
acoustic-impedance-versus-shear- anisotropy parameters, 24 chemical diagenesis, 57, 65
impedance (AI-SI), 89, 105 full-wave processing of dipole sonic Class IIp, 90, 96
AVO classification, 89, 90 logs, 25 AVO curves, 99
from AI-SI template to synthetic-gather lithology color-coding, 27 AVO responses, 97, 98
models, 91 rock properties of sands, 25 clay-free components, 2
lithology and fluid on prestack attributes SH-wave velocity surface, 25–26 compressional-wave velocity, 67
effects, 118–119 anisotropic mudrocks. see organic compressive strength test, 180
log editing and AI-SI template, mudrocks Consolidated Sandstone Model, 59,
105–108 static and dynamic elastic parameters in, 61, 80
modeling N/G, 109–111 169–171 consolidated sandstones, 58
seismic N/G computation from anisotropy crack effects, 61
simultaneous inversion, 111–117 anisotropy-corrected acoustic– dry frame moduli, 58
AI-SI. see acoustic-impedance-versus- impedance, 112 dry P-wave and shear moduli vs.
shear-impedance (AI-SI ) core measurements, 137–139 porosity, 60
AI-SI inversion, 111 effect of kerogen, 144–153 in situ logging-data analysis, 59
arbitrary lines, 112–114 intrinsic velocity, 141–144 modified HSb models, 59
horizon amplitude maps, 112 parameters, 24 shape factors, 59
NMO-corrected synthetic angle, 113 prediction in organic and conventional consolidation porosity, 57–58, 61, 63, 65,
of prestack seismic data, 191, 192 shales, 174–175 70–72, 102, 109
AI-SI relationships, 161–165 stress dependence, 153–155 contact models, 51–52
AI-SI crossplot of log data, 165 aspect ratio effect, 46 continuum-mechanics-based theory, 56
bedding-normal core measurements, 163 normalized saturated bulk conventional shales, 66–69
bedding-normal VP-versus-Vs modulus, 47 core/log database, application to, 4
template, 164 normalized spheroidal pore compressional velocity vs.
in siliciclastics, 76–78 compliances, 46 porosity, 4, 6
vertical P- and S-wave velocity asymmetric TOC distribution, 186 density of major siliciclastic rock, 5
logs, 162 AVO analysis. see amplitude variation with histogram of volumetric clay
AI-SI template, 78–80, 105, 110 offset analysis (AVO analysis) content, 5
additional log editing, 107 microporosity, 4
anisotropy correction of sonic velocities, P-wave and shear moduli, 7
106–107 B solid-matrix properties, 8
crossplot, 107–108 XRD data, 5–6
Aki-Richards approximation, 88 background trend, 91, 95, 100, 102, core measurements of velocity and
amplitude variation with offset analysis 108, 118 anisotropy, 137–139
(AVO analysis), 19, 43, 84, 87 Backus averaging technique, 24, bulk-density vs. TOC values, 139
class III and zero-gradient case, 91–93 109, 110 P-wave and SH-wave velocities vs.
classes, 109 Backus inversion, 196 confining pressure, 138
classification in AI-SI space, 89 bedding plane, 141, 180 three-plug principle, 137
fluid factor in prospect risk mitigation, bedding-subparallel microcracks, 177 crack(s)
100–101 Biot flow. see viscoelastic equilibration aspect ratios, 47
from AI-SI template to synthetic-gather brine-saturated P-wave velocity, 22 crack-closure stress, 52
models, 91 brine-saturated sandstone density, 22 density, 46, 49, 55, 58
linearized AVO equations and features, brittleness, 169 in dry and fluid-saturated rocks, 46–47
87–89 bulk clay volume, 8 porosity, 46
tuning effects in AVO synthetic bulk density, 89 cracks and pores, mixture of, 47
modeling, 101–104 expression, 132 ellipsoidal pore-shape factors, 48
VTI anisotropy effect, 97–99 bulk-modulus equation, 44 pore-shape factors, 48

209

10-Vernik_Index.indd 209 12-08-2016 19:58:18


210 Seismic Petrophysics in Quantitative Interpretation

critical-porosity, 1, 60 elasticity-based brittleness index, 171 I


model, 4 ellipsoidal shape, 47
range in young sediments, 57 empirical model, 32 IGV. see intergranular volume (IGV)
threshold, 57 engineering parameters, 167 illite-clay platelets, 127
crossplot, 107–108 initial intergranular porosity, 56
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

crushed-core technique, 139 intergranular volume (IGV), 2, 56


F intrinsic velocity and anisotropy, 141–144
database of physical properties, 142–144
D FE-SEM images. see field-emission
scanning-electron-microscopy images
DEM model. see differential-effective- (FE-SEM images) K
medium model (DEM model) FIB-SEM. see focused ion beam-scanning
density log, 12. see also anisotropic electron microscopy (FIB-SEM) kerogen
correction of sonic logs field-emission scanning-electron- Backus model, 152
conditioning, 20–22 microscopy images (FE-SEM bedding-parallel and bedding-normal,
mud-invasion correction, 22–23 images), 127 150
density model for stress computation, “fizz-gas” reservoir, 101 effect on velocities and anisotropy,
31–32 flow-zone indicator (FZI), 17 144–153
depth of investigation (DOI), 23 fluid factor, 87, 92, 100 elastic properties, 149
DFITs. see diagnostic fracture-injection in prospect risk mitigation, 100–101 intrinsic elastic stiffnesses, 148, 151
tests (DFITs) fluid-saturation effect, 43–45, 86 kerogen content, 151
DHI. see direct hydrocarbon indication fluid substitution, 78–80 substitution, 189
(DHI) fluid vectors, 118 ultrasonic velocity and anisotropy
diagnostic fracture-injection tests (DFITs), focused ion beam-scanning electron measurements, 146–148
173, 178 microscopy (FIB-SEM), 121, 129 velocity measurements, 150
differential-effective-medium model (DEM 4D modeling of sands and sandstones, kerogen-fraction log, 131
model), 157, 159 80–86 Kozeny-Carman equation, 14
direct hydrocarbon indication (DHI), 87, 91 full-wave processing of dipole sonic Kuster-Toksöz equations, 50
DOI. see depth of investigation (DOI) logs, 25
drained bulk modulus, 40 FZI. see flow-zone indicator (FZI)
drained rock-frame moduli, 45–46 L
contact models, 51–52
cracks in dry and fluid-saturated rocks, G laminar limestone fraction, 196
46–47 LEFM. see linear elastic fracture mechanics
effective-stress, 52–56 gamma-ray index, 8 (LEFM)
effects of pore/crack interactions, 49–51 gamma-ray-log (GR), 8 levels of organic metamorphism
mixture of cracks and pores, 47–49 Gas Research Institute protocol (GRI (LOM), 124
drained saturated crack compliances, 46 protocol), 139 linear elastic fracture mechanics
dry volume fraction, 9 Gassmann’s relations, 43, 44, 47, 55, (LEFM), 179
62, 78 linear-regression
GR. see gamma-ray-log (GR) equations, 6
E grain-supported texture, 56 model, 74
Greenberg-Castagna equations, 74 linearized AVO equations and features, 87
Eagle Ford core data set, 183 GRI protocol. see Gas Research Institute lithologic parameters, 8
Eagle Ford marl, 127, 128, 151, 159–161, protocol (GRI protocol) fragments of turbidite reservoir, 10
169–171, 174, 179, 188, 194, 196 group velocity hydrogen index, 11
effective stress, 29, 31, 52 P-wave phase velocity and, 140 normalized GR histograms, 9
bedding-normal compressional–wave phase velocity vs., 139–141 lithology
velocity models, 36 snapshot of numerical simulation of and fluid on prestack attributes effects,
crack-density parameter, 55 P-wave propagation, 141 118–119
elastic-contact stiffening, 53 color-coding, 27
shale VP-versus-porosity model, 35 vectors, 118
sonic velocity vs., 34 H local squirt flow. see viscoelastic
VP and Vs values in dry, 54 equilibration
effective-stress tensor half-space models, 87 log-based N/G evaluation, 105
alternative methods of fracture-gradient harmonic-averaging equation, 44 log-based permeability prediction, 14
computation, 41 heuristic exponential model, 54 log editing, 20, 105
and Shmin gradient, 40 heuristic model, 34 additional log editing, 107
Hooke’s law, 40 HI. see hydrogen index (HI) anisotropic correction of sonic logs,
Mohr’s circles for different stress high thermal-expansion coefficient, 177 23–27
states, 41 higher-porosity muds, 30 anisotropy correction of sonic velocities,
normal-faulting stress regime, 42 Hooke’s law, 40 106–107
elastic-impedance logs, 112 hydrocarbons, 105 crossplot, 107–108
elastically isotropic pore, 45 hydrogen index (HI), 11, 122, 125 density log, 20–23

10-Vernik_Index.indd 210 12-08-2016 19:58:18


 Index 211

sonic log, 20 crack-density-tensor components, 158 estimation of kerogen density from


log model for unconventional shales, 129 NIA models of anisotropic, 156 vitrinite reflectance, 126
kerogen-fraction log, 131 P-wave and SV-wave phase-velocity lithology, solid-grain density, TOC, and
microstructural observations, 129–131 surfaces, 158 maturity data, 123–124
model applications, 134–137 MPa. see megapascals (MPa) organic shale maturation and
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

total porosity and kerogen porosity, mud-filtrate-saturation profile, 13 hydrocarbon generation, 124
131–133 mud-invasion correction, 22–23 plot of HI vs. vitrinite reflectance, 125
water saturation, 133–134 TOC content vs. rockeval HI, 125
logging while drilling (LWD), 17 organic shales
LOM. see levels of organic N brittleness estimation from elastic
metamorphism (LOM) properties, 168–169
low-frequency background model, 106 nano-computerized tomography elastic-property relationships in TI
low-porosity sandstone, 56 (nano-CT), 129 shales, 167–168
lower-impedance dipole, 92 NCT model. see normal–hydrostatic– elasticity-based brittleness of organic
LWD. see logging while drilling (LWD) compaction–trend model (NCT model) mudrocks, 171–173
near-incidence reflectivity, 88 geomechanical elastic parameters, 167
net-to-gross curve, 9 geomechanics of maturation-induced
M net-to-gross ratios (N/G ratios), 101, 105 microcracking, 176–182
and AI-SI inversion, 105 Shmin stress-gradient estimation,
maturation-induced microcracking lithology and fluid on prestack attributes 173–176
geomechanics, 176 effects, 118–119 static and dynamic elastic parameters in
effective crack-closure stress vs. crack log editing and AI-SI template, anisotropic mudrocks, 169–171
aspect ratio, 181 105–108 orientation distribution function (ODF),
effective stress state, 181 modeling, 109–111 68, 175
excess pore pressure, 180 N/G mapping, 114–117 overburden gradient (OBG), 178
fracture toughness of kerogen, 179–180 seismic N/G computation from overburden stress. see vertical
fracturing of source rocks, 177–178 simultaneous inversion, 111–117 effective stress
frictional equilibrium, 178–179 N/G ratios. see net-to-gross ratios OWC. see oilwater contact (OWC)
hypothetical bottomhole-stress- (N/G ratios)
concentration effect, 177 NIA. see noninteraction
microcracks, 177 approximation (NIA) P
stress-relief and cooling-mechanism NMO. see normal moveout (NMO)
explanations, 177 noninteraction approximation (NIA), 45, PDF. see probability density function
tensile-strength anisotropy, 180 48, 49, 155, 158 (PDF)
total-stress state, 178 normal-faulting stress regime, 42 permeability, 7, 13
mechanical diagenesis, 61, 65 normal-hydrostatic-compaction-trend nanodarcy, 179
megapascals (MPa), 33 model (NCT model), 31 prediction in siliciclastics, 13–17
mg/gOC. see milligrams per gram of normal moveout (NMO), 19 reservoir, 168
organic carbon (mg/gOC) petrographic data and petrophysical
microcracks, 177 classification, 1
microporosity, 4 O clay-free components, 2
microstructural observations, 129–131 grain-supported sands and sandstones, 2
milligrams per gram of organic carbon OBG. see overburden gradient (OBG) sorting, 2–3
(mg/gOC), 122 OBM. see oil-based mud (OBM) texture of siliciclastics, 2
mislabeling, 43 ODF. see orientation distribution TOC, 3–4
mode I fracture toughness, 179 function (ODF) turbidite sandstone cores, 3
model applications, 134–137 oil-based mud (OBM), 13, 70 petrologic data in unconventional
Eagle Ford well, log model applied oilwater contact (OWC), 104 shales, 121
in, 134 organic-matter transformation, 130 organic richness and thermal maturity,
Haynesville well, log model applied organic mudrocks, 121, 122. see also 122–126
in, 136 anisotropic mudrocks rock composition, 121–122
Woodford well, log model applied bulk-density-versus-TOC rock texture, 126–129
in, 135 relationship, 132 petrophysical log modeling, 191
modeling N/G, 109 classification scheme, 122 petrophysical model, 131
AI-SI template, 110 elasticity-based brittleness of, 171–173 building, 8
P-wave reflection coefficient vs., incident formations, 168 lithologic parameters, 8–11
angle, 109 intrinsic anisotropy parameters of, 151 total porosity, 11–13
plane-wave synthetic angle gathers solid organic matter in, 122 phase velocity
gas-sand reservoir, 111 textural characteristics, 137 group velocity vs., 139–141
porosity and impedances of gas unconventional properties, 162 P-wave group-velocity and, 140
sand/shale pairs, 109 organic richness and thermal maturity, snapshot of numerical simulation of
modeling of conventional shale, 56 122–126 P-wave propagation, 141
modeling stress dependence, 155–158 empirical relationship between measured PI. see Poisson impedance (PI)
Bakken shale core, 157 TOC, 126 Poisson impedance (PI), 108

10-Vernik_Index.indd 211 12-08-2016 19:58:19


212 Seismic Petrophysics in Quantitative Interpretation

Poisson’s ratio, 40, 45, 89, 167 rock physics, 43 seismic ties, 184–186
Poisson’s reflectivity, 88 core measurements of velocity and Shmin stress–gradient mapping, 194–198
polar anisotropy. see transverse anisotropy, 137–139 TOC estimation from AI-SI template,
isotropy (TI) drained rock-frame moduli, 45–46 186–190
polarity reversal, 90, 96, 98 effect of kerogen on velocities and TOC mapping, 191–194
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

pore-pressure equilibration, 43 anisotropy, 144–153 in unconventional shales, 183


pore-shape factor, 57 fluid-saturation effect, 43–45, 86 seismic detection limit, 102
pore/crack interactions, effects of, 49 intrinsic velocity and anisotropy, 141–144 seismic forward modeling, 18
comparison of dry frame P-wave modeling in sand/shale sequences, 56–72 seismic N/G computation from
modulus, 51 modeling stress dependence, 155–158 simultaneous inversion
Mori-Tanaka model, 49 of unconventional shales, 137 AI-SI inversion, 111–114
pore-shape factors, 50 phase velocity vs. group velocity, 139–141 sand volume computation and net/gross
water-saturated VP-versus-porosity rock-physics model, 158–161 mapping, 114–117
space, 49 science of, 18 seismic petrophysics, 18, 19–20
pore pressure stress dependence of velocity and engine, 43
and stress state, 29 anisotropy, 153–155 log model for unconventional shales,
effective–stress tensor and Shmin gradient, rock-physics modeling, 158–161, 191 129–137
40–42 bedding-normal P- and S-wave petrologic data in unconventional shales,
gradient, 176 velocities, 160 121–129
prediction from shale velocity, 34–40 consolidated sandstones, 58–61 rock physics of unconventional shales,
shale-compaction model, 29–32 conventional shales, 66–69 137–161
vertical effective stress, 32–34 Eagle Ford formation, 159 VP-Vs and AI-SI relationships, 161–165
pore-pressure prediction, 36 in sand/shale equences, 56 seismic reservoir characterization, 105
fracture, and overburden gradients pore shapes and porosity thresholds in seismic rock properties, 43, 82
estimation, 37, 39 sands, 56–58 seismic ties, 184
from shale velocity, 34 sandstone diagenesis models, 64–66 statistical wavelet extracted from seismic
good agreement, 38 two case histories, 69–72 data, 185
sonic velocity vs. effective stress in unconsolidated and poorly consolidated synthetic angle, 186
shales, 34–36 sands, 61–64 SEM. see scanning electron microscopy
pore shapes and porosity thresholds in rock-physics templates in siliciclastic (SEM)
sands, 56 sequences, 69–72 SH-wave velocity surface, 25–26
thin-section photomicrographs, 57–58 rock properties of sands, 25 shale anisotropy, 106
poroelastic modulus, 43 rock texture, 126–129 shale–compaction model, 29, 35, 67
poroelastic theory, 43 compacted organic shales with elevated density model for stress computation,
porosity reduction, 29 clay contents, 128 31–32
centroid effect, 31 Eagle Ford limestone, 128 porosity reduction, 29–31
NCT model, 31 Eagle Ford marl, 127 Shale Model, 67
shale porosity of vertical effective high-resolution FE-SEM image of Eagle line, 107
stress, 30 Ford marl, 128 shale-volume fraction, 8
probability density function (PDF), 89 typical Bazhenov shale core, 127 shapes of pores, 58
root-mean-square (rms), 59 Shmin gradient, 173
rule of one-quarter of wavelength, 102 alternative methods of fracture-gradient
Q RVE. see representative- computation, 41
volume-element- (RVE) effective-stress tensor and, 40
quantitative interpretation (QI), 105. Hooke’s law, 40
see also quantitative reservoir S Mohr’s circles for different stress states
description in sandstone reservoirs, 41
quantitative reservoir description, 19 sand and shale diagenesis models, 89 normal-faulting stress regime, 42
quartz, 34, 54, 153 Sand Diagenesis Model line, 107 Shmin stress-gradient estimation
cementation, 56, 61 sand volume computation, 114 anisotropy prediction in organic and
arbitrary line through Vsand volume, 115 conventional shales, 174–175
average well values of N/G from log stress profiling from log data, 175–176
R model, 117 uniaxial-strain-based approach,
thickness–average N/G map, 116 173–174
reflection curvature, 183 sandstone Diagenesis Models, 64–66, 70, Shmin stress-gradient mapping, 194–198
relative dip angle, 23–26, 67, 106 72, 74, 76, 119, 162 for Eagle Ford Formation, 197–198
representative-volume-element (RVE), 43 scanning electron microscopy (SEM), 121, 1D stress-gradient profiles, 195
reservoir-scale anisotropy 126, 129 velocity-versus-effective-stress lines, 196
estimation, 183 seismic analysis Shuey approximation, 88
VTI anisotropy, 184 AI-SI inversion of prestack seismic siliciclastic rocks
rms. see root-mean-square (rms) data, 191 log editing, 20–27
rock composition, 1, 87, 121–122, reservoir–scale anisotropy estimation, permeability prediction in siliciclastics,
161, 175 183–184 13–17

10-Vernik_Index.indd 212 12-08-2016 19:58:19


 Index 213

petrophysical classification of textural information, 126 Vernik-Kachanov model, 160


siliciclastics, 1–8 texture of siliciclastics, 2 vertical effective stress, 29, 32
petrophysical model building, 8–13 thermal maturity, organic richness and, density profiles, 34
seismic petrophysics, 17–20 122–126 schematic geologic, 33
siliciclastics Thomsen’s anisotropy parameters, 97 TVD, 33
Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

AI-SI relationship in, 76–78 3D oil-wet kerogen network, 177 VSP, 33


application to core/log database, 4–8 3D seismic, 105 vertical seismic profiling (VSP), 27, 33,
petrographic data and petrophysical three-term Shuey equation, 88 67, 107
classification, 1–4 TI. see transverse isotropy (TI) vertical transverse isotropy (VTI), 97, 167,
petrophysical classification of, 1 time frames, 80 174, 180
VP-Vs relationship in, 72–76 total-organic-carbon (TOC), 3, 121, 124, anisotropy effect, 97–99
soft inclusions, 53 161, 162 anisotropy parameters, 184
soft porosity, 53 AI-SI template, 188, 192 of clay-rich conventional shale
soft sand models, 60, 62 cumulative–TOC map of Eagle Ford sequences, 99
solid black line, 41 interval, 194 symmetry, 161
sonic logs, 20, 33, 105, 157 curve, 189 verticalization, 106
anisotropic correction, 23–27 Eagle Ford well data, 189 viscoelastic equilibration, 44
sonic velocities estimation from AI-SI template, 186 VP-versus-porosity model, 35
anisotropy correction of, 106–107 mapping, 191 VP-Vs relationships, 161–165
bedding-normal compressional-wave mean–TOC map of Eagle Ford AI-SI crossplot of log data, 165
velocity models, 36 Formation, 193 bedding-normal core
effective stress in shales, vs., 34 reflection coefficients, 190 measurements, 163
shale VP-versus-porosity model, 35 synthetic angle, 190, 195 bedding-normal VP-versus-Vs
sorting, 2–3 TOC vs. bulk density, 187 template, 164
stiff sand models, 60 total porosity, 11–13 geotechnical applications, 75
stress, 32 and kerogen porosity, 131–133 Greenberg-Castagna equations, 74
bedding-normal P-wave velocity vs. relationship between bulk density ρb and in siliciclastics, 72
stress curves, 154 TOC content, 132 shear-velocity-prediction
dependence of velocity and anisotropy, total-porosity-versus-TOC-content accuracy, 76
153–155 relationship, 133 vertical P- and S-wave velocity
maturity level IVb, 155 transverse isotropy (TI), 24, 106, 137 logs, 162
profiling from log data, 175–176 formations, 24 VP-Vs transform coefficients for wet
stress-anisotropy effects, 53 TI shales, elastic-property relationships sands and shales, 73
stress state in, 167–168 water-saturated VP-Vs plot for soft
effective-stress tensor and Shmin gradient, Traugottpower-law model, 30 arenites, 73
40–42 true vertical depth (TVD), 33 VSP. see vertical seismic profiling (VSP)
pore pressure and, 29 tuning-free half-space model, 186 VTI. see vertical transverse
pore-pressure prediction from shale TVD. see true vertical depth (TVD) isotropy (VTI)
velocity, 34–40 two-term Aki-Richards approximation, 88
shale-compaction model, 29–32 two-term Fatti equation, 88
vertical effective stress, 32–34 W
subhorizontal microcrack, 196
symmetry axis, 167 U water-based mud system (WBM), 13
synthetic AVO models, 107 water saturation, 133–134, 135
synthetic forward modeling, 87–88, 185 unconsolidated and poorly consolidated Widess’s rule, 101
synthetic-gather models, AI-SI template sands, 61–64 Winland model, 14, 15
to, 91 unconventional shales
AVO class IV, 93–95 log model for, 129–137
AVO class III and zero-gradient case, petrologic data in, 121–129 X
91–93 rock physics of, 137–161
AVO classes I and II, 95–97 uniaxial-strain-based approach, 173–174 X-ray diffraction (XRD), 4, 121
synthetic modeling, tuning effects in AVO, unloading mechanisms, 29
101 upscaling, 184
AI-SI template, 102 Y
P-wave reflectivity, 103
wedge modeling, 102–103 V Yield stress, 168
Young’s moduli, 167–171
Van Baaren’s equation, 14
T velocity
core measurements, 137–139 Z
tensile strength, 179 effect of kerogen, 144–153
anisotropy in organic shales, 180 intrinsic anisotropy, 141–144 zeolites, 5, 17, 59
direction, 180 stress dependence, 153–155 Zoeppritz equations, 87, 89, 107, 112

10-Vernik_Index.indd 213 12-08-2016 19:58:19


Downloaded 12/09/16 to 155.247.166.234. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

This page has been intentionally left blank

You might also like