You are on page 1of 286

M IN N

LIBRARY

7 #r5T/S
3 5£0

University of Nevada

Reno

The Geochemistry of Lime-Induced Heave in

Sulfate Bearing Clay Soils

A dissertation submitted in partial fulfillment of

the requirements of the degree of Doctor of Philosophy,

in Geology

by

Robert Dalton Hunter

August 1989
©1989
Robert Dalton Hunter
All rights reserved
The dissertation of Robert Dalton Hunter is approved

University of Nevada

Reno

August 1989
ii

ACKNOWLEDGEMENTS

Where do I start? Bob and Robbie Hunter taught me to do

my best and finish what I started. Anyone who reads this thing

will notice that it took guite awhile. My parents gave me the

gift of tenacity, among other things. Dr. Roger Jacobson of

the Desert Research Institute provided outstanding technical

and moral support throughout this lengthy ordeal. His efforts

and patience are greatly appreciated. SEA, Consulting

Engineers, Inc., has been very supportive of this project and

is gratefully acknowledged. Adrienne Frederick of Adrienne's

Office, a long-time friend, typed this beast, including the

dreaded "big nasty awful table" and numerous chemical

equations. Thanks Adrienne— you probably saved my marriage.

Finally I want to thank my family: Candace, my wife, for

her support and patience; and my children, William and Susan,

who, to this date, have never known their father when he wasn't

in school.
iii

ABSTRACT

Clay soil is commonly stabilized in-place by mixing with

anhydrous (CaO) or hydrated (Ca(OH)a) lime. The resulting

chemical disequilibrium causes various cementing and exchange

reactions between the lime and clay. On Stewart Avenue, Las

Vegas, Nevada, unusual reactions between lime, clay, and

sulfate minerals caused severe swelling in lime-treated

subbase. Reconstruction costs approached $3,000,000. Lime-

induced heave has also occurred in California, Utah, Kansas,

Texas, Tennessee, and Paris, France. On Stewart Avenue, heave

resulted from growth of thaumasite, a complex multi-hydrated

calcium silicate-carbonate-sulfate mineral. Thaumasite forms

a solid solution series with ettringite, a multi-hydrated

calcium aluminum sulfate mineral. At temperatures above 15 C,

ettringite probably forms first and later converts to

thaumasite. At least 50% of the sulfate migrated to the

nucleation centers, mostly in capillary water derived from

ponding in utility trenches. Mass flux estimations indicate

that the needed ions could probably have reached the nucleation

sites within the observed time frame of two years.

Solubility determinations were made on ettringite

synthesized during this research. Study results, which

considered ion pairs, are reasonably represented by the

equation log Ket = 196.429 - 1.6699T + 0.002846T2, and

corresponds to —48.49 at 25 "C. The logarithmic expression


iv

leads to AG° = -66.14 ± 0.27 kcal/mol, AH” = -11.0 ± 0.3

kcal/mol, ACP° = 2.4 ± 0.1 kcal/mol, and ASr° = -196 ± 52 cal/deg

mol for the dissociation of ettringite at 25°C. Using this

data, ettringite was found to be the stable phase at high pH

and sulfate concentrations above 15 mg per kg of soil. On

Stewart Avenue approximately 25% of the heave was due to added

mineral mass with the remaining 75% resulting from density loss

and increased soil voids. Complete reaction of 10,000 mg

sulfate/kg soil should induce a one-inch (2.5 cm) heave from £ , ~ # /

a 12-inch (30.5 cm) thick lime-treated soil. Because of ion

mobility, lime treatment of soils containing measurable sulfate

is risky. Swelling can probably be held to tolerable levels

by careful control of drainage and capillarity. Costs of

preventing chemical heave in soils containing sulfate minerals

may exceed the potential savings from lime or cement

stabilization.
V

TABLE OF CONTENTS

ACKNOWLEDGEMENTS ii

ABSTRACT iii

LIST OF FIGURES viii

LIST OF TABLES xi

LIST OF APPENDICES xiii

1 INTRODUCTION 1

2 THEORY AND PRACTICE IN LIME STABILIZATION OF CLAY


SUBGRADES 10
Historical Background 12
The Chemistry of Normal Lime-Soil Reactions 14
Cation Exchange 16
Flocculation/Agglomeration 20
Carbonatation 23
Pozzolanic Reactions 24
Design of Lime Stabilized Subbase for
Pavements 30
Basic Construction Methods 32

3 LIME-INDUCED HEAVE OF SULFATE BEARING CLAY SOILS:


CASE HISTORIES OF STEWART AVENUE AND OWENS AVENUE
LAS VEGAS, NEVADA 34
Site Conditions 34
Stewart Avenue 34
Owens Avenue 37
Geology and Soil Conditions 38
General Geology 38
Project Soil Conditions 40
Exploration 44
Field Observations 57
Laboratory Analysis 63
Soil Testing 63
Index Testing 63
Moisture-Density Curves 63
Volume Change and Expansion 64
Chemical Testing 66
Soluble Salts 66
Cation Exchange Capacity 69
Mineral Identification 69
Analysis and Conclusions 71

4 OTHER CASES OF LIME-INDUCED HEAVE AND RELATED


LABORATORY RESEARCH 85
Other Cases of Lime-Induced Heave 85
Previous Laboratory Research on
Lime-Clay-Sulfate Reactions 91
vi

TABLE OF CONTENTS (Continued)

5 THE MINERALOGY OF THAUMASITE AND ETTRINGITE 93


Thaumasite 94
Ettringite 96
Synthetic Ettringite and Thaumasite 98
Ettringite 98
Thaumasite 99

6 METHODOLOGY FOR EXPANSIVE POTENTIAL AND MINERAL


SYNTHESIS EXPERIMENTS 103
Objectives 103
Sulfate Leaching 104
Expansive Potential 106
Expansive Pressure 106
Volume Change Experiments 107
Experiment One 110
Experiment Two 110
Experiment Three 111
Experiment Four 113
Experiment Five 114
Experiment Six 116
Summary of Experiments on Expansive
Potential 118
Ettringite/Thaumasite Growth in Other Soils 122
Ettringite/Thaumasite Concentration
by the Heavy Liquid Method 125
Ettringite/Thaumasite Concentration
by Elutriation 126
Synthesis of Ettringite and Thaumasite 128
Synthesis of Ettringite 128
Experiment One 129
Experiment Two 130
Experiment Three 131
Synthesis of Thaumasite 134
Experiment One 134
Experiment Two 138
Experiment Three 140
Experiment Four 141
Experiment Five 142
Experiment Six 143
Summary 145

7 THE THERMODYNAMICS OF ETTRINGITE DISSOCIATION 146


Experimental Methodology 146
Evaluation of the Analytical Data from the
Solubility Experiments 154
Summary 159
The Solubility of Gypsum 160
The Solubility of Ettringite 164
The Thermodynamics of Ettringite Solubility 167
vii

TABLE OF CONTENTS (Continued)

8 ANALYSIS AND EVALUATION OF EXPERIMENTAL


AND FIELD DATA 170
Laboratory Simulation of Lime-Induced Heave 170
Expansive Pressure 170
Volume Change 170
Thaumasite/Ettringite Growth in Other Soils 176
Synthesis of Thaumasite 178
Synthesis of Ettringite 178
Mass-Volume Relationships 179
Mass-Flux Calculations 189
Geochemical Reaction Model 191

9 CONCLUSIONS AND PRACTICAL APPLICATIONS 198


Stability Diagrams 199
Portlandite (Ca(OH)2) 199
Gypsum (CaS04-2H20) 201
Ettringite and the Saturation Index 202
Evaluating Susceptibility to
Lime-Induced Heave 207
Engineering Considerations 211

REFERENCES 215

APPENDICES 220
viii

LIST OF FIGURES

1 Photograph of heaved pavement on Stewart


Avenue, Las Vegas, Nevada 3

2 Photograph of heaved median on Stewart


Avenue, Las Vegas, Nevada 4

3 Diagram of the structure of kaolinite 17

4 Diagram of the structure of montmorillonite 19

5 Diagrammatic representation of the fixed and


diffuse layers 22

6 Diagram showing effects of pH on silica


solubility at 25°C 27

7 Diagram showing effects of pH on dissolved


aluminum species in equilibrium with
gibbsite at 25°C 28

8 Location map for Stewart and Owens Avenues,


Las Vegas, Nevada 35

9 Photograph of typical distress along Owens


Avenue, Las Vegas, Nevada 39

10 Generalized soils map in the area of


Stewart and Owens Avenues,
Las Vegas, Nevada 41-43

11 Location of test pits along Stewart


Avenue 45-49

12 Location of test pits along Owens Avenue 50

13 Typical test pit profiles for Stewart


Avenue 52-55

14 Typical test pit profile for Owens Avenue 56

15 Photograph showing concentration of large


crystals along contact between native soil
and subbase from test pit S-7 60

16 Photograph showing specimen of undamaged


subbase from test pit S-26 61

17 Photograph of test pit S-10 showing lens


shape of damaged subbase 62
ix

LIST OF FIGURES (Continued)

18 Trilinear diagram from untreated soils from


Stewart Avenue 73

19 Trilinear diagram from lime-treated soils of


Stewart Avenue 74

20 Typical stiff diagrams from treated and


untreated soils on Stewart Avenue 76

21 Graph of soluble calcium versus soluble


sulfate 78

22 Trilinear diagram for lime treated and


untreated soils along Owens Avenue 79

23 Graph showing percentage of clay sized


particles along Stewart Avenue 83

24 Graph of expansive pressure from growth of


ettringite/thaumasite 108

25 Results of volume change experiment five 117

26 Results of volume change experiment six 119

27 Graph showing estimated percentage of


calcite in synthetic ettringite 133

28 Graph showing the change in the equilibrium


constant of gypsum as a function of
temperature 162

29 Graph showing the change in the equilibrium


constant of ettringite as a function of
temperature 166

30 Diagram showing relative changes in soil


volume from growth of ettringite/
thaumasite 188

31 Diagrammatic representation of the


mechanism of lime-induced heave 195-197

32 Stability diagram for portlandite and


ettringite 200

33 Stability diagram for gypsum and ettringite 203

34 Graph showing untreated and lime treated


soils from Stewart Avenue relative to
ettringite stability field 205
X

LIST OF FIGURES (Continued)

35 Graph showing changes in geochemical


equilibrium induced by treatment with
4.5% CaO 208

36 Graph of calculated maximum volume change


as a function of initial soil sulfate 210
XI

LIST OF TABLES

1 List of Minerals Identified in Stewart


Avenue Samples 70

2 Mineralogic Properties of Thaumasite 95

3 Mineralogic Properties of Ettringite 97

4 Initial Chemistry (in Grams) for


Volume Change Experiment Five 115

5 Summary of Experiments on Expansive


Potential 120

6 Sample Designations and Loctions of Various


Soils for Ettringite/Thaumasite Growth 122

7 Mechanical Analysis of Various Soils


for Ettringite/Thaumasite Growth 123

8 Results of X-Ray Analysis of Other Soils


for Ettringite/Thaumasite Growth 127

9 Summary of Ettringite Synthesis Experiments 134

10 Thaumasite Synthesis
Initial Chemistry of Experiment One Group 135

11 Estimated Percentage of Thaumasite


in Experiment One Samples 137

12 Thaumasite Synthesis
Initial Chemistry of Experiment Two Group 138

13 Estimated Percentages of Thaumasite in


Remaining Samples of Experiment One and
Experiment Two 139

14 Thaumasite Synthesis-Initial Chemistry of


Experiment Three 140

15 Results of Thaumasite Synthesis


Experiment Three 141

16 Results of Thaumasite Synthesis


Experiment Five 143

17 Initial Chemistry of Thaumasite Synthesis


Experiment Six 144

18 Composition of Solutions Used in


Determining Ettringite Solubility 147
Xll

LIST OF TABLES (Continued)

19 Conductivity (in Micromhos) and pH of


Ettringite Solubility Experiment at 10 "c 149

20 Conductivity (in Micromhos) and pH of


Ettringite Solubility Experiment at 25 °C 150

21 Conductivity (in Micromhos) and pH of


Ettringite Solubility Experiment at 4 0 °C 151

22 Chemical Analysis of Preliminary Sample


for Ettringite Solubility 152

23 Analytical Results of Ettringite


Solubility Samples 154

24 EPM Balance of Analytical Data 156

25 Stoiciometric and Measured Mole Ratios 157

26 Calculated Equilibrium Constants for


Ettringite 165

27 Summary of Thermodynamic Values for


Ettringite at 2 5 ° C (298.16°K) 169

28 Data for Mass-Volume Calculations 181

29 Data for Estimation of Mass-Flux on


Stewart Avenue 189
xiii

LIST OF APPENDICES

1 Summary of mechanical test results for


Stewart and Owens Avenues 220

2 Unified soils classification 232

3 Typical volume change and expansive


pressure curves for Stewart Avenue 233

4 Trail dilution ratios for extract of


soluble ions, Stewart Avenue 234

5 Summary of soluble ion chemistry from


extraction of Stewart Avenue soils 235

6 Summary of test results of cation exchange


capacity for Stewart Avenue soils 243

7 Typical x-ray diffraction patterns,


Stewart and Owens Avenues 244

8 Comparison of thaumasite and ettringite


x-ray patterns 249

9 Typical scanning electron microscope


patterns, Stewart and OwensAvenues 250

10 X-ray diffraction pattern from U.S.


Highway 41, Davidson and Sumner
Counties, Tennessee 252

11 Description of experiments on lime-induced


heave in progress, University of
California, Berkeley 253

12 List of x-ray diffraction peaks for


thaumasite 256

13 List of x-ray diffraction peaks for


ettringite 257

14 X-ray diffraction pattern for experiment


VC-5C 258

15 X-ray diffraction pattern for experiment


VC-6A 259

16 X-ray diffraction pattern for sample CS 260

17 X-ray diffraction pattern for sample ES-2 261

18 X-ray diffraction pattern for sample ES-3 262


XIV

LIST OF APPENDICES (Continued)

19 X-ray diffraction pattern for sample


TS1-3A 264

20 X-ray diffraction pattern for sample


TS3-C 265

21 X-ray diffraction pattern for sample


TS6-ED 266

22 WATEQDR ion activity calculations for


gypsum, sample G-25 267

23 WATEQDR ion activity calculations for


ettringite, sample ET-25C 269
1

CHAPTER 1

INTRODUCTION

Lime treatment of clay-bearing soils has been in common

use as an effective means of stabilizing low-strength roadway

subgrade for over 40 years. The procedure involves thoroughly

mixing fine-grained native soils with 3-10 weight percent

hydrous (Ca(OH)2) or anhydrous (CaO) lime. stabilization

proceeds through a combination of four mechanisms, three of

which are chemical reactions directly involving clay minerals

in the soil. A fourth reaction mechanism combines calcium from

lime with carbon dioxide in the water and atmosphere to form

calcium carbonate.

When the ideal lime-soil-water mixture is properly mixed

and compacted, increase in subgrade strength is dramatic. In

addition, the shrink-swell potential of any high plastic,

expansive clays in the soil is virtually eliminated. The rate

of strength increase is rapid at first and then declines. The

net result is a continued increase in strength that approaches

some maximum value asymptotically for the life of the pavement

structure. This maximum is a function of lime content, clay

mineralogy, soil chemistry, moisture, and temperature.

Lime treatment has not always produced the dramatic

stabilization anticipated or even necessary for adequate


2

performance of the pavement system. Under certain conditions

failure occurs through heave induced by reaction between lime

and soluble sulfate in the treated soil (SEA INC., 1983;

Mitchell, 1986; Hunter, 1988). In 1975, a 2.75 mile section

of Stewart Avenue in Las Vegas, Nevada, was reconstructed and

improved to a 100-foot right-of-way. Three years later Owens

Avenue, parallel to and one mile north of Stewart, was also

reconstructed. Both roadways incorporated lime-treated subbase

into their pavement structural sections using 4.5 weight

percent quicklime (CaO). Within one year of completion, both

Stewart Avenue and Owens Avenue began showing major signs of

distress. Over the next few years, the damage steadily

increased, particularly along Stewart Avenue. This distress

included severe cracking and bulging of asphaltic concrete

pavement (Figure 1) as well as cracking and uplift of concrete

medians (Figure 2). Sidewalks were generally undamaged but did

not overlie lime-treated soil. Along Stewart Avenue bulges in

the pavement often reached a height of 12 inches above adjacent

areas of pavement. At Owens Avenue the maximum magnitude of

the heave is approximately 8 inches.

The original 1975 cost to widen and improve Stewart Avenue

was approximately 1.4 million dollars. Two thirds of the

pavement section was reconstructed in 1985 at a cost of 2.7

million dollars. The worst areas of Owens Street have been

repaired; however, it has not yet been reconstructed. Recent


3

Figure 1 View west along Stewart Ave. from test pit


S-7 Heave in this area exceeded 12 Inches. Note
6 inch wide crack in pavement.
Figure 2 View north along Stewart Awe., just east of
Mojave Blvd. Concrete median has heaved over 6 inches.
5

reevaluation (SEA INC., 1988) indicates that the Owens Avenue

distress has become more extensive since 1983.

Forensic investigations by three consultants, including

the soils engineer responsible for the design, were

inconclusive and conflicting. The reported cause of heave

ranged from expansive clay to growth of gypsum induced from

chemical reactions between lime and sulfate in the native soil.

Without a thorough understanding of the distress mechanism, it

was impossible to evaluate the potential for further damage and

make appropriate geotechnical or legal recommendations. In

1982 SEA Engineering, Inc., contracted with the City of Las

Vegas to perform a geotechnical investigation of Stewart

Avenue. The objective was to establish the cause of the

distress and make recommendations for reconstruction or salvage

of the roadway. During that study it was found that the heave

resulted from lime-induced growth of thaumasite, a rare

hydrated calcium-silicate-carbonate-sulfate mineral.

Thaumasite forms a complex solid solution series with

ettringite, a hydrated calcium-aluminum sulfate mineral. Under

the geochemical conditions of Stewart Avenue ettringite

probably grew first and later converted to thaumasite.

The SEA investigation (conducted by this writer in 1982-

1983), and the many questions it necessarily left unanswered,

form the basis for this dissertation. During the SEA study an

extensive literature review found no published cases of lime


6

induced heave and few laboratory studies that considered the

effects of sulfate on lime-treated soil. Telephone

conversations with highway departments in the Western United

States, however, indicated that the problem was not unique.

Since this dissertation was initiated, and an interim

publication completed (Hunter, 1988), other major occurrences

of lime-induced heave have been documented. Problem areas have

included Southern California; Wichita, Kansas; Interstate

Highway 70, Grand County, Utah; U.S. Highway 10, Texas;

Davidson County, Tennessee, and even a warehouse foundation and

floor slab in Paris, France. Experiments performed as a part

of this research have demonstrated the potential for this

problem in many other areas.

In 1985, James K. Mitchell presented a lecture to the

American Society of Civil Engineers as the invited speaker for

the Twentieth Annual Terzaghi Lecture. This prestigious

lecture is in honor of Karl Terzaghi (deceased) who is

generally recognized as the father of modern soil mechanics.

In this lecture, entitled "Practical Problems from Surprising

Soil Behavior," Professor Mitchell discussed four newly

recognized and surprising problems related to geotechnical

engineering. One of the surprises involved the lime-related

distress on Stewart Avenue in Las Vegas, Nevada. The lecture,

which was later published (Mitchell, 1986), was the first

mention of the phenomenon in U.S. literature. The paper

correctly diagnosed the cause of the distress as being growth


7

of ettringite and thaumasite, but recognized that much

additional research would be needed to learn where, why, and

how the problem can occur.

The overall objective of this dissertation was to develop

a detailed understanding of the geochemical mechanism of

ettringite/thaumasite growth in lime-treated soils. Clearly

there is scientific need for this information in that, as

Professor Mitchell pointed out, it is one of the newly

recognized mysteries in soil mechanics. From a practical

standpoint, the need is even more apparent. Lime treatment of

Stewart Avenue saved, perhaps, $300,000 in initial costs over

other techniques commonly used to stabilize clay roadway

subgrade. Subsequent evaluation, repairs, and reconstruction

necessitated by the swelling reactions, however, cost the City

of Las Vegas nearly ten times that amount. It was expected

that the understandings gained from this research could be

applied to other soils in other areas to prevent similar costly

mistakes. Although lime treatment of clays containing any

sulfate could simply be avoided, this would preclude many soils

that respond well to the technique. As a consequence, the

potential cost benefits of lime stabilization would be

sacrificed out of fear rather than on the basis of sound

technical judgement.

The dissertation research can be divided into three main

segments, beginning with a thorough review of published, and


8

other available, literature on lime stabilization. The review

considered the geochemistry of lime/soil reactions, design of

lime stabilized pavement sections and construction methods.

Detailed case histories of Stewart and Owens Avenue were also

provided. Other cases of lime-induced heave were uncovered and

described, demonstrating the widespread nature of the problem.

The physio/chemical mechanism by which ettringite/

thaumasite causes swelling in lime treated soils is detailed

in the second segment. The evaluation involved experiments to

duplicate the swelling in the laboratory and measure the

maximum volume change and pressure generated by the mineral

growth. Mass volume relationships were analyzed confirming

that the in situ sulfate was, at best, capable of generating

only one-half of the heave observed in the field. Mass flux

estimations were developed to verify that sufficient sulfate

and other ions could reach the system within the observed time

frame. On the basis of the overall analysis, a series of

stepped equilibrium reactions were proposed to describe the

geochemical changes in the soil system that result in

precipitation of ettringite/thaumasite.

The final segment of the research involved determination

of the solubility thermodynamics of ettringite. Under most

conditions ettringite probably occurs first, and in the

presence of carbonate and at temperature below 15 C, slowly

converts to thaumasite. Solubility determinations required


9

synthesis of ettringite. Although this has been performed by

previous researchers, synthesis of ettringite pure enough for

thermodynamic work is difficult and required considerable

experimentation. The solubility study was used to calculate

the dissociation (equilibrium) constant, standard state Gibbs

free energy, enthalpy, entropy, and heat capacity for

ettringite. These data, which were not available prior to this

research, were used to generate phase diagrams depicting

geochemical conditions under which ettringite becomes the

stable phase. Such diagrams are the basis for making an

educated evaluation of the potential for a given soil to be

damaged by lime-induced heave.


10

CHAPTER 2

THEORY AND PRACTICE IN

LIME STABILIZATION OF CLAY SUBGRADES

Lime stabilization of roadway subgrades has become common

engineering practice, particularly during the last 25 years.

The relatively simple procedure can be performed with readily

available construction equipment and greatly improves the load-

bearing capacity of certain soils. As a consequence,

thicknesses of the more expensive pavement components,

asphaltic or Portland cement concrete and crushed aggregate

base, can be significantly reduced. Although lime treatment

is expensive, reductions in pavement and base, in many cases,

more than offset the cost of stabilization. The net result is

an economical alternative to increasing the thickness of the

structural section to compensate for weak or expansive subgrade

soils. Lime treatment can also be an effective solution where

existing utility lines preclude increasing the thickness

(depth) of the structural section. This was partially the case

for reconstruction of Stewart Avenue in Las Vegas.

Lime treatment modifies the physical characteristics of

clay-rich soils in a variety of ways and to varied degrees.

These changes are summarized below (National Lime Association,

1976) :
11

1. Decrease plastic index as much as fourfold. This

occurs through an increase in liquid limit and

decrease in plastic limit.

2. The soil is agglomerated, i.e., the percentage of

particles passing the No. 40 mesh sieve are

substantially decreased.

3. Accelerates the breakdown of clay clods increasing

the soils workability.

4. Aids in drying out wet soils— particularly when

anhydrous (quicklime) is used.

5. Greatly reduces the shrink-swell characteristics of

expansive clays thereby minimizing the potential for

heave or settlement.

6. Greatly increases the unconfined compressive strength

of the soil. In some cases lime-treated soils have

forty (40) times the unconfined compressive strength

of untreated soils.

7. Load bearing capacity, as measured, by various tests,

increases substantially. For example, R-values

(ASTM D 2844) often increase by a factor of ten (10) .


12

8. The stabilized material develops significant tensile

strength which is a critical factor in pavement

performance.

9. Forms an impermeable horizon that resists penetration

of gravity water from above and capillary water from

below.

10. Provides a stable working base and prevents pumping

of soft, underlying subgrade.

HISTORICAL BACKGROUND

Lime treatment was used early in man's history to improve

a soils workability and potential for cultivation (Eades and

Grim, 1960). Lime stabilization of clay soils dates back, at

least, several thousand years to roadways constructed by the

Roman Empire (McDowell, 1959). Since the end of World War II,

lime stabilization has been used throughout this country and

the world with increasing regularity.

As commonly happens, the technique was put into practice

long before its physio-chemical mechanisms were understood, or

even of concern. Some of the earliest, modern, lime

stabilization projects involved reconstruction of paved parking

lots and roadways on military bases in Arkansas, Louisiana, New

Mexico, Oklahoma, and Texas during the late 1940's. A follow­


13

up investigation conducted 25 years later demonstrated the long

range durability of the procedure (Kelley, 1977).

Not all of the early attempts at lime stabilization were

successful, however (Taylor and Arman, 1960). It was likely

the failures rather than the successes that brought lime

stabilization to the attention of the engineering and academic

communities. The years from approximately 1959 through 1976

saw a tremendous surge of research and development in the field

of soil stabilization. State highway departments throughout

most of the United States as well as the Federal Highway

Administration and Army Corps of Engineers became involved in

laboratory experiments, test sections, and major lime

stabilization projects. Often this work was conducted in

collaboration with local universities. As a consequence of

these efforts, we now have a reasonably good understanding of

normal lime-soil reactions.

In the past decade, there has been relatively little

research in lime stabilization. There have also been very few

pavement failures that were not the direct result of poor

workmanship during construction. The most remarkable

exceptions are Stewart Avenue and Owens Avenue in Las Veg a s ,

Nevada. Both of these pavements were badly damaged by adverse

chemical reactions between lime and salts in the native soils

(SEA INC., 1983, Mitchell, 1986; Hunter, 1988).


14

THE CHEMISTRY OF NORMAL LIME/SOIL REACTIONS

Stabilization using pure lime is generally restricted to

soils with a moderately high content of clay minerals. In

addition to increasing the bearing strength, lime treatment

decreases the plastic index of the soils, greatly reducing the

shrink-swell characteristics of high plastic clays. Fine

grained soils with low clay content are better stabilized with

lime-fly ash mixtures or Portland cement.

Lime is a rather broad term for a variety of calcium

bearing compounds used in soil stabilization, cement, and

numerous other products. Pure lime is available in either

hydrated (Ca(OH)2) or anhydrous (CaO) form. Anhydrous lime is

commonly called quicklime. In the United Stats, hydrated lime

tends to be favored, largely because it is a safer compound to

work with. Quicklime is generally used in Europe and offers

several advantages. Chemically, quicklime has approximately

25% more available reactant (CaO) than hydrated lime so that

3% quicklime is as effective as 4% hydrated lime. The

significant advantage lies in lower transportation costs— an

important consideration on-large scale projects. In addition,

the reaction of quicklime with water:

CaO + H20 = Ca(OH)2


15

is spontaneous and strongly exothermic. This is useful in

drying wet soils and for working at near-freezing temperatures.

Heat from this reaction also accelerates the various other

reactions involved in stabilizing soil.

The major disadvantages of quicklime are its higher cost

per ton and its highly caustic nature. Application of

pulverized quicklime during hot, humid, windy conditions is a

significant hazard and should be avoided.

Reagent grade lime is not used in construction for obvious

reasons of cost. Commercial grade, high-calcium lime typically

contains a minimum available lime (CaO) of 90%. High dolomitic

quicklime containing at least, 95% total oxides (MgO + CaO) is

also available. A high calcium quicklime with 94% available

CaO was used on Stewart Avenue and Owens Avenue.

Numerous additives have been tried with lime in attempts

to improve its properties of stabilization. The most commonly

used additive is fly ash, however, virtually every reasonably

common salt, oxide, or silicate has been researched (California

Division of Highways, 1967).

Stabilization of lime treated soils occurs through a

combination of four mechanisms (Thompson, 1966):


16

1. Cation exchange

2. Flocculation/Agglomeration

3. Carbonatation

4. Pozzolanic reactions

The first two increase soil workability and are the result of

changes in the electrical charges of the clay minerals. The

second two mechanisms involve cementation reactions which

generate the stabilization and conseguent increase in bearing

strength.

Cation Exchange

The common clay minerals are comprised of alternating

sheets of ions of two basic types. The octahedral or gibbsite

sheet consists of the anions 02' and OH' in octahedral

coordination around Al3+. The second layer, known as the

tetrahedral or silica sheet, is made up of 02' and OH' in

tetrahedral coordination around Si‘


,+.

Two layer clays, such as kaolinite, consist of a

repetition of a combined octahedral and tetrahedral sheet unit

cell (Figure 3). Such clays are held together by Van der Waals

forces acting between the unit cells. The attraction is

relatively strong so that the silica-gibbsite unit cells do not

readily separate from one another. As a consequence, two layer

clays do not absorb large amounts of water between the layers


18

and are not expansive. In addition, cations are not easily

absorbed between the layers so their cation exchange capacity

is very low (Grim, 1968 Berner, 1971).

Unit cells in three layer clays, such as the smectite

(montmorilIonite) group, consist of octahedral sheets

sandwiched between two tetrahedral sheets, as shown in Figure

4. Again, the structural units are bonded by Van der Waals

forces. In this case, however, the attraction between two

tetrahedral layers is less than that between the octahedral

sheet and silica sheet observed in kaolinite. In addition,

with montmorillonite, there is some substitution of Al3+ for Si4+

in the tetrahedral sheet and considerable substitution of Mg2+

for Al3+ in the octahedral sheet. This results in a large net

negative charge which is balanced, to some extent, by

absorption of cations between each unit cell. Since these

cations are loosely held, they are readily replaced by other

cations with more favorable bonding characteristics (Lambe and

Whitman, 1969). Typically, the order of exchange is

Na+ < K+ < Ca2+ < Mg2+ where each cation will tend to replace any

to its left (Thompson, 1966). Because water is a polar

molecule, it is also easily absorbed between the unit cells of

montmorillonite. This gives rise to the shrink-swell

characteristics of certain clay minerals, particularly

montmorillonite and vermiculite.


19
20

The cation exchange capacity of two layer clays is

relatively low while that of three layer clays, particularly

the smectites, is quite high (Grim, 1968; Berner, 1971). As

a consequence, the benefits of cation exchange in lime treated

soils are far more dramatic in expansive clays. In these

clays, divalent calcium ions typically replace monovalent

sodium or potassium ions in the interlayer positions. This

results in much stronger bonding between the unit cells and a

remarkable drop in the soils shrink-swell characteristics. The

cation exchange process is generally complete within 24 hours,

however, significant improvement is immediately apparent. Two

layer clays are generally lime-treated only to increase bearing

strength since they are not severally expansive.

In "real world" soils, the situation is somewhat more

complicated since pure, well-ordered clay minerals are not

common. More often, the clays are of "mixed layer" structure

consisting of random or orderly alternations of various two-

or three-layer unit cell types. The behavior of lime treated

soils is, therefore, difficult to predict from simple tests of

cation exchange capacity. Test specimens mixed, molded, and

cured are normally required.

Flocculation/Aaalomeration

All soil particles carry a net electrical charge. In

theory this charge can be either positive or negative, however,


21

only negative charges have been measured. By virtue of their

small size (less than 0.002 mm), clay minerals have a high

specific surface and hence, a high net charge. Although other

factors are involved, this negative charge is largely the

result of isomorphous substitution within the crystal lattice.

Typically this consists of Al3+ replacing Si4+ or Mg2+ replacing

Al3+.. The clay particle then attracts loosely bonded,

exchangeable cations to neutralize the net charge (Lambe and

Whitman, 1969). The result is an electrical double layer

consisting of a negatively charged surface surrounded by a

diffuse cloud of cations. The structure of this double layer

has been modelled mathematically and is thought to consist of

a fixed layer of cations essentially attached to the solid

surface and a diffuse layer in which the ions are free to move.

The inner, fixed layer is known as the Stern layer while the

diffuse cations lie within the Gouy layer (Figure 5).

When two clay particles approach they are repelled from

one another by the like charges of their Gouy layers. When the

ionic strength of the pore water is increased, as happens with

lime treatment, the Gouy layer is compressed, closer to the

solid surface. The electrostatic repulsion keeping the

particles apart is then no longer sufficient to counteract the

Van der Waals forces which tend to bring them together (Drever,

1982). As a result, the clay minerals flocculate forming

agglomerations of particles and a net increase in the effective

grain size of the soil.


22

Layer of
fixed
cations Diffuse layer

Stern Gouy

Figure 5 Diagrammatic representation of the fixed (stern)


and diffuse (Gouy) layers. (Source: Drever, 1982, p. 80.)
23

In reality, the mechanisms of flocculation/agglomeration

are extremely complex and, perhaps, the least understood

component of lime stabilization.

Carbonation

Reactions between calcium (or magnesium) and carbon

dioxide are the lesser of the two groups of cementation type

reactions involved in lime stabilization. The reaction

proceeds through one of the most common and best understood

mechanisms in geochemistry:

C02 + H20 = H2C03

H2C03 = 2H+ + C032' (at pH > 10.5)

Ca2+ + C032' + 2H+ + 2 (OH)* = CaC03 + 2H20

In this case both calcium and hydroxyl ions are supplied by the

dissolution of lime:

Ca(0H)2 = Ca2+ + 2 (O H )"

Carbon dioxide is available directly from the atmosphere, from

C02 dissolved in water added to the soil-lime mixture and from

biogenic C02 present in the soil itself.


24

In reality, calcium carbonate cementation is simply an

unavoidable by-product of the lime-soil-water system and not

altogether desirable. Calcium tied up in calcite is

unavailable for the pozzolanic reactions that produce a

considerably stronger cementing agent (Mateos, 1965; Thompson,

1966; Zube, et. a l . , 1966). Laboratory tests with lime-treated

bentonite cured in a vacuum have shown a 68% increase in 28-

day unconfined compressive strength compared to specimens cured

in a C02 saturated atmosphere (Mateos, 1965). Indeed, care

must be taken in the storage of lime to prevent carbonation

and decrease of reactivity.

Pozzolanic Reactions

By definition (ASTM C-593-762) a pozzolan is "a siliceous

or alumino-siliceous material that in itself possesses little

or no cementitious value but that in finely divided form and

in the presence of moisture will chemically react with alkali

and alkaline earth hydroxides at ordinary temperatures to form

or assist in forming compounds possessing cementitious

properties." Pozzolanic recations are the major sources of the

cementation and are the dominant factor in stabilization and

increased strength in lime treated soils (Thompson, 1966).

Research has demonstrated that significant pozzolanic

reactions may not occur until enough lime is added to satisfy

the potential for cation exchange and flocculation/


25

agglomeration reactions (Taylor and Arman, 1960; Hilt and

Davidson, 1960). This natural affinity of the soil for calcium

ions is termed lime fixation. Lime fixation is a function of

clay mineralogy and grain size (Hilt and Davidson, 1960).

Once the soil is sufficiently saturated with lime, the pH

of the soil-water system stabilizes at 12.3. This value

exceeds any known in nature and is sufficiently high that

silica becomes extremely soluble (Krauskopf, 1979). Silica

solubility is strongly affected by both temperature and pH

(Barnes, 1980). At surface temperatures, however, pH is the

dominant factor. More recent research has demonstrated that

silica solubility also increases in the presence of various

salts (Fournier, 1983). In all cases, silica is thought to

dissolve through the following series of stepped reactions:

Si02 + 2H20 = H4Si04 ; log K, = -3.8

H4Si04 = H+ + H3Si04" ; log K2 = -9.9

H3Si04 = H+ + H2Si042- ; log K3 = -11.7

At a pH of 12.3, H2Si04" is the dominate species. Most natural

waters have a pH between 6 and 9. From pH 2 to 8, the

solubility of quartz (Si02) and silica gel (Si02) remain nearly

constant at, approximately, 6 ppm and 120 ppm respectively.

Solubility of both forms of silica increases exponentially

above a pH equal to 8.5. At pH 12.3, silica solubility is

quite high (Figure 6).


26

The solubility of aluminum is also largely controlled by

pH, however, the behavior is somewhat more complex than that

of silica. Nevertheless, at high pH levels, the solubility of

aluminum also becomes relatively high (Figure 7).

Lime stabilization research has shown that clay minerals

become unstable and begin to deteriorate at pH levels above

10.5 (Eades and Grim, 1960; Davidson, Demirel and Handy, 1966).

The breakdown of these clays provides the siliceous and

aluminous pozzolans for reaction with calcium (Thompson, 1965).

Lime stabilization of clay-poor soils has been effectively

accomplished by adding fly ash or volcanic ash (pozzolanic

materials) to the lime-soil mixture (California Division of

Highways, 1967). Portland and other hydraulic cements

typically consist of pozzolanic material and a few minor

additives. As a consequence, the basic chemistry of the

reactions involved in lime and hydraulic cement are very

similar.

When calcium reacts with silicon and/or aluminum at high

pH, numerous compounds form (Taylor, 1964). In general they

are difficult to study since they are present in small amounts

relative to the mass of the soil. In addition, most of the

compounds are initially amorphous and become more crystalline

through aging or increased temperature. The combination of

small mass and lack of an ordered internal crystal structure


2 4 6 8 10 12 14
PH
Figure 6 Activities of dissolved silica species in equilibrium with quartz at 25 C.
(Source: Drever, 1982, p. 92)
with gibbsite at 25 C. (Source: Drever, 1982, p. 95.)
makes the compounds difficult to identify since they are not

easily recognized by x-ray diffraction.

The most commonly cited pozzolanic reaction product is a

complex hydrated calcium silicate (Taylor, 1964; Ruff and

Clara, 1966). Its formation in lime-treated soils can be

represented by the following series of stepped reactions:

5CaO + 5H20 = 5Ca(OH)2

5Ca (OH) 2 = 5Ca2+ + 10 (0H‘)

Clay minerals = 6H3Si04‘ + 6H* (at pH 12.3)

/ /
5Ca2+ + 6H3Si04" + 10 (OH') + 6H+ = Ca5(Si018H2) • 4H20 + 12H20

The compound is chemically similar to the mineral tobermorite,

however, tobermorite is fully crystalline. The semi­

crystalline "tobermorite-like" compound is generally considered

to be the main cementing agent in both lime and cement treated

soils and in concrete (Taylor, 1964; Ruff and Clara, 1966).

The compound forms as a gel and actually goes through several

transitions with increasing crystallinity (Ruff and Clara,

1966) ;

CSH(gel) = CSH (II) = CSH(I)

where CSH abbreviates calcium-silicate-hydrate. This slow

transition from gel to semi-crystalline accounts for the steady


30

strength increase of concrete and cement or lime treated soils

that continues for many years.

Numerous other hydrous calcium-silicate, calcium-

aluminate, and calcium-aluminum-silicate compounds have been

identified in lime or cement treated soils and concrete

(Taylor, 1964; Ruff and Clara, 1966).

DF.SIGN OF LIME STABILIZED SUBBASE FOR PAVEMENTS

Over the years, many laboratory test methods have been

employed to evaluate the strength and load bearing capacity of

lime treated soils. These have ranged from simple methods

involving only the plastic index of the soil (McDowell, 1966,

1972) to elaborate schemes of triaxial shear testing (Taylor

and Arman, 1960). The most commonly used test procedure,

however, has been unconfined shear strength of test cylinders

molded under various conditions. All of these tests, however,

have serious limitations in that they are not routinely used

to evaluate subgrade strength of pavement design. Most modern

design methods use Resistance (R) value testing (ASTM D-2844)

as the measure of subgrade strength for asphaltic pavements and

K values (modulus of subgrade reaction) for Portland cement

pavements. Since K-value testing (ASTM D-1196) is a rather

arduous field chore, K—values are usually approximated from

R-values which are easily measured in the laboratory.


31

In the R-value test, 2 1/2-inch high, 4-inch diameter soil

samples are compacted in a metal mold using a kneading

compactor. The R-value is determined from a combination of the

pressure at which water exudates from the sample under a

specific uniaxial load and a modified measure of its triaxial

shear strength. The test is normally run at 3 or more moisture

contents and the R-value arbitrarily assigned to be that

occurring at an exudation pressure of 300 psi. R-value testing

can also be conducted with different percentages of lime mixed

with the soil to determine the weight percent lime needed for

adequate stabilization. Since the test is generally completed

within a 48-hour period, increases in R-value do not reflect

all of the longer term pozzolonic reactions so important to the

stabilization process. Even so, weak clay soils with initial

R-values on the order of 5 usually show increases to the 50 to

60 range (California Department of Transportation, 1976). Such

values would be typical for silty sands and gravelly sands— a

significant strength increase.

Unfortunately, none of the test methods commonly used for

evaluating lime-treated subbase would indicate the potential

for lime induced heave. Such tests normally last only a few

days. In most cases, weeks or months might be required before

any distress becomes apparent.


32

BASIC CONSTRUCTION METHODS

The basic steps of lime stabilization involve

scarification, partial pulverization, lime spreading, mixing,

curing and compaction. This can be accomplished satisfactorily

by a variety of methods, however, specialized equipment is

preferred for major projects.

For small-scale projects the lime is often mixed in place.

The subgrade is thoroughly scarified to the appropriate depth

with rippers on a grader or bulldozer. The lime can then be

spread out from bags in the required amount and mixed with the

soil by further scarification. During mixing, water is added

to raise the moisture of the soil-lime mixture to at least five

percent (5%) over optimum for hydrated lime (Ca(OH2) and even

higher for quicklime (CaO). The treated soil is then cured for

24 to 48 hours prior to compaction. This allows the cation

exchange and flocculation/agglomeration reactions to occur,

greatly improving the soil's response to compactive effort.

The curing period is sometimes extended to as much as seven

days for highly expansive clays.

In larger projects, the initial scarification is generally

performed with a disc harrow and the lime brought in bulk

trucks, rather than bags. Since thorough mixing of lime and

soil is crucial to the process, it is generally preferable to

further pulverize and blend the materials with a rotary mixer.


33

For projects requiring the maximum possible stabilization, a

two-step application of the lime is commonly used. In the

first application, approximately one-half of the required

percentage of lime is thoroughly mixed with water and the soil.

Following a curing period of approximately one week, the

remaining lime is added. The first treatment preconditions the

soil making it much easier to mix and compact with the

remaining lime. The second application supplies the lime for

the major stabilizing pozzolanic reactions (National Lime

Association, 1976).

Numerous variations in the methodology have been tried.

In all cases, the most critical factors are thoroughness of

mixing, proper moisture conditioning, curing time and

compaction. Lime treated soils are generally densified to

either ninety or ninety-five percent (90-95%) compaction

relative to ASTM D-1557-78).


34

CHAPTER 3

LIME-INDUCED HEAVE OF SULFATE BEARING CLAY SOILS:

CASE HISTORIES OF STEWART AVENUE AND OWENS AVENUE

LAS VEGAS, NEVADA

SITE CONDITIONS

Stewart Avenue

Stewart Avenue is a major east/west trending roadway

through downtown Las Vegas. The section covered by this

investigation extends approximately 2.75 miles (4.44 km)

eastward from 28th Street to Nellis Boulevard (Figure 8). The

street includes four twelve-foot (3.66 m) traffic lanes, two

parking lanes, and a concrete or asphalt median. Left turning

lanes have been provided at all intersections.

As designed and constructed in 1976, Stewart Avenue used

two separate structural sections. Between 28th Street and

Pecos Road four inches (10.2 cm) of asphaltic concrete, five

inches (12.7 cm) of Type II aggregate base, and a twelve-inch

(30.5 cm) thickness of lime-treated native soil (subbase) were

required. From Pecos eastward to Nellis Boulevard, the


NELLIS I BLVD.
35
36

aggregate base thickness was increased to eight inches

(20.3 c m ).

The Stewart Avenue distress consisted of severely damaged

asphaltic pavement, concrete and asphalt medians, as well as

occasional curbs and gutters. Damage to sidewalks, which were

not intentionally lime treated, was rare. The most badly

damaged asphalt pavement occurred between Mojave Boulevard and

Pecos Road. In this area a major ridge paralleled the roadway

in the south parking lane (refer to Figure 1). The ridge rose

as much as twelve inches (30.5 cm) above the adjacent pavement

and ranged from one to two feet (30.5 to 61 cm) in width.

Often the adjacent pavement had also been heaved so that the

maximum uplift was hard to gauge. The ridges generally

displayed large fractures in the asphalt— typically one to

three inches (2.5 to 7.5 cm) in width.

The remaining asphaltic pavement and medians between

Mojave and Pecos showed extensive damage with the pavement

exhibiting a generally uplifted and undulating surface.

Concrete median slabs were displaced three to six inches (7.5

to 15.2 cm) (refer to Figure 2). From 28th Street to Mojave

and from Pecos to Sandhill the distress was very similar in

form but of a somewhat lower magnitude. Damage was

predominately in the median and on the south of the street.


37

Between Sandhill and Lamb Boulevard the heaving was

generally confined to isolated areas on the south side of

Stewart or in the median. East of Lamb damage became more

isolated but occurred on both sides of the street as well as

in the median.

In 1986 the area of Stewart Avenue between 28th Street

and Lamb Boulevard was reconstructed, without lime stabil­

ization. Damaged areas east of Lamb Boulevard were repaired.

Owens Avenue

Owens Avenue was reconstructed between Eastern Avenue and

Pecos Road in 1978 (refer to Figure 8). This segment

incorporated an eight-inch (20.3 cm) lime treated base which

was directly overlain by ten inches (25.4 cm) of asphaltic

concrete.

Distress along Owens Avenue is similar in form to that

of Stewart Avenue but of a considerably lesser magnitude.

Maximum heave along linear ridges is six to eight inches (15.2

to 20.3 cm). The ridges are also less continuous than at

Stewart Avenue and occur about equally on both sides of the

street. The majority of the roadway is undamaged or only


38

slightly damaged, often limited to cracking of the pavement.

No displacement of sidewalks, curbs, or gutters was observed,

however, damage to the median is common. A photograph typical

of the Owens Avenue distress is presented as Figure 9. As of

late 1988 the roadway had not been reconstructed. Some repair

work, however, has been performed in the most seriously

damaged areas. A follow-up study (SE&A, Inc., 1988) showed

that the aerial extent of the damage has increased since 1983.

GEOLOGY AND SOIL CONDITIONS

General Geology

The Stewart Avenue and Owens Avenue alignments lie on the

eastern side of the Las Vegas Valley in an area of thick basin

fill sediments. These materials typically consist of fine

grained sandy silts and sandy clays originally deposited in

marshy or playa environments during the Plio-Pleistocene

(Bell, 1981, Mifflin and Wheat, 1979). The deposits were

derived from a variety of source rocks which, for the most

part, are dominated by carbonates. Some of the source areas

contain abundant bedded gypsum and possibly other evaporites.

A considerable sequence of undifferentiated, Tertiary volcanic

rocks lie along the southeast rim of the basin (Longwell, et

al, 1962).
40

Project Soil Conditions

A generalized soils map of the area has been included as

Figure 10. Soils underlying the two roadways are of Recent

Age, derived from reworking of the older basin fills by

drainages along the Las Vegas Wash (Dinger, 1977). They are

poorly to slightly stratified and, on a gross scale, fairly

homogeneous. On a much finer scale, they are complexly

interbedded and contain areas with significant differences in

engineering and chemical properties. Dinger (1977) used

groundwater chemistry to divide the Las Vegas Valley into

carbonate and sulfate geochemical lithofacies. Stewart Avenue

and Owens Avenue both lie in the sulfate lithofacies, slightly

east of the boundary with the carbonate lithofacies.

The soils in the project area contain abundant soluble

salts indicative of playa environments. In most soil samples,

salts were visible to the unaided eye. The upper soils unit

typically consists of moist, slightly stiff to stiff, brown

to tan, low plastic sandy clay or sandy silt with seventy to

ninety percent (70-90%) fines. Thickness of this horizon

varies from three-and-one-half feet (106.6 cm) to over eight-

and-one-half feet (259.1 cm).


N ellis Blvd
(Source: USDA, 1967.)
DESCRIPTIONS OF SOIL UNITS

r,M^ ara^isc siIt loani> drained, slightly saline (Pa)— proved by means of drainage ditches and by pumping
Uns soil occurs as a small area in the central part of Las ground water. RunolT is very slow. Permeability is me­
Vegas valley. It adjoins and extends into the city of dium as far down as the hardpan and slow through the
Las Vegas. pan. I ho inherent fertility and the water-holding capac­
Representative profile: ity arc high. The erosion hazard is slight.
Surface soil— I(or tho most part, this soil is under natural vegeta­
0 to 2 inches, clark-gniy silt loam; weak, medium to fine,
granular structure; slightly hard when dry, friable when tion. If irrigated and properly managed, it would he
moist; p I [ S.4.
2 to 7 inches, gray loam; massive; hard when dry, friable
well suited to crops. It would need fertilizer containing
when moist; p ll S.4. nitrogen and phosphorus, even though its surface soil is
Subsoil— high in organic matter. Also, it would need deep leaching
7 to 14 indies, lifiht-gray silt loam; massive; slightly hard
when dry, friable when moist; p lf 8.2. and drainage to remove the excess salts. The slow perme­
14 to 31 inches, light-gray silt loam; massive; slightly hard ability of the hardpan and the pressure of artesian water
when dry, friable when moist; pH 8.2. are likely to interfere with tho installation of an artificial
Substratum—
31 to 52 inches + , white, weakly lime cemented silt loam;
drainage system. Irrigation, unless carefully controlled,
m any extremely firm lime nodules; massive; very hard is likely to result in a perched water tabic above the
when dry, firm when moist; p it 8.4. hardpan, waterlogging of the soil above tho hardpan,
The thickness of the surface soil ranges from 2 to 12 and accumulation of excess salts in the surface layer.
inches, depending on the amount of leveling. The depth The surface soil and subsoil have severe to very severe
(o the lime-cemented horizon ranges from 21 to more limitations as foundation-bearing material; the substra­
than GO inches. This horizon is several feet thick in tum has moderate limitations. The shrink-swell potential
spots and is harder and more cemented with increasing is moderate to high, depending on the sodium sulfate
depth. & concent ration. (Capability unit IIw-G)
Tho natural drainage was poor, but it has been im­

Figure 10b Generalized soils map for the S tew art/O w ens Ave. area.
DESCRIPTIONS OF SOIL UNITS

Glendale silt loam, slightly saline (Gf).—This soil Land silty clay loam, wet, strongly saline (|_h).—This
occurs as large'tracts in ah area that extends along Las soil occurs mainly in Las Vegas Wash. Excess water
Vegas Wash from a point east of Las Vegas northeastward from the sewage disposal plant has seeped into the soil
toward Nellis Air Force Base. and raised the water table, and the soil is saturated
[Representative profile: at a depth of 2 to 3 feet. All of this soil is in natural
Surface soil— vegetation. It is unsuitable for cultivation and is best
0 to 5 inches, pink silt loam ; moderate, medium to coarse,
platy structure; slightly hard when dry, friable when used as native pasture. No attempt should be made
moist; pH 9.0. to cultivate it or to alter the native vegetation. The
Subsoil and substratum — forage is adequate for grazing but could be improved
5 to GO inches + , very similar to the surface soil, but stratified
w ith either slightly finer textured or slightly coarser
by water spreading. (Capability unit VIw-36)
textured m aterial; massive; pH S.4.
The color ranges from pink to very pale brown. The
texture of the strata in the subsoil and substratum ranges
from fine sandy loam to light silty clay loam, and the
strata vary in thickness and in amount of deposition. Land very fine sandy loam, wet, strongly saline (Ln).-—
There is some mica in the profile and, in most places, This soil occurs at the lower extremities of Las Vegas Wash,
some gypsum. where a geologic barrier..forces ground water upward
Natural drainage is good, runoff is slow', permeability into the soil profile. It is saturated below a-depth of
is moderately slow, and infiltration is slow. Both the 12 inches during most of the year. In winter the -water
water-holding capacity and the natural fertility are high. table is likely to rise to the surface, and the soil becomes
This soil is poor or unsuitable for most engineering swampy. It is unlikely that this soil could be leached
purposes. The shrink-swell potential is low to high, de­ of excess salts unless it is drained. The native vegetation
pending upon the concentration of sodium sulfate salts. should be maintained. Preferably, it should be used as
The sulfate hazard to concrete is severe. (Capability unit food and cover for wildlife, but it would provide some
I-A) ' forage for livestock. (Capability unit VlIw-6)

Figure 10c Generalized soils map for the S tew art/O w ens Ave. area.

U)
44

The upper horizon grades downward into a white to gray,

gravelly sandy clay. Because the test pits were generally

shallow this horizon was not frequently observed. The clay

shows a definite increase in calcium carbonate, however, a

hard caliche layer was not encountered during exploration.

A lens containing up to fifty percent (50%) gypsum and

approximately two feet (61.0 cm) in thickness was encountered

on Stewart Avenue near Mojave Road.

EXPLORATION

Thirty-four test pits and three borings were placed along

Stewart Avenue. Six test pits were excavated along Owens

Avenue. The location of each Stewart Avenue test pit and

boring is shown on Figure 11, while those of Owens Avenue are

presented on Figure 12. Test pits typically ranged from three

to four-and-one-half feet (91.4 to 137.2 cm) deep and reached

a maximum depth of eight-and-one-half feet (259.1 cm) on

Stewart Avenue just east of Mojave Road. The three Stewart

Avenue soils borings were advanced to depths ranging from five

to fifteen feet (152.4 to 457.2 cm).

Test pits were excavated through the asphalt in severely

damaged, moderately damaged, and undamaged areas of travel

lanes, turning lanes, shoulders medians, gutters, and curbs.


otoo 6t 00

T
NTS

I T E S T PIT L O C A T IO N
Figure 11a L o c a tio n of te s t pits along S te w a rt Ave.
00
38 TOO

NTS

I T E S T PIT L O C A T IO N
Figure 11b L o ca tio n of te s t pits along S te w a rt Ave.
IIO +OO M2 tOO IK too

id
S-30
iEli

: t 00

IT E S T PIT L O C A T IO N
Figure 11 d L o c a tio n of te s t pits along S te w a rt Ave.
51

Control test pits were placed along Stewart Avenue in unpaved

areas, adjacent to badly damaged sections of road as well as

adjacent to undamaged sidewalks.

In paved areas asphaltic concrete was sawcut and

carefully removed with a backhoe. A sand cone density test

(ASTM D1556) was then taken in an undisturbed portion of

aggregate base. At Owens Avenue, where aggregate base was not

used, this step was eliminated. The base section was then

gently removed and a sand cone density test taken in the

underlying lime treated subbase. The lime treated subbase was

next removed to expose the native soils. In undamaged areas,

the subbase was extremely well-cemented and could only be

removed with a jack-hammer. The native soils were density

tested and the test pit extended to its full depth. A density

test was occasionally taken at the bottom of the test pit,

particularly where a change in the soil was noted. The

results of the density testing are presented in Appendix 2.

Typical descriptions of the soils encountered as well as

test pit profiles are presented as Figure 13 for Stewart

Avenue. Figure 14 presents typical soils descriptions and a

test pit profile for Owens Avenue.


DESCRIPTION

<8>
0 -0 .4 6 ' A s p h a lt ic Co n c ra ta

0 .4 6 -1 .0 ' A p ora oa ta Basa w ith 1B1


n o n p la s t ic fin a s ^ 491 sa nd, 331 g re v a l

1 .0 - 2 .O' Lima T ra a ta d Subbasa c o n -


sI s t i n g o f w a t, s o r t , l ig h t g r a y , Sandy
SI I t w ith 541 s l i g h t l y p la s t i c f i n a l ,
2AJ sa n d s , 221 camantad g ra v a ls t o 3/4"
d lo m a ta r. M a ta rle l d is p la y s a m o rrlad
appaaranca and Is com prlsad o f 50-601
w h lta m ln a ra ls up t o */2" d la m a ta r. Soma
c r y s t a l s t r u c t u r a v l s l b t a unoar m a g n ifi­
c a tio n .

3 .0 - 4 .5 ' M o is t , s t i f f , l ig h t brown,
Sandy Claw w it h 671 maclim p la s t i c
T T n e s T - T 7 T w a ry fln a se nd. M inor s a lts
v l s l b l a uncar m a g n if ic a tio n , a s p a c le lly
from 3 - 4 ' .

2 ■I VERTICAL EXAGGERATION A SAMPLE LOCATION

Figure 13b Test pit p ro file S t e w a r t Ave. S - 8 .

Ol
U>
DESCRIPTION

0 -0 .5 ' A s p h a lt ic C o n c ra ta
1761
0 . 5 -1 .0 8 ' Ao gra qa ta Beta w it h 4J
noop la s t I c t fn a s , j t in a -c o a r s a sand.
3 2 J g ra va f t o !■ d l a r a t a r .

1 .0 8 -1 .8 3 * Lima Tra<_______________
'a d Subbasa c o n -
s i s t l o g o f m o is f , v a r y cans* (s x c a v a tlo n
ra p u lra d Jaekham m ar), brow n, C w a o ta d
1760 S o *1 la c k in g th a » o t t I ad apparanca of
a Is rra s s a d a ra a s . W hlta c r y s t a l s up to
Y 2 " d la a a ta r com prlsa lass than 2 S of
s o i l m s s . M a t a rla l shows a w h lta s a l t y
s u rfa c a whan d r y . D i s t i n c t l y s o f t a r In
, °**r 0 . 2 ' , su g g a s tln g soata chaailcal
ra e c t lo n has o c c u r ra d .

1 759 1 .8 5 -3 .8 3 ' M o is t , s o f t , brow n, Sandy


S n * T C la y w ith 86 f low p la s t i c Y l n a s ,
14J v a r y f l n a s a n d . V a ry p o r o u s , no
s a l t s v l s l b l a un d a r m a g n if ic a t io n .

1758

1757

8 10 12
DISTANCE (F E E T )
2 ' I VERTICAL EXAGGERATION
A SAMPLE LOCATION

Figure 13c Test pit p ro file S t e w a r t Ave. S - 2 2 .


U1
U1
D«T *EwAT./E WOShmE CLAV
DESCRIPTION

0 ~ 0 .9 2 ' A s p h a lt i c C o n c re te

^ • 9 2 - 2 . OB' Lim a T r e a te d Subbase c o n s is t in g


100 o f m o is t , s o f t , 1 1 g'h+ g r a y , SI l t v Sandv
G re v a l w it h 1 9X n o n p le s t lc f i n e s , 2 B I
sa n o , 5 2 t g r a v a l t o 2 " d la m a t a r . w h it*
m in a r e ts up t o 1 /2 " d la a > a ta r c o m o rls a 50-401
o f th a t o t a l s o i l mess g i v in g The m e r e r le I
a m o tt le d a p p e a ra n c e . Sum* c r y s t a l s t r u c ­
t u r e v i s i b l e u n d e r m a g n !f Ic e t J o n .

2 . 0 5 -2 .5 ' U t i l i t y b a c k f i l l c o n r ls t in o o r
Sandy C -raval c o v e r s m ost o f th e tr e n c h
boTto m . N a t! v s s o l i in n o r th * .* :
c o n s is t s o f a m o is t , s t i f f , brown, Sandy
C la y a l t h 8 5 J s l l g n r t o low p l a s t i c f in e s ,
151 v e ry f i n e sa nd. M ino r s a l t s v i s i b l e
In n a t iv e under m a g n if ic a tio n .

EAST SIDEWALL

10 12
DISTANCE (F E E T )

Z-\ VERTICAL EXAGGERATION A SAMPLE LOCATION

Figure 14 Test pit p ro file Owens Ave. 0 - 5 .

Ol
Q\
57

Representative samples for mechanical and chemical

testing were collected from the aggregate base, lime treated

subbase, and native soils. Bulk samples for mechanical

testing were collected from trench sidewalls and sealed in

plastic containers. Samples for chemical testing were also

collected from trench sidewalls and were placed in air tight

jars. Soil temperatures were measured to accuracies of plus

or minus 0.5 degrees centigrade at various levels within

selected test pits to give an indication of the temperature

range appropriate for geochemical modeling.

Field Observations

The effects of lime treatment on the native soil varied

drastically between two extremes. In badly damaged areas the

subbase consisted of soft, light gray mineral aggregate with

the mechanical properties of a plastic silt (Figure 13a). In

undamaged areas lime treatment resulted in an extremely hard,

cemented, medium brown material that could not be described

as soil (Figure 13d). All gradations between the two extremes

were observed in the test pits. Often, the extremely hard

material graded to the soft, wet subbase in the length of a

single exploration trench. Vertical gradation was also

observed in some instances.


58

The undamaged subbase was medium brown in color and

generally contained less than five percent (5%) white

crystals. These white minerals were clearly of secondary

origin with a maximum size of approximately three-fourths inch

(19.1 mm). The damaged subbase was a light gray to almost

white color, and showed very little of the brown coloration

typical of the native soils or undamaged subbase. In the most

severe cases the subbase consisted of an estimated 30-60%

light colored secondary minerals grown subsequent to the lime

treatment process. Maximum size for individual minerals was,

again, approximately three-quarters inch (19.1 mm), however,

the majority was much smaller. Large crystals were

occasionally concentrated along the contacts between base and

subbase or subbase and native soil (Figure 15).

The contacts of undamaged subbase with the overlying

aggregate base and the underlying native soils were straight

and parallel to each other. The layer was of uniform

thickness, typically twelve inches (30.5 cm) on Stewart Avenue

and eight inches (20.3 cm) on Owens Avenue. These thicknesses

are in good agreement with the design. The material contained

obvious, darker brown fragments of native soils that had not

been mixed with lime. These fragments reached a maximum size

of approximately two inches (5.1 cm). When air dried in the

sun, the undamaged subbase developed a white salty appearance


59

attesting to the abundance of dissolved salts in the soil pore

water. A photograph of this material is shown in Figure 16.

Areas of major damage were almost always found adjacent

to an obvious source of water. This included permeable

utility trench backfill (Figure 14), yard drainage, areas of

poor surface drainage, and construction joints such as occur

at the concrete median and asphaltic pavement intersection.

In most instances the worst damage was found to be adjacent

to or overlying trench backfill. Occasionally, damage vras

observed on both sides of a utility trench, but not directly

over the trench.

The damaged subbase was distinctly lens shaped, thickened

in the most distressed areas and thinning progressively with

lower degrees of damage. This is clearly visible in Figure

17. The maximum measured thickness was 2.25 feet (68.6 cm)

on Stewart Avenue— more than twice that of the undamaged

material. The lower contact was often slightly convex

downward indicating some expansion into soft, native soils.

The upper contact was sharply convex demonstrating that the

expansion was predominantly upward (refer to Figures 13 and

17). Figure 14 from Owens Avenue is a rather unusual case

where upward and downward deflections were about equally

divided.
Figure 16 Photograph of undamaged subbase from test pit S-7.
Note salty appearance.
63

t.
&ROT? ATOR Y ANALYSIS

Soil Testing

Index Testing

Samples of aggregate base, lime treated subbase, and

native soil from each test pit on Stewart Avenue and Owens

Avenue were analyzed in the laboratory to determine their

grain size distribution and plasticity characteristics.

Testing included sieve analysis, hydrometer analysis, liquid

limits, and plastic limits. The results of these tests are

summarized in Appendix 1 for both Stewart Avenue and Owens

Avenue. Index testing allows the soils to be classified in

accordance with the Unified Soil Classification System

(Appendix 2) and correlated with the range of mechanical

properties that could be expected for each soil type.

Moisture-Densitv Curves

Moisture density curves were developed in accordance with

ASTM D-1557-78 for most of the native soil and lime treated

subbase samples. The results of these tests are summarized

in Appendix 1 and were used for many of the other tests and

experiments.
64

Volume Change and Expansion

Volume change tests were conducted on remolded samples

of lime treated subbase and native soil to determine the

potential volume change of the soil in accordance with AASHTO

Designation T-116. The purpose of these tests was to evaluate

the expansion potential of the native clays since previous

consultants had implicated swelling clay in their distress

analysis. In addition, it was hoped that the tests would give

some indication of the potential for continued expansion in

the lime treated subbase.

Samples were tested at four different moisture contents:

in place, five percent (5%) below optimum, optimum, and five

percent (5%) over optimum in accordance with the appropriate

moisture-density curves. Samples of native soil were tested

at ninety percent (90%) compaction relative to ASTM D1557-78.

Samples of lime treated subbase were tested at ninety-five

percent (95%) relative compaction in accordance with the

construction recommendations of the original soils report

(Converse-Davis and Associates in 1974). Lime-treated subbase

and native soil samples were tested with surcharges

appropriate for the overlying structural section.


65

Expansion tests were also conducted on the native clays

and lime treated subbase using an FHA swell meter. This test

provides a measure of the potential expansive pressure of the

soil sample. Tests were run at various moisture contents in

order to evaluate the relationship of moisture content to

expansive pressure. Again, the native soils were run at

ninety percent (90%) and the lime treated subbase at ninety-

five percent (95%) relative compaction. Typical volume change

and expansive pressure curves are presented as Appendix 3 for

Stewart Avenue.

Samples of native soil obtained from subbase grade in

unpaved shoulder areas adjacent to badly damaged pavement

(Sample S-9B) were mixed with 4.5 weight percent quicklime

(CaO) at a moisture 4% over optimum moisture and quicklime

contents of our laboratory mixture were selected in accordance

with those reported by Converse Davis and Associates in their

field inspection summary dated 1975. The laboratory mixture

of lime treated soil was allowed to cure for 24 hours and then

separated in three six-inch (15.2 cm) high, three-inch (7.6

cm) diameter, clear lexan, cylinders. Soil was placed m the

cylinders in five uniform layers densified to approximate the

original field density of ninety-five percent (95%) relative

compaction (ASTM D1577-78). The three cylinders of lime

treated soils were each placed in a 500 ml glass beaker. One


66

beaker was filled with a magnesium sulfate solution containing

20,000 ppm S042' and another a sodium sulfate solution

containing 20,000 ppm S042". The third beaker was filled only

with de-ionized water. A fourth test cylinder containing

untreated native soil at ninety-five percent (95%) relative

compaction was placed in a beaker containing the 20,000 ppm

magnesium sulfate solution. Temperatures were allowed to

fluctuate diurnally between 3° and 18° centigrade. After a

period of only 48 hours all three of the quicklime-treated

specimens had expanded, developing semi-vertical cracks in the

lexan cylinders. Increased damage and cracking could be

observed on a daily basis. After a period of three weeks the

damage to the three cylinders was severe. Cracks had

increased to one—half inch (12.7 mm) in width. The cylinder

containing untreated native soils showed no damage after 120

days. It is important to note that the experiment was set up

so as to require the water to move upward through the soil by

capillarity.

chemical Testing

Soluble Salts

Samples of aggregate base, lime treated subbase, and

native soils from various depths were analyzed for total


67

extractable ions of calcium, sodium, potassium, magnesium,

sulfate, chloride, carbonate, and bicarbonate. Selected

samples were also analyzed for manganese, aluminum, and

silica.

In order to remove most of the soluble ions, initial

samples were extracted at trial dilution ratios of 5:1, 10:1,

20:1, and 50:1. Curves generated from these data (Appendix

4) demonstrated that soluble ions could best be extracted from

native soils and lime treated subbase using dilution ratios

of 50:1. Even at this ratio, not all of the calcium carbonate

dissolved. Dilution ratios of 5:1 were used for aggregate

base samples which contained a much lower concentration of

soluble salts.

For the fine grain soils, 400 gram samples were oven

dried at a temperature of 30-35 degrees centigrade in

accordance with the methods recommended by the Soil

Conservation Service (USDA, 1972). Dried samples were

screened through a No. 40 sieve, thoroughly mixed and split

into four fractions. A 70 gram sample from one of the splits

was added to 500 milliliters of de-ionized water and mixed for

five minutes using a high speed soil dispersion mixer. This

soil-water suspension was transferred to a one-gallon plastic

storage vessel and the stainless steel mixing cup thoroughly


68

rinsed into the vessel with 700 milliliters of de-ionized

water. An additional 2300 milliliters of de-ionized water was

then added to complete the soil dilution and the vessel

sealed. For the coarse grain aggregate base samples 200 grams

of dry soil were added directly to 1000 milliliters of de­

ionized water in the storage vessel. Accuracy of preparation

was double checked by weight and found to be within plus or

minus 0.05%.

Storage vessels were periodically shaken and then allowed

to equilibrate for a minimum of 48 hours. A 300 milliliter

water sample was obtained by filtering the soil-water solution

into a sealable glass vessel. These samples were analyzed by

atomic absorption in accordance with the method described in

Standard Methods for Examination of Water and Wastewater, 15th

Edition, 1980. The results of the extraction analysis are

tabulated in Appendix 5 for Stewart and Owens Avenues.

The extraction solution of each sample was measured for

pH. In addition, all lime subbase samples and selected

samples of native soil were measured for pH. This was

accomplished by mixing the soil with enough de-ionized water

to form a thick slurry and then making the measurement with

a Hach Model 1975 pH meter. Typical soil pH values are shown

on the test pit logs (Figures 13 and 14).


Cation Exchange Capacity

Ten samples of the native soil were tested for cation

exchange capacity using the sodium acetate method described

by the Soil Conservation Service (USDA, 1972). The cation

exchange capacity of a soil is function of the types of clay

minerals it contains and the geochemical environment. The

more expansive clays such as those of the smectite and

vermiculite groups generally have a higher capacity for cation

exchange. The results of this testing are presented in

Appendix 6.

MTNERAT, IDENTIFICATION

Representative samples of damaged lime treated subbase,

undamaged subbase, and native soil were analyzed by x-ray

diffraction to determine their constituent mineralogy.

Typical diffraction patterns and the indicated minerals are

presented in Appendix 7.

Analysis of selected samples was also conducted using a

Scanning Electron Microscope. The SEM uses energy dispersal

analysis to identify the constituent elements of a compound.

This is particularly useful for differentiating minerals such


as thauxnasite and ettringite which, as shown in Appendix 8,

have very similar x-ray patterns. Typical energy dispersal

analyses from the Scanning Electron Microscope are presented

in Appendix 9. The analysis was performed at the University

of Nevada, Mackay School of Mines laboratory.

The minerals identified in the samples are summarized

below in Table 1.

Table 1

I.ist of Minerals Identified in Stewart Avenue Samples:

Mineral Composition

Native Soil

gypsum (primary) CaS04-2H20

calcite (primary) CaC03

quartz (primary) sio 2

lime Treated Subbase

thaumasite (secondary) Ca6[Si(0H)6]2-(C03)2- ( S 0 J 2-24H20

gypsum (secondary) CaS04 •2H20

quartz (primary) Si02

polyhalite (secondary?) K2HgCa2(S04)4*2H20

The x—ray work shows a significant increase of calcite

and gypsum in the subbase, relative to the native soils. This


71

indicates that secondary calcite and gypsum have formed as

reaction products of the lime/soil system.

ANALYSIS a n d c o n c l u s i o n s

Results of the analysis for soluble salts at a 50:1

dilution ratio have been summarized in Appendix 6 both in

units of milligrams per 1000 grams soil and milliequivalents

per liter. This information was plotted on trilinear and

stiff diagrams in order to interpret the original soil

chemistry and the chemical effects of lime treatment.

Figure 18 shows a trilinear diagram of native (untreated)

soil soluble ion chemistry from Stewart Avenue. The soluble

ions are dominated by calcium and sulfate with variable ratios

of sodium + potassium and total carbonate, HC03' + C032 . The

percentages of magnesium and chloride are very uniform

throughout the soils. The most significant factor demon­

strated by the diagram is the broad range of soluble ions

shown by the native soils. There is no distinctive grouping

of soils from undamaged versus damaged areas. This indicates

that the lime/soil reactions were not noticeably affected by

variations in initial soluble ion chemistry.


72

As would be expected lime treatment resulted in major

changes in the soil chemistry. These changes are apparent in

Figure 19, a trilinear diagram of the lime treated soils. The

diagram shows a pronounced relative migration of anions toward

the S042" apex. The cations show a significant increase in

relative percent calcium, however, it is far less than would

be expected from the addition of massive amounts of lime.

Calcium added by lime treatment is two to ten times that

contained in the native soil so that the points should all

plot very near the Ca2+ apex. The fact that they do not

indicates that either the calcium has been leached away or

that it is tied up in insoluble minerals. The grouping of

points on the cation diagram from undamaged samples (24, 25,

29, 33) is critical in this regard. These soils contain

abundant secondary calcite which is relatively insoluble and

is clearly tying up some of the added calcium. The grouping

of these samples in the anion triangle and the composite

diamond also indicate that these areas contain abundant

insoluble calcite.

The relative increase in sulfate indicates that either

additional S042' has migrated into the soil or that total

carbonate and chloride have decreased. In fact, additional

carbonate had to be introduced into the soil to form secondary

calcite. Likewise, secondary gypsum is also present and,


73

rooO

CATIONS PERCENT OF T O T A L
M I L L I E Q U I V A L E N T S PER LITE R

Figure 18 Trilinear diagram for untreated


soils from S tew art Ave.
soils from Stew art Ave.
75

consequently, the shift toward the S042' apex must represent

an influx of sulfate. This is readily apparent from stiff

diagrams, examples of which are presented as Figure 20.

The remaining calcium, sulfate, and carbonate, are

thought to be tied up in thaumasite. The solubility of

thaumasite was inferred from the diagrams and mineralogy, to

be slightly less than that of calcite, and considerably less

than that of gypsum. For example, sample S-10B shows abundant

gypsum in the x-ray pattern. The corresponding stiff diagram

shows a large increase in both soluble sulfate and calcium

relative to the untreated soil (S-10C). Sample S-27B shows

very little gypsum, but abundant thaumasite and calcite in the

x-ray pattern. A relatively minor change in the stiff polygon

is noted after lime treatment. The most significant change

is a slight decrease in the amount of soluble carbonate. The

x-ray pattern for S-30B shows only traces of thaumasite and

gypsum, but a very high content of calcite. A small increase

in the sulfate and calcium axis of the stiff polygon is

observed, largely attributable to dissolution of gypsum.

Again, a slight decrease in soluble carbonate is observed.

The decreased solubility of carbonate in the lime treated

soils is a general but not infallible trend throughout all the

chemical data. The relationship is complicated somewhat by

the increase of carbonate with depth shown in the control test


MILLIEQUIVALENTS/LITER MILLIEQUIVALENTS/LITER

CO2
3

and u n tre a te d S te w a r t Ave. soils.


77

pits. This effect is probably minor over the distance

separating the samples of subbase and native soils. Clearly,

if thaumasite was more soluble than calcite the damaged

subbase would show increased soluble carbonate. As a

consequence thaumasite was inferred to be no more, and likely

somewhat less, soluble than calcite.

The calcium-to-sulfate-to-carbonate ratio in thaumasite

is three to one to one (3:1:1). This accounts for the

distinctive difference in the relative migration of ions

toward the calcium and sulfate apexes of the stiff diagrams

after lime treatment. Three times more calcium than sulfate

is tied up in thaumasite, which is relatively insoluble.

Figure 21, a graph of calcium versus sulfate, shows that both

of these ions, before and after lime treatment, are mostly

derived from dissolution of gypsum. If thaumasite was near

gypsum in solubility the lime treated points would plot well

above the "gypsum line."

Figure 22 is a trilinear diagram for both lime treated

and untreated soils obtained along Owens Avenue. Here the

trends are not as clear as those observed for Stewart Avenue,

however, there is still no obvious association with damaged

areas and native soil chemistry.


79

•LIME TREATED SOIL


°UN TREATED SOIL

Figure 22 Trilinear diagram for untreated and


treated soils from Owens Ave.
80

The damaged subbase was extremely soft and contained an

unusually high moisture content— often ranging from thirty to

fifty weight percent (30-50%). In-place dry densities

measured in damaged subbase typically ranged from 65 to 85

pounds per cubic foot for relative compactions of 50% to 80%.

During construction this material was densified to at least

ninety percent (90%) relative compaction (Converse-Davis and

Associates, 1975). This sharp decrease in density can be

directly correlated to increase in volume and growth of

minerals with a low specific gravity.

The results of this investigation clearly demonstrate

that the distress on Stewart Avenue and Owens Avenue was due

to chemical reactions between quicklime and the native soils.

The new minerals grown include thaumasite, polyhalite, gypsum

and calcite. The seemingly sporadic occurence of distress

along Stewart Avenue was found to be the result of two primary

factors. Neither of these factors relate significantly to

native salt chemistry or the construction process. In all

cases the critical factor was the ability of water to enter

the soil/lime system. Both ettringite and thaumasite are

highly hydrous minerals. Without an abundance of water they

simply cannot form. Virtually all of the distress occurred

near an obvious water source.


81

Typically, heave paralleled a utility trench backfilled

with highly permeable material. Localized (non linear) areas

of damage could generally be traced to specific houses which

channeled roof and yard drainage to the street. Storm drains

in Las Vegas are often intentionally undersized and the street

itself expected to carry the bulk of the water during

occasional thunderstorms. During one small thunderstorm

(November 1982) water was observed running well above

sidewalks on Stewart Avenue. Ponded water exceeded a depth

of four feet (122 cm) in some places. It was concluded,

therefore, that stormwater entered the subsurface along

construction joints and migrated through the permeable

aggregate base and into the utility trench backfill. After

ponding in low lying areas of the trench, capillarity and

diffusion supplied saline solutions to ettringite/thaumasite

nuclei. Damage also was common along the construction joints

themselves, especially between the asphalt and the concrete

curb and the asphalt and the median. As the swelling

progressed construction joints widened and the pavement

cracked. This condition made it increasingly easier for water

to enter the system, doubtless accelerating the heave.

Although the availability of water is the dominant

factor, it does not explain the distinctive decrease in

observed damage on Stewart Avenue east of Lamb Boulevard.


82

Analysis of the hydrometer data, however, demonstrated a

dramatic drop in the percentage of clay sized particles east

of Lamb Boulevard. These data are shown graphically as Figure

23. It was concluded that the decrease of clay sized

particles correlated with a decrease in clay minerals

available for reaction. Where sufficient silica and alumina

could not be generated from dissolution of clay, the calcium

reacted to form calcite, gypsum and, presumably pozzolanic

compounds, rather than large quantities of thaumasite/

ettringite. Although minor thaumasite is present in undamaged

subbase, it is apparently not of sufficient quantity to cause

distress.

Thaumasite is a rare mineral forming a solid solution

series with the more publicized mineral, ettringite (Kollman

and Strubel, 1981):

Thaumasite: Ca6[Si (OH)6]2*(S04)2* (C03)2-24H20

Ettringite: Ca6[Al (OH)6]2* (S04)3*26H20

In theory, this means that any composition between the

two end members is possible and that the pure minerals might

be uncommon. The x-ray diffraction patterns of thaumasite and

ettringite are nearly indistinguishable, as shown in Appendix

8. The two minerals can, however, be distinguished by energy


dispersal analysis on the electron microscope (Appendix 9).

The thaumasite identified during this study very closely

approaches the composition of the pure synthetic end member

and contains very little aluminum.


85

CHAPTER 4

OTHER CASES OF LIME-INDUCED HEAVE AND

RELATED LABORATORY RESEARCH

OTHER CASES OF LIME-INDUCED HEAVE

Deterioration of concrete in high sulfate environments

through ettringite growth has been recognized as a problem for

many years (Taylor, 1964; Lukas, 1975). Considerably less

attention has been given deterioration of lime and cement

treated soils through adverse reaction with sulfate. Although

the Stewart/Owens cases are the most severe examples known,

other cases of lime-induced heave have been documented.

An early reference to sulfate—lime distress (Sherwood,

1962) states: "An instance has been recorded in South

Australia where the presence of calcium sulfate in the soil

resulted in disintegration of a soil cement road." This single

sentence description references an unpublished report which

could not be located.

A lime-treated school parking lot in the east Las Vegas

area has suffered extreme heave subsequent to the Stewart

Avenue distress. A lime treated parking lot in Wichita,

Kansas, is also reported to have failed through growth of

ettringite sometime after 1980 (Mitchell, 1986). In 1975


86

(Gouda, Roy, and Sarkar, 1975) thaumasite, and an intermediate

phase between thaumasite and ettringite, were described in a

deteriorated soil cement. The soil cement was from a seven-

year-old pavement section in Pennsylvania. In this case no

significant heave was mentioned. The pavement simply failed

prematurely due to loss of support in the cement treated base.

Conversations with the State of Texas Highway Department

(Gerald Peck, personal communication, 1983) revealed the

occurrence of lime-induced heave on U.S. Highway 10 between

Amarillo and Fort Worth, in 1975. Here lime treatment of

native soils produced isolated swelling after placement of

asphalt treated base, but prior to paving. The base material

was removed from the damaged areas and the underlying lime-

treated subbase was simply recompacted. There has apparently

been no report of recurring or additional damage. Damage was

blamed on the growth of gypsum and it was stated that it was

an occasional problem where lime reacted with sulfate in the

native soils. Maximum heave was on the order of 1/2 inch (12.7

mm). No further investigation of the phenomena was undertaken;

the entire documentation of the event being limited to a memo

to the file.

The Utah Department of Transportation (Loren Rausher,

1983, personal communication) experienced lime-induced heave

on Interstate 70 in Grand County, eastern Utah. Swelling

occurred in lime treated subgrade prior to paving. Subgrade


87

was rerolled and no further incident has been reported. The

growth of the mineral ettringite was thought to have been the

cause. Maximum heave was less than one inch. No testing was

conducted and no written documentation is available.

In both the Texas and Utah cases swelling occurred prior

to placement of pavement. Repair was limited to recompacting

isolated areas of the subgrade which was not alarming enough

to make the occurrence any more than a curiosity. No

additional cases of lime-induced heave were known by the main

offices of state highway departments in Arizona, New Mexico,

California, nor Nevada.

In 1972 the Orange County Road Department (Santa Ana,

California) observed severe cracking of asphaltic concrete

pavement on a street with lime stabilized clay base. An

expansion study initiated soon thereafter used laboratory

specimens of the native clay soil treated with three percent

quicklime (CaO). The treated soil was compacted at optimum

moisture and allowed to swell against a surcharge load equal

to the weight of the overlying pavement, approximately 46

pounds (101.2 kg). One specimen was soaked in tap water and

the other in high sulfate water. After 20 days both specimens

had expanded approximately one percent. After 600 days the

sulfate specimen had expanded 10.4 percent while the tap water

specimen had a volume increase of 6 percent. Neither specimen

showed any sign of a declining expansion rate. The same clay,


88

untreated and soaked in tap water expanded 8.8 percent in 20

days and 9.6 percent in 300 days (Graf, 1986; personal

communication). Although it will never be known whether the

pavement distress was caused by ettringite/thaumasite, the

laboratory experiments were such that ettringite should have

grown, at least in the sulfate specimen.

A recent paper (Hunter, 1988) presented the early findings

of this dissertation. The response to that paper uncovered

several other major cases of lime-induced heave. The Tennessee

Department of Transportation (TDOT) forwarded an in-house

report (Gaffron, et. al., 1988) describing pavement distress

on U.S. Highway 41 in Davidson and Sumner Counties, Tennessee.

The aggregate base for this roadway had been blended with a

mixture of lime and fly ash which normally results in a product

much like cement treated base. Fly ash provides the highly

reactive (metastable) silica necessary for the pozzolanic

reactions. A description of the distress is quoted from the

TDOT report:

rin-v-i noroinhpr 1987 a number of distortions of

l i n e a r b u lg e s w e re a ls o n o te a
w ith th e ro ad w ay. C o n cre te
s ta b le th ro u g h o u t th e p r o je c t .
89

The investigation by TDOT was fairly extensive and

included field measurements of the distress, test pits, core

samples, density testing, and a variety of physical testing in

the laboratory. No chemical or mineral identification tests

were performed. As a result of their evaluation, the following

conclusion was reported:

The bulging of the pavement began in December. An


increase in moisture during this period changed the
strength of the subgrade and the base since the base
was still in a somewhat plastic state. Hydration
and strength gain of the lime fly-ash base is known
to be sensitive to temperature. The cooler ground
temperatures of late fall may not have been
sufficient to totally cure the base. The base acting
as a plastic material began to displace under traffic
or expansion loading. When the material displaced,
it became less dense and gained volume. The bulges
in the pavement are thought to be the result of
plastic displacement. Some bulges occur in wheel
paths and many occur in the center turn lane. There
is also little sign of rutting in the wheel paths.
This may indicate that the bulges are related to
expansions of the base and are not totally of load
or traffic origins.

As can be seen from the above statements, the evaluation

was somewhat inconclusive. However, the distress described

was strongly suggestive of lime-induced heave. Telephone

conversations with the TDOT (Gaffron, 1988, personal

communication) produced a set of samples for further analysis.

Native subgrade soil and material from a core of the damaged

base were subjected to x-ray diffraction. The results of this

analysis (Appendix 10) revealed the presence of significant

quantities of ettringite/thaumasite in the matrix of the lime

fly ash-base mixture. Since there was no indication of sulfate


90

bearing minerals in the subgrade, it was assumed that the

sulfate source was most likely in the aggregate itself. The

aggregate is rock derived from coal mining in the area and

might be expected to contain some sulfur. As an alternative,

sulfur could be supplied by sulfate reducing bacteria in the

soil. In this case the sulfur would migrate into the base as

a gas, H2S, and oxidize to sulfate in the high pH environment

(Garrels and Christ, 1965). The linear nature of the bulges

is thought to be a surface reflection of the pathway of water

entering the aggregate base.

On the basis of the TDOT report, photographs of the

damaged roadway, and the x-ray diffraction work, it was

concluded that the distress was clearly the result of lime-

induced heave caused by growth of ettringite/thaumasite. Since

no calcite (or other carbonate) occurs in the native soil or

base, the swelling mineral is probably near the end-member

composition of ettringite. In addition to being a well-

documented case of lime-induced heave, the damage to U.S.

Highway 41 represents the first known case where gypsum was not

involved. Previously it was assumed that the phenomenon was

restricted to arid climates where gypsum and other sulfate­

bearing evaporate minerals are common. The Tennessee

occurrence greatly expands the set of geochemical conditions

under which lime-induced heave can occur. Another similar

situation reportedly occurred in Kentucky (Gaffron, 1988,


personal communication), however, the details are not readily

available at this time.

A third and somewhat different case of lime-induced heave

was presented in a publication received from France. In this

paper (Habib and Aversenc, 1988) described an industrial

building in Paris that was damaged by lime-induced heave.

Subgrade fill beneath the concrete floor slab had been mixed

with both lime and cement in an effort to increase the subgrade

modulus (bearing strength) of the material. Gypsum in the fill

reacted with the lime and/or cement producing ettringite. The

resulting volume change produced heave along the edges of the

floor slab in the range of 5 to 20 millimeters. The swelling

occurred slowly over the course of two years and was continuing

at a steady rate at the date of the publication. Deflection

of the floor slab also resulted in uplift of the foundation and

deformation of the exterior walls.

PREVIOUS LABORATORY RESEARCH ON LIME-CLAY-SULFATE REACTIONS

Several researchers have investigated the effects of

sulfate on lime and cement stabilized soils in the laboratory.

Lambe, Michaels and Moh (1960) and Ladd, Moh, and Lambe (1960),

Hollis and Fawcett (1966) concluded that small amounts of

sulfate produced a significant increase in the strength for

certain soil types. Evidence for detrimental effects of high

sulfate concentrations on cement and lime treated soils


92

presented by Mehra, et. al. (1955), Sherwood (1958, 1962),

Lambe, Michaels and Moh (1960) and Cordon (1962). A report

published in 1974 by the California Division of Highways

stated: "Sulfates can be detrimental to lime-treated soils

because they enhance swelling and may induce disintegration

when the mixture is saturated." In a major laboratory study

of lime stabilization, they demonstrated that potential for

swell increased with the sulfate content of the treated soil.

Maximum volume increases were on the order of 4% and attributed

to growth of ettringite.

A major laboratory study of ettringite/thaumasite growth

in various soils is presently in progress at the University of

California, Berkeley (Mitchell, 1988, personal communication).

The writer was asked to review the proposed experiments

(Appendix 11) and provide suggestions that might improve the

outcome or answer additional questions. The experimental

results are not yet available, however, they are expected to

be very helpful.
93

CHAPTER 5

THE MINERALOGY OF THAUMASITE AND ETTRINGITE

Thaumasite and ettringite are both complex hydrated

calcium sulfate minerals. The two minerals are relatively

rare, occurring in scattered locations throughout the world.

Commonly, but not necessarily, thaumasite and ettringite are

found in association with one another. As previously noted,

they form a complete solid solution series which makes their

association considerably less surprising (Kollmann and Strubel,

1981). Several minerals with compositions between the end

members have also been identified (Edge and Taylor, 1971;

Murdoch and Chambers, 1985; Hurlbut and Baum, 1960; Kollmann

and Strubel, 1981). Another mineral, jouravskite

(Ca6Mn2(S04)2(C03)2(O H )2•24H20 ) , is also isostructural with

ettringite and thaumasite (Edge and Taylor, 1971). Although

jouravskite is apparently found only on the dumps of the

Tachgagalt No. 2 vein, Anti Atlas, Morroco, under certain

geochemical conditions, a more complex solid solution series

involving jouravskite, ettringite and thaumasite might be

possible. In jouravskite, manganese replaces aluminum or

silica in the ettringite/thaumasite structure.

Ettringite and thaumasite are most often found in zones

of hydrothermal alteration or low temperature thermal

metamorphism. Both minerals are commonly associated with


94

zeolites and are clearly of supergene origin (Hurlbut and Baum,

1960; Kollmann and Strubel, 1981). No reference to either

mineral occurring naturally in low temperature agueous

environments has been found. There is no indication that

either ettringite or thaumasite occur in playas or in

association with evaporite minerals. Synthetic ettringite, in

minor amounts, is a normal reaction product of Portland cement

and some expanding grouts. Excess ettringite, such as occurs

from sulfate attack, causes rapid mechanical deterioration of

concrete (Taylor, 1964). Synthetic thaumasite probably occurs

in many concretes but has yet to be recognized. Again, this

may be largely due to the similarity of the thaumasite and

ettringite x-ray diffraction patterns. In concrete an

intermediate phase would be a likely end product.

THAUMASITE

Thaumasite comes from a Swedish word meaning to be

surprised— doubtless referring to the mineral's remarkable

composition (Dana, 1919). In this regard, the hydrated

calcium-silicate-carbonate sulfate-hydroxide is unparalled

among minerals (Dana, 1919). Thaumasite was first described

by A. E. Nordenskold in Langban, Sweden in 1878 (Nordenskold,

1878, Kollmann, 1978). Some of the samples Nordenskold

examined had been collected as early as 1802 (Dana, 1919).

Since that time, the mineral has been recognized at numerous

locations including Crestmore, Riverside County, California,


95

Beaver County, Utah; Cochise County, Arizona; Paterson and West

Paterson, New Jersey; and Fairfax County, West Virginia (Dana,

1919; Font-Altaba, 1960).

Thaumasite has been reported with various compositions,

largely due to its tendency toward solid solution with

ettringite. The true composition of the pure thaumasite end

member is Ca6[Si(0H)4]2-(C03)2* (S04)2-24H20 (Welin, 1956; Edge and

Taylor, 1970; Kollmann, 1978; Kollmann and Strubel, 1981). The

following crystalographic and physical properties have been

ascribed to thaumasite (Roberts, et al., 1974; Dana, 1919).

Table 2

MineraJocric Properties of Thaumasite

Crystal System Hexagonal


Class: 6
Space Group: P63
Z: 2
Lattice Constants: a = 21.80 c + 20.58
3 Strongest Diffraction Lines: 9.66 (10) 3.792 (75)
4.582 (65)
Optical Constants: w = 1.500-1.507
e = 1.464=1.468 (-)
Hardness: 3.5
1.91 (measured)
Density:
1.916 (calculated)
(101-1) in traces; brittle
Cleavage:
Crystals acicular to
Habit: filiform; usually massive,
compact
Colorless to white;
Color: transparent to translucent
Vitreous to somewhat silky
Luster:
or greasy

diffraction peaks has been


* The complete list of x-ray
included as Appendix 12.
96

Differential thermal analysis of thaumasite shows a large

endothermic peak at 206 °C corresponding to loss of H 20 and C02.

Between 200° and 550°C, thaumasite changes to a glass-like

amorphous state (Kirov and Pouieff, 1968). A weak endothermic

peak at 709°C occurs as the mineral recrystallizes to a mixture

of anhydrite and larnite. Above 900°C calcium silicosulfate

and anhydrite are the thermal decomposition products (Fonta-

Altaba, 1960).

The thaumasite structure is based on columns of empirical

composition (Ca3Si(0H)6'12H20)4+ running parallel to the C axis.

Groups of C022' and SO<2' occur between the columns. Thaumasite

was the first structure in which octahedral groups of Si(OH)6

were proven to exist (Edge and Taylor, 1971).

ETTRINGITE

The type locality for ettringite is near Ettnngen, Rhine

Province, Germany. The mineral occurs there in cavities of

metamorphosed limestone inclusions in volcanic flows.

Ettringite was first described at that location by J. Lehmann

in 1867. In the United States ettringite has been reported m

Crestmore, Riverside County, California, Cochise County,

Arizona; and Franklin, New Jersey.

Like thaumasite, ettringite of varied composition has been

reported, largely because of the solid solution phenomena


97

(Murdoch and Chalmbers, 1958, 1960). The actual, pure end

member composition is Ca6[Al(OH)6]2* (S04)3*26H20 (Moore and

Taylor, 1968; Kollmann, 1978; Kollmann and Strubel, 1981).

Ettringite has the following crystallographic and physical

properties (Roberts, et al., 1974).

Table 3

Mineraloaic Properties of Ettringite

Crystal System: Hexagonal


Class: 6/m 2/m 2/m
Space Group: P63/mmc
Z: 8
Lattice Constants: a = 22.46 c = 21.44
3 Strongest Diffraction
Lines*: 9.65 (100) 5.58 (75) 3.21 (65)
Optical Constants: w = 1.491 e = 1.470 (-)
Hardness: 2-2.25
Density: 1.77 (measured) 1.79 (calculated)
Cleavage: (101 - 1 )
Habit: Flattened hexagonal dipyramids;
also prismatic, usually without
terminal faces; also as thin
fibers
Color: Colorless and transparent; also
milky white
Luster: Vitreous

* The complete list of x-ray diffraction peaks has been


included as Appendix 13.

Ettringite occurs as columns of empirical composition

(Ca3(A l (O H )6) *12H203+, which run parallel to the C axis. Between

the columns lie the sulfate ions, along with the remaining

water molecules (Kollmann, 1978).


98

gVNTHETir: ETTRINGITE AND THAUMASITE

Ettrinaite

The synthetic preparation of ettringite was discovered in

1874 by J. Lehmann. Lehmann simply mixed an aqueous calcium

sulfate solution with an aqueous solution of calcium aluminate

at room temperature. The immediate white precipitate was found

to be ettringite. Since that time at least 15 other chemical

combinations have been found to yield synthetic ettringite at

temperatures between 5 and 50 C. The simplest technique

(Michaelis, 1892) precipitates ettringite through aqueous

reaction of calcium hydroxide and aluminum sulfate:

6Ca (OH) 2 + Al2(S04)3*18H20 + 6H20 =


Ca6(A1 (O H )6)2•(SO,)3•2 6H20

(Kollman and Trost, 1977; Kollman, 1978).

Synthetic ettringite is produced in minor amounts as a

normal reaction product in Portland cement concrete. The

addition of gypsum in the manufacture of Portland cement must

be controlled such that ettringite formation is completed

before the time the initial concrete set. Excessive gypsum (or

other soluble sulfate compounds) leads to internal expansive

forces larger than those that can be absorbed by pore spaces

in the concrete. The result is a decrease in both flexural and

compressive strength whenever the optimal sulfate content is

exceeded (Taylor, 1964).


99

Thaumasite

Thaumasite is considerably more difficult to synthesize

than ettringite. This appears to be primarily the result of

two nearly mutually exclusive reactions that must occur in

order for thaumasite to form. The formation of synthetic

thaumasite can be represented by the following series of

stepped reactions:

1. 2Ca(OH)2= 2Ca2+ + 40H' (raises pH to 12.3)

2. Si02 + 2H20 = H4Si04 (formation of silicic acid)

3. 2H4Si04 = 2H2Si042‘ + 4H+_


(dissolution of silicic acid, at high pH)

4. 2CaC03= 2Ca2+ + 2C032" (dissolution of calcite)

5. 2CaS04•2H20 = 2Ca2+ + 2S042‘ + 2H20 (dissolution of


gypsum)

6. 6Ca2+ + 2H2Si04" + 2C02‘+ 2S02'+ 28H20 =


Ca6[Si (OH) 6]2*(C03)2*(S04)2*24H20

The difficulty lies in the need for carbonate and silica

both to be soluble in the same solution. Both silica and

carbonate solubility are strongly pH dependent, however, they

are affected in an opposite manner. Crystalline silica is

soluble in pure water at 60 ppm below pH—9. Above a pH of

the solubility increases dramatically and exceeds 6000 ppm at

the pH of lime saturation, 12.3. Amorphous silica, while

significantly more soluble than crystalline silica, behaves in

a similar way (Krauskopf, 1979). The relationship between

silica solubility and pH is shown graphically in Figure


100

To the contrary, calcite is highly soluble in acid

solutions and nearly insoluble under alkaline conditions. In

aqueous solutions, calcite dissolves through attack of carbonic

acid, as represented by the reaction:

CaC03 + H2C03 = Ca2+ + 2HC0f = Ca2+ + 2H+ + C032'

The log of the dissociation constant for the bicarbonate

reaction:

HC03' = C032' + H+

is -10.3 indicating that precious little C032 should be

available in the solution. From Le Chateliers rule, any

increase in the chemical potential of calcium will tend to

drive the calcite dissolution reaction back toward the left,

precipitating calcite.

It is also possible to produce thaumasite using C02

derived from the atmosphere as a source of carbonate (Kollman,

1978). Although this eliminates the necessity of dissolving

calcite the problem remains. Carbon dioxide dissolves in water

through the reaction:

C02 + H20 = H2C03,


101

the formation of carbonic acid. Carbonate is then released

through the dissociation of carbonic acid which, at a pH

greater than 10.5, occurs predominately through the reaction:

H2C03 = 2H+ C03‘.

This can be seen to be the same net reaction involved in the

dissolution of calcite. Again, any increase in the chemical

potential of calcium, such as from the dissociation of lime or

gypsum, should result in immediate precipitation of calcite.

Thaumasite is not known to have been prepared

synthetically prior to 1973 (Bensted, et. al., 19730. The

original synthesis was achieved from a stoichiometric solution

of gypsum, calcium silicate and calcite. The experiment was

conducted between 1°C and 4°C and required many months to

complete. The reaction products were thaumasite, calcite, and

aragonite. The secondary carbonate minerals formed from

reaction of atmospheric carbon dioxide with the solution.

Since that time, several other workers have synthesized

thaumasite through similar methods (Van Aaudt et. al., 1975,

Kollmann, Strubel, and Trost, 1976). The major differences in

technique are the source used for the silica. These range from

spurrite which gives a very slow reaction, to silica gel, the

fastest of the known procedures. Generally, any clay mineral

works well as clay minerals are broken down at pH greater

10.5 (Eades and Grim, 1950; Davidson et. al., 1965; Thompson,
102

1966). Even using the more soluble silica gel, however,

reactions require several months to reach equilibrium and

produce reasonably well-crystallized thaumasite.

Thaumasite does not appear to form at temperatures above

15 °C (Kollmann, 1978). This is likely a function of the

retrograde solubility of calcite. As discussed, calcite is

nearly insoluble at the pH required to dissolve appreciable

silica. This situation is apparently improved by decreasing

the temperature of the reactions which raises the solubility

of carbonate. Silica solubility, however, decreases with

temperature but remains considerably more soluble than

carbonate at lower temperatures and high pH.


103

CHAPTER 6

*v

METHODOLOGY FOR EXPANSIVE

POTENTIAL AND MINERAL SYNTHESIS EXPERIMENTS

nR.TECTIVES

Lime-induced heave through crystallization of ettringite

and/or thaumasite is a complex geochemical phenomenon. A

series of laboratory experiments were conducted under

controlled, repeatable conditions in order to further the

understanding of the physio/chemical mechanisms involved. The

experiments selected are listed below:

1. Concentration of sulfate in water leaching through soil


from Stewart Avenue

2. Measurement of expansive pressure

3. Laboratory duplication of the expansion observed on


Stewart Avenue
4. Growth of ettringite/thaumasite in soils from other areas

5. Synthesis of ettringite

6. Synthesis of thaumasite
7. Determination of approximate solubility of ettringite

8. Determination of approximate solubility of thaumasite

These experiments included attempts to synthesize and

isolate pure ettringite and thaumasite end members for use m

determining the approximate solubility of each mineral.

Solubility data can be used to calculate the mineral s


104

equilibrium constant, enthalpy of formation, heat capacity, and

free energy of formation. From these data the phase

relationships to other minerals can be determined and stability

diagrams can be calculated. Experiments were also designed to

investigate volume change under several geochemical conditions,

thaumasite/ettringite growth in various soils, nucleation, and

other supporting factors.

All of the water used in the experiments was de-ionized,

doubly distilled and had a conductivity of approximately 18

micro mhos. This water will be referred to herein as purified

water. All high (>8.0) pH solutions were mixed and Stored in

polyethylene containers. Glassware becomes significantly

soluble at a pH greater than 9. The ionized silica could

affect the reactions and would actually substitute for aluminum

in the ettringite structure.

GIJLFATE LEACHING

During the initial investigation of Stewart Avenue

(Chapter 3), it was postulated that water moved to

ettringite/thaumasite nucleation sites by capillarity. The

trilinear and stiff diagrams (Figures 18, 19, 20, and 22)

indicated an influx of sulfate ions, presumably carried m the

capillary water. In order to calculate the ion flux and

mass/volume relations it was necessary to obtain an estim


105

of the concentration of sulfate ions carried in the pore water

of the soil.

For this experiment soil sample S-9C from Stewart Avenue

was selected. This sample was collected at subbase elevation

in the unpaved shoulder adjacent to an area of severe damage.

Chemical analysis of the soil extract (50:1 water to soil

extraction) showed 6500 milligrams of soluble sulfate per

kilogram of soil which is not unusually high.

The soil was compacted into a one-inch (2.54 cm)


diameter, 24-inch (61 cm) long PVC plastic tube. The
soil mass was sufficient to provide a dry density of
107 pounds per cubic foot (pcf; 1.72 gm/cc) or 90-s
compaction relative to ASTM D1557-78. Moisture at
compaction was 10% which was optimum for this soil.
The tube was capped on both ends. One end was
attached to a source of purified water while the
other was connected to a tube draining to ^
collection flask. The system was pressurized at 5
pounds per square inch and maintained at that,
pressure for 20 days. During that time approximately
100 ml of water flowed through the tube.

The water sample was analyzed at the Water Resources

Laboratory of the Desert Research Institute, Reno, Nevada, and

found to contain 5710 mg/liter of sulfate. Since sulfate from

gypsum would only be soluble in pure water at approximately

1250 mg/1, it was hypothesized that most of the sulfate had

been derived from a more soluble sulfate bearing mineral. No

other sulfates had been noted in the original x-ray pattern.

For sample S-9C, however, such minerals would likely be present

in quantities too small for detection. As a check a sample of

soil s-9C was moistened to slightly above optimum and allowed


to air dry in a 250 ml beaker. After a period of a few days

a white efflorescence developed on the soil surface. X-ray

diffraction of this material demonstrated the presence of

thenardite, Na2S04. Sodium sulfate is extremely soluble, in

excess of 10,000 mg/1 sulfate. On this basis it was concluded

that much of the sulfate was probably derived from dissolution

of thenardite or its hydrated counterpart, mirabilite

(Na2S04•10H2O) .

EXPANSIVE POTENTIAL

Expansive Pressure

In order to roughly evaluate the potential expansive

pressure generated by growth of ettringite, a remolded sample

was placed in an Federal Housing Authority (FHA) soil swell

meter. Although the test procedure has no American Society of

Testing Materials (ASTM) or American Association of State

Highway Transportation Officials (AASHTO) sanction, it provides

a reasonably good measure of expansive pressure.

The meter consists of a rigid steel frame supporting


a proving ring and dial indicator above e
base. The soil sample is placed on a porous stone
in an aluminum ring that forms the samp e c •
The specimen, which is one-half inch (1. ^ . +-r>
and 2.5 (6.4 cm) inches in diameter, is compacted to
the desired density with a 5 -pound (2.27 kg) ramm ,
as used for ASTM D-698. Another porous ®t0"® l
placed over the compacted soil fol ov
permeable (drilled) aluminum disc. T h e proving ring
is then brought into contact with t h e disc_ by
adjustment of a threaded steel bar. Wit
107

completed, the apparatus can be submerged to allow


water to reach the soil.

The soil selected for this experiment was sample S-


22C from Stewart Street. Soil was mixed with 6.0%
hydrated lime (4.5% CaO) and densified to 105 pcf
(1.68 gm/cc) at 20% moisture. The sample chamber was
emersed in distilled water and readings taken
periodically.

After 70 days the expansive pressure had leveled off at

7535 pounds per square foot. As shown by Figure 24, the curve

of expansion pressure versus time indicates that the expansion

had reached the maximum value for the test conditions.

The soil specimen was removed from the meter and analyzed

by x—ray diffraction. The pattern indicated the presence of

a small percentage of ettringite or thaumasite. Since no other

new minerals were found it was concluded that the expansive

pressure was, most likely, the result of ettringite/thaumasite

growth. It is also possible, however, that amorphous calcium

silicate or calcium aluminum silicates were, at least, partly

responsible.

Volume Change Experiments

In order to fully evaluate and understand the expansion

mechanism it is a fundamental requirement to duplicate the

expansion measured in the field with a laboratory experiment.

The earliest attempts to measure potential volume change were

based on the assumption that thaumasite, the mineral identified


EXPAN SIVE P R E S S U R E IN POUNDS PER SQ U A R E FO O T
1000 ____________ 2 0 0 0 ____________ 3 0 0 0 ____________ 4 0 0 0 ____________ 5 0 0 0 ____________ 6 0 0 0 ____________ 7 0 0 0

of ettringite/thaumasite.
TIME IN DAYS
Figure 24 Expansive pressure from growth
109

in the field, had formed directly. Later research showed this

to be unlikely, however, the first volume change experiments

were conducted at low temperatures, in the range of 2°C to

22 °C .

The test apparatus selected was an AASHTO T-116 volume

change meter. Although this test has been replaced by more

popular methods it as the advantage of a relatively large,

thin, wide sample with a high surface area to volume ratio.

The brass volume change meter consists of a four-inch (10.2 cm)

diameter mold with a porus stone in the base. A 1 9/16-inch

(4 cm) thick soil sample is compacted into the mold with a

hydraulic press. The mass of the sample is selected so as to

attain the desired density within the given volume. Another

porous stone rests on top of the sample and is attached to a

shaft extending above the top of the mold. A fitted brass

cover slides over the shaft and caps the mold. This cover

stabilizes the shaft but still allows water to enter the top

of the mold. Water also enters the sample from the bottom

through holes in the base and the porous stone. A dial

indicator attached to a tripod rests on the lip of the mold and

measures vertical deflection of the shaft. The mold portion

of the apparatus can be submerged to allow influx of water to

the top and bottom of the sample.


110

T’vpfiriment One fVC-1'1

The four initial tests used sample S-9C taken from an area

of Stewart Avenue with moderate levels of both clay sized

particles and soluble sulfate. The grain size distribution and

soil chemistry of this material are given in Appendices 2 and

6, respectively.

The soil was mixed with 4.5% quicklime (CaO) at 4 %


over optimum moisture and densified to 90% compaction
relative to ASTM D1557. Four reaction solutions were
tried including purified water, water saturated with
Ca(0H)2, water saturated with CaS04, and water
saturated with Ca(0H)2 and CaS04. In all cases the
volume increase was initially rapid and then
gradually stopped within a few weeks. _ _ The
temperature of the experiment was initially
maintained at 2 to 4°C and later raised to 22 C.
Finally the temperature was brought to 45 C.
significant temperature related effects were
observed. The volume changes all were in the range
of 4% to 5.5%. The experiments were then postponed
until a better understanding of the geochemical
mechanism was developed.

Experiment Two TVC-21

Later evaluation and analysis on other aspects of the

research suggested that CaS04*H20 might be too insoluble to

yield the large quantity of sulfate ions required for the

observed heave. Also, research in Germany (Kollmann, et. al.,

1977; Kollmann, 1978) demonstrated that ettringite should have

grown first in the Stewart Street physio/chemical environment

and then slowly changed to thaumasite when the soil temperat

dropped below 15°C. As a consequence of this hypothesis,


Ill

subsequent testing was performed at room temperature in order

to facilitate ettringite growth.

Two experiments were set up for the second phase of


volume change testing. Again, a highly reactive soil
sample from Stewart Avenue (S-9C) was selected and
mixed and compacted with quicklime (CaO), as
previously described. One test apparatus was
submerged in purified water and the other, prepared
identically, was placed in a solution containing
13,450 ppm S042" from sodium sulfate.

After one month the sample in the sulfate solution had

stabilized at a volume increase of 1.5%. The sample saturated

in pure water had only expanded half that amount. At that time

the pure water bath was saturated with Ca(0H)2 and both samples

were allowed to react for another month. No significant volume

increase was observed in either sample and the experiment was

discontinued. X-ray analysis showed only a hint of ettringite,

or possibly thaumasite, in either sample.

Experiment Three fVC-31

The third phase of volume change experiments was an

attempt to reduce the number of variables involved. For this

purpose the mineralogically complex Stewart Street soils were

replaced by "synthetic soil."

The soil test specimen consisted of reagent gr


kaolinite clay blended with 6 . 0 weight percent
C a (OH)2 which is equivalent to 4.5* CaO. Two
specimens were then prepared by blen ing ^ ^
weight percent S0 42'. One sample used CaS04 2H20 an
the other Na2S04 as the sulfate source. ^he y
material was then thoroughly mixed with 40 „ water
112

(by weight) and allowed to cure in a sealed container


for 24 hours prior to compaction. Compaction was to
a dry density of 90 pounds per cubic foot (1.44
gm/cc). The samples were next allowed to rebound for
24 hours. During this time period, a rebound volume
increase of 5.7% occurred in each sample during the
first 11 hours and then stabilized.

The volume change meter containing the calcium


sulfate treated sample was submerged in water
saturated with CaS04-2H20 and Ca(0H)2. In this sample
the concentration of S042' was approximately 1350 ppm.
The Na2S04 bearing clay sample was submerged in an
aqueous Na2S04 solution containing 5500 ppm SO/' and
also saturated with Ca(0H)2. The 5500 ppm was
selected on the basis of the soil leaching
experiment.

This third volume change experiment was kept at a constant

temperature of 22°C for 60 days. By that time both samples

were continuing to expand, however, the rate of change had

become extremely slow. During the first thirty days the Na and

Ca sulfate samples had grown in volume 20.2% and 20.5^;

respectively. In the following thirty days the total volume

increase for sodium and calcium respectively, reached only 20.6

and 20.9%. After the experiment was terminated both specimens

were analyzed by x-ray diffraction. Both samples showed minor

ettringite or thaumasite. Although significant, neither sample

appeared to contain sufficient mineral growth to account for

all of the observed volume increase. Since kaolimte, and

particularly lime-treated kaolinite, is not expansive, it was

concluded that the rapid portion of the volume increase was,

at least, partially the result of overcompaction.


113

TCypariment Four (VC-4)

The fourth volume change experiment used a mixture


(by weight) of 64.46% 30 mesh Monterey sand, 28.2%
kaolinite, 6.6% Ca(OH)2 and 0.74% Na2S04. The sand
was added to simulate the ratio of sand to clays
observed in lime-treated native soil on Stewart
Avenue. The change greatly increases the effective
concentration of lime and sulfate relative to the
reactive (clay) components of the soil. The mixture
was blended with 25% (by weight) water and allowed
to cure in a sealed container for 24 hours prior to
compaction. Compaction was performed hydraulically
in a 2.50-inch (6.35 cm) diameter mold. The mass of
sample used resulted in a dry density of 110 pcf
(1.76 gm/cc) for a sample height of 2.0 (5.08 cm)
inches. The specimen was extruded from the mold and
placed in the 4-inch (10.2 cm) diameter volume change
test meter. The space between the mold and the
specimen was filled with hand—compacted 20 mesh
Monterey sand. The sand had previously been brought
to optimum moisture with an aqueous solution of
Na2S04 (5500 ppm S042') . No significant rebound was
noted in the 24-hour period between placing the
sample in the volume change meter and addition of
the 5500 ppm S042‘ (Na2S04) solution.

Initially the apparatus and the surrounding container of

reactant solution were placed in a constant temperature bath

at 25 ± 0.05 °C . After a period of 63 days the total volume

change was 3.1%. At that time the temperature of the bath was

increased to 40 ± 0.05°C and allowed to react for an additional

30 days. Virtually no additional expansion occurred beyond

that which could be attributed to thermal effects. Analysis

by x-ray diffraction showed only a hint of ettrmgite or

thaumasite.
114

yvpfiriment Five (VC-51

Since ettringite and thaumasite can readily be synthesized

and have been demonstrated to form in lime treated soils, it

was not clear why they could not be easily grown in the volume

change meter. This was especially curious in light of the fact

that ettringite/thaumasite was induced to grow in the previous

test for expansive pressure. For these reasons, it was

hypothesized that the problem might involve a chemical reaction

between the brass (copper and zinc alloy) test apparatus and

the sulfate or hydroxyl ions. A blue efflorescence had been

noted on the mold and was most likely a copper sulfate

compound.

Experiment five of the volume change testing was


conducted in non-reactive devices. Three 2 . 5 - m c h
(6.4 cm) cubes of clear styrene plastic were bored
with a 1.010-inch (2.565 cm) diameter round hole
a depth of 2.0 inches (5.08). A 1.00 inch (2-54 )
diameter PVC piston was then machined to a length
2.5 (6.4 cm) inches. A 0 .0 6 2 - m c h (0.15 £
diameter hole was drilled through the four sides of
the cubes to allow water access to the sample.

Sample charges containing lime-treatedkaolinite were


mixed with sodium sulfate (sample V C-SA), cabci
sulfate dihydrate (VC-5B), and both sodium and
calcium sulfate with calcium carbonate ( .
initial chemistry of the Phase 5 e x p e n m
summarized in Table 4.
115

Table 4

Initial Chemistry fin Grams') for


Volume Change Experiment 5

Sample

Component VC-5A VC-5B VC-5C

Kaolinite 8.75 8.61 8.18


Ca(OH)2 0.50 0.50 .50
Na2S04•10H2O 0.75 0.00 0.37
CaS04•2H20 0.00 0.89 0.45
CaC03 0.00 0.00 0.50
Total Mass 10.00 10.00 10.00

The charges were mixed with 40% doubly distilled,

deionized water and densified in a hydraulic press to 85 pcf

(1.26 gm/cc). The height of the compacted samples was one-half

inch (.13 cm). Each of the three samples were allowed to

rebound for 24 hours prior to being placed in the reaction

solution. Samples VC-5A and VC-5C were submerged in a covered

container filled with a 5500 ppm sulfate solution derived from

sodium sulfate. Sample VC-5B was placed in a sealed container

with a solution saturated with CaS04-2H20. Volume change was

monitored daily with a tripod-mounted dial indicator that

straddled the cube and measured the piston deflection. The

sodium sulfate reaction solution was kept at a nearly constant

ionic strength by monitoring conductivity and making the

necessary adjustments. The calcium sulfate solution was always

saturated by excess gypsum in the container. The temperature

of the experiment was allowed to fluctuate between 15°C and

25 C
116

The three samples were monitored for a period of 100 days

on one- to five-day intervals. The results are shown

graphically on Figure 25. The expansion ranged from a minimum

volume change of 28.8% for sample VC-5C to a maximum of 49%

for VC-5B. Sample VC-5B had the most rapid expansion but

slowed and was surpassed by VC-5A at 31 days. However, by

approximately 55 days the growth of sample VC-5A ceased at

41.4% while sample VC-5B continued to expand. Sample VC-5B was

monitored for another 56 days and finally ceased expansion at

a volume change of 55.0%. X-ray diffraction verified the

presence of significant ettringite/thaumasite in the three

samples. The x-ray pattern for sample VC-5C is included as

Appendix 14.

Experiment Six ('VC-6')

Samples VC-5A and VC-5C were removed after the 100-day

period. These were replaced by two native soil samples from

Stewart Avenue. Test specimen VC-6A consisted of soil from

Test Pit S-15. The soil was collected from native subgrade

beneath damaged lime-treated subbase. Test specimen VC-6B was

soil collected from Test Pit S-18. These soils were selected

because of their high percentage of clay and colloid sized

particles. The grain size distribution and chemistry of the

two soils (S-15C and S-18C) are presented in Appendices 1 and

6, respectively.
PERCENT VOLUME IN CREASE
118

Both soils were mixed with 4.5% by weight CaO and


compacted into the plastic volume change meters at
95% relative compaction and 4% over optimum moisture.
The compaction was performed hydraulically in two
lifts brought to equal pressure of approximately 248
psi. The meters were allowed to rebound for 8 hours
and were then placed in sealed containers containing
purifed water. After two weeks of no response, the
water was replaced with a solution saturated with
gypsum and calcium hydroxide. Response was sluggish
but both samples began expanding at a nearly steady
rate of approximately 1% a week. The temperature
fluctuated between 10°C and 25°C with little effect
on the reaction rate.

The expansion was monitored for over 6 months. The

results of the Phase 6 expansion testing are shown graphically

on Figure 26. The x-ray pattern for experiment VC-6A is

included as Appendix 15.

Summary of Experiments on Expansive Potential

The expansive pressure experiment was immediately

successful and demonstrated the ability of ettringite/

thaumasite to generate pressure sufficient to damage pavements

and lightly loaded structures. Volume change experiments,

while initially hindered by unexpected chemical reactions with

the test apparatus, also succeeded. The results are summarized

in Table 5.
PER CEN T VOLUME IN C R EA SE
120

Table 5

Sumarv of Experiments on Expansive Potential

Expansive Pressure

Added Enersion Tenp. Days Expansive Ettringite1


Soil Line Solution Range Duration Pressure Thaunasite

61 Ca(OH)j Distilled 20-25'C 106 7535 psf Weak


S-22C2
H,0

Volune Chanae

Added Enersion Tenp. Days ? Volune Ettringite1/


Line Solution Ranqe Duration Increase Thaunasite
Exoerinent2 Soil

4.5? CaO Distilled H,0 25-32'C 26 4.3 ND


VC-1A 2S-9C

4.5? CaO *Ca(0H)2 25-32'C 26 5.1 HD


VC-1B 2S-9C

2CaS04 25/32'C 26 5.5 ND


VC-1C 2S-9C 4.5? CaO

‘CaS0,+Ca(0H)j 25-32'C 26 5.4 ND


VC-ID 2S-9C 4.5? CaO

Distilled 20-22'C 63 <1 Trace


VC-2A 2S-9C 4.5? CaO
h 2o /
*Ca(0H)j

*HajS04 20-22'C 63 1.5 Trace


VC-2B 2S-9C 4.5? CaO
CaS04.2H304 22'C 60 20.9 Weak
VC-3A Kaolinite + 6.0? Ca(OH)2
0.05? + Ca(0H)j
so2; -
(CaS0.-2H20)
60 2 0 .6 Weak
VC-B Kaolinite + 6.0? Ca(OH)a *Ca(0H)a+5500 22‘ c

0.05? ppn S024


S02;(Na2S04) (N3aS04)

93 3.1 Trace
VC-4A 64.46? Sand + 6.6? Ca(OH)a 5500 ppn 25/40'C
28.2? S042“(Ha2S04)
Kaolinite +
0.74? KajSO,
100 41.4 Koderate
VC-5A 92? Kaolinite 5? Ca(OH)a 5500 ppn 15-25'C
+ 8? S02;(Na2S04)
N a jS O / lO H jO

156 55.0 Hoderate


VC-5B 90.6? 5? Ca(OH)a *CaS04.2Hj0 15-25'C
Kaolinite +
9.41
CaS04-2H30
121

Added Etersion Tenp. Days 1 Volune Ettringite /


Fxperinent3 Soil line Solution Ranee Duration Increase Thautasite1

86.11 51 Ca(OB)j 5500 ppt S0: 15-25'C 100 28.8 Moderate


VC-5C
Kaolinite (HajS04)
+ 3.91
Na2S04-10H20

S-15C123 4.51 CaO *CaS04.2H:0 10-25'C 180 26.4 Weak


VC-6A
+ Ca(0H)2

S-18C2 4.51 CaO ‘CaS04.2H20 10-25'C 180 20.6 Weak


VC-6B
+ Ca(OH)a

1 Relative intensity of Bain x-ray diffraction peak. ND = not detected.

2 Stewart Avenue soil sanple.


3 Experinents VC-1 through VC-4 used AASHTO T-116 brass volune change seter. VC-5 and VC-6 used non-reactive plastic devices.

* Solution aaintained at saturation with these conpounds.


122

■ETTRINGITE/THAUMASITE g r o w t h i n o t h e r s o i l s

Previous research has demonstrated that any clay mineral

will provide an adequate source of aluminum and silicon for

growth of ettringite and thaumasite (Kollmann, et. al., 1977;

Kollmann, 1978). Since numerous areas in the west and midwest

are known to have saline soils, the obvious question is why

there have been so few reported cases of lime-induced heave.

In order to study this problem 14 soil samples from various

locations were mixed with lime to see if thaumasite or

ettringite would grow. The sample locations of these soils are

described in Table 6. Grain size analysis, liquid limit and

plastic index for each of these samples is presented in

Table 7.

Table 6

sample Designations and Locations


of Various Soils for Ettrinaite/Thaumasite Growth

Samole No. nescriotion of Location

BR Black Rock Desert, Humboldt County, Nevada

BS Bonneville Salt Flats, Tooele County Utah


East of Wendover, Nevada, on Interstate 80

CS Carson Sink, Churchill County, Nevada, near


junction of Interstate 80 and U.S. Highway 95

MF Malheur River Floodplain, near Vale, Malheur


County, Oregon

ML Mono Lake, Mono County, California, south shore

NC Near Statesville, Iredell County, North Carolina

NR Northwest Reno, Evans Subdivision, Washington


Avenue, Washoe County, Nevada
123

Pyramid Lake, near Sutcliffe, Washoe County,


Nevada

SF Snake River Floodplain, Homedale, Owyhee County,


Idaho

SA Stewart Avenue, sample S-3C, native soil from


Test Pit No. 3., Las Vegas, Clark County, Nevada

SB Stewart Avenue, sample S-7B, native soil from


Test Pit No. 7, Las Vegas, Clark County, Nevada

SC Stewart Avenue, sample S-27B, native soil from


Test Pit No. 27, Las Vegas, Clark County, Nevada

WL White Lake (Playa) north side, Washoe County,


Nevada

YC Yolo Causeway, Interstate 80 overpass, Yolo


County, California

Table 7

Mechanical Analysis of Various Soils


for Ettrinaite/Thaumasite Growth

a* NC M PL SF SA SB SC WL YC
Sample m BS CS MF

Sieve
Size Percent by Weight Passing

100 100
3/4 97
1/2 93
86 100 95
3/8 94
100 100 100 72 98
No. 4 100 91
100 94 98 100 98 55 95 100
No. 10 100 100 100 100
92 86 95 93 96 94 83
No. 40 99 99 99 99 81 78
80 77 90 42 92 81 83 82
No. 100 98 97 94 98 60 66
73 65 80 38 86 66 74 81
No. 200 97 95 90 94 59 50

Hydrometer
Analysis

22 25 21 15 5 38 44
0.005mm 72 49 35 29 12 14 39
7 6 2 40 1 11 23
0.001mm 31 26 22 11 4 3 21

Liauid 21 47
36 58 51 37 57 52 53
Limit 46 52 44 <45

Plastic 27 6 25
19 42 2 11 43 27
Index 20 22 20 NP

* Weakly cemented with calcium carbonate


124

The soils were mixed with 5% by weight CaO. Enough water

was added to the mixtures to form a thick slurry to allow easy

movement of the ions and, presumably, accelerate the reactions.

All samples were stored in a refrigerator at a temperature of

4 °C +i° C . Each was analyzed by x-ray diffraction after a

reaction period of one year. None of the samples showed

definite thaumasite or ettringite. Seven samples showed

extremely weak peaks in the area of the main thaumasite/

ettringite peak. However, these could not be reasonably

separated from the background.

Three of the soil samples (S-3C, S-7C and S-27C) were from

areas of Stewart Avenue where thaumasite is known to have grown

and damaged lime-treated subbase. The question was, therefore,

why thaumasite or even ettringite, apparently, did not grow in

the experiment. It was hypothesized that, perhaps, these

minerals were simply not present in sufficient quantity to be

detected by standard x-ray powder diffraction. On this

assumption, the maximum percent of ettringite that could be

grown was calculated for Sample S-3C (untreated native soil)

from Stewart Avenue. Chemical analysis of this sample had

shown it to contain 24,500 mg of sulfate per 1000 grams of soil

(Appendix 6 ). On the basis of stoichiometry, specific gravity,

and molar volume it is readily demonstrated that the quantity

of sulfate present could only result in a maximum growth of 4%

by volume of ettringite. This assumes that all of the sulfate


125

is gypsum and that it all reacts to form ettringite. In fact,

the amount of gypsum present in the damaged subbase is often

about equal to that prior to lime treatment. Clearly, not all

of the sulfate reacts. It was concluded, therefore, that even

in high sulfate soils, ettringite/thaumasite may not grow to

the 3-5% by volume necessary for detection by x-ray powder

diffraction. This concept also led to a detailed analysis of

mass-volume relationships presented in Chapter 7.

TC-H-ri naite/Thaumasite Concentration

by the Heavy Liquid Method

Since ettringite/thaumasite might be present in quantities

too small to detect, an attempt was made to concentrate these

minerals in one of the samples using a heavy liquid. Sample

WL was selected as it had shown a very weak, possible peak at

d = 9 .65 X. The heavy liquid chosen was tetrabromoethane which

had a measured specific gravity of 2.95 gm/cc. This was

thinned with methanol to achieve the desired specific gravity

of 2.5. A liquid at this density would allow the lighter

minerals such as gypsum and ettringite/thaumasite to float

while letting quartz, clay, and calcite sink. The net effect

would be to increase the relative percentage of ettringite/

thaumasite in the float. As it turned out heavy liquid

separation was messy and probably contaminated the

ettringite/thaumasite anyway. Many minerals appeared to

dissolve into the liquid and particles that should have sunk
126

(such as quartz) were often kept afloat by surface tension or

trapped air bubbles.

The float was removed, rinsed, dried at room temperature

and x-rayed. The result was a distinctive ettringite/

thaumasite peak in the area of d=9.65 *. The small percentage

hypothesis was concluded to be sound, however, concentration

by elutriation was thought to be a better technique.

Ettrinaite/Thaumasite Concentration by Elutriation

The 14 lime-treated soil samples were air dried at a

maximum temperature of 48°C. Approximately 35 grams of sample

was passed through a No. 200 sieve and elutriated.

In the elutriation procedure the sample was placed


in a 36-inch (91.5 cm) long, one inch (2.5 cm) I.D.
clear plastic tube. Air pressure was applied from
the bottom until the finest particles }ust started
to exit the top of tube. Once the finest particles
were removed the air pressure was increased very
slightly. The material coming out of the tube
this increased pressure was captured on a gla
slide coated with petroleum jelly.

If all of the grains were individual minerals of the same

shape and size those exiting the tube at the lowest air

velocity would be those with the lowest density. Ettringite/

thaumasite, with a specific gravity of approximately 1.8 is

considerably less dense than the clay or other salt minerals

present. Gypsum, for example, is the next lightest min

present in significant quantity and has a specific gravity of


127

2.32. Unfortunately, the minerals are complexly intergrown and

do not separate readily. It is possible, however, to markedly

increase the relative concentration of low density minerals

using the elutriation technique.

Following elutriation x-ray diffraction showed that 6 of

the samples clearly contained ettringite/thaumasite while 4

definitely did not. The remaining 4 (very weak) probably

contained some of the mineral, however, the peaks were not

entirely separable from the background. The presence of common

sulfate minerals was also noted, as summarized in Table 8.

The x—ray diffraction pattern for sample CS is included as

Appendix 16.

Table 8

Results nf X-Rav Analysis of Other Soils


for Ettrinaite/Thaumasite Growth

Sample Ettrinai te/Thaumasite Peak Sulfate Mineral Peaks.

Very weak Very Weak


BR
Very strong None
BS
Strong Strong
CS
None None
MF
None None
ML
None None
NC
Strong Weak
NR None
PL Moderate-strong
Very weak None
SF
SA Weak-moderate
Weak None
SB None
SC None
Strong Very weak
WL None
YC Very weak
128

gVNTHESIS OF ETTRINGITE AND THAUMASITE

As discussed in Chapter 5 both ettringite and thaumasite

have been previously synthesized by a number of researchers.

The problem, then, is not how to synthesize the minerals, per

say, but how to synthesize and isolate them as the purest

minerals possible. This turned out to be a non-trivial, and

not entirely successful, pursuit. Very few of the experiments

were initially successful. The methodology had to be developed

through an evolutionary trial-and-error process. The results

of each experiment were analyzed and compared so that the next

experiment could be designed to answer a specific question or

solve a particular problem.

Synthesis of Ettringite

Ettringite is readily precipitated by mixing a saturated

solution of calcium hydroxide with an undersaturated solution

of aluminum sulfate. When the aluminum sulfate solution was

too concentrated, gypsum was found to precipitate with the

ettringite. In addition, any contact between the calcium

hydroxide solution and carbon dioxide derived from either the

water or the atmosphere resulted in co-precipitation of calci

with the ettringite. Previous researchers had addressed the

gypsum problem (Kollmann, 1978) however, the difficulty with

calcite had not been discussed. For solubility determinations

the purest possible compound is always desirable. Once


129

and/or gypsum form with ettringite the minerals are virtually

impossible to separate in a practical manner.

•Experiment One

Initially the calcium hydroxide solution was prepared


by placing 4.0 grams of Ca(0H)2 on a weighed filter
paper. One thousand milliliters of double distilled,
deionized (purified) water was then passed through
the filter under vacuum. By weighing the residual
lime it was determined that 1.8 grams had dissolved
in 1000 ml of water at 27°C. The amount of aluminum
sulfate that could be added to 1000 ml of saturated
calcium hydroxide solution was then calculated
assuming only 1.5 grams of lime had dissolved. This
lower concentration was used as a safety factor to
help prevent oversaturation of the calcium hydroxide/
aluminum sulfate solution with gypsum. From the
stoichiometry of ettringite, it was calculated that
0.00507 moles (2.89 grams) of Al2(S04)3•18H20 would be
needed to precipitate 1.5 grams of Ca(0H)2 as
ettringite. A solution containing 80 grams per
liter Al,(S04)2*18H20 was made up, again with purified
water. It was calculated that 36 ml of this^sulfate
solution would precipitate the 1.5 grams of (Ca(0H)2.
The 36 ml of sulfate solution were added via a buret
to the hydroxide solution at a rate of 2 3 ml per
minute. Too rapid a mixing rate also results m th
precipitation of significant quantities of 9 YPSum
(Kollmann, 1978). This was confiraed here
accidentally when the mixing rate increased due to
a faulty buret.

As predicted in the literature, a whi^ e ' mil^


precipitate formed immediately Wlth ®ach dr°i\ °
sulfate solution. The solution was th^ filtered,
open to the atmosphere, rinsed with purifi ,
and dried at 35°C.

The dried precipitate was analyzed by x-ray diffraction

using a Phillips Norelco diffractometer. Results of the early

experiments showed poorly to well crystallized ettringite in

approximately equal proportions with well crystallized

Minor gypsum was also present.


130

-E x p e r im e n t Two

It was next assumed that the source of the carbonate was

atmospheric carbon dioxide dissolving in the three aqueous

solutions during preparation and mixing. As a consequence,

the next phase of experiments involved preparation and mixing

of all solutions in an atmosphere of nitrogen.

The calcium hydroxide solution was prepared by adding


4.0 grams of Ca(OH)2 to a 1000 ml flask of purified
water. Air was kept from entering the flask by a
constant, low pressure purge with nitrogen. The
solution was allowed to equilibrate and settle for
at least 48 hours prior to filtration. This was done
in an effort to decrease filtering time and,
consequently, the potential for atmospheric
contamination. During the filtering, a positive
nitrogen purge was applied to the top of the
apparatus, again to minimize contact with C02. m e
solution was stored in a polyethylene erlenmeyer
flask flushed with nitrogen prior to sealing with a
rubber stopper lightly coated with petroleum lelly.

The aluminum sulfate solution was prepared as m


Phase One except that the flask was flushed with
nitrogen as the water was added to the salt. Once
again 36 ml of sulfate solution were added to the
calcium hydroxide solution at a rate of ?h
minute. This time both the buret Gontaining the
aluminum sulfate solution and the PolY ® ^ y l e n
reaction vessel were maintained under a constant
nitrogen purge. The reaction vessel was mixed using
a magnetic stirring device during the en
the aluminum sulfate solution was b em g added. The
precipitate was separated by filtration ae
this time under the shelter of a nitrogen purge.

X-ray diffraction of the precipitate revealed the presence

of poorly crystallized ettringite with an estimated 25% calcite

(Appendix 17). It was found that the crystal structure of the


131

ettringite improved greatly if the precipitate was allowed to

age at least 48 hours prior to filtration. Additional aging

did not appear to improve crystalinity. Also, the crystal

structure showed significant improvement when the precipitate

was dried at 25-30°C. When the drying temperature exceeded

approximately 40°C crystal structure was very poor. One

sample, accidentally dried at 82°C, showed only calcite.

Several attempts were then made to remove calcite from

the precipitate using an acid rinse. The first effort used

purified water saturated with C02. It was hoped that the

carbonic acid would be strong enough to dissolve calcite, yet

too weak to have a significant effect on ettringite. After

drying at room temperature, x-ray analysis revealed that

ettringite had been completely replaced by calcite and gypsum.

A second attempt using 0.1 molal HCl resulted in dilution of

the entire precipitate. As a consequence, the acid rinse

approach was abandoned.

Experiment Three

The source of the C02 contamination remaining, in spite

of the careful nitrogen purge, was eventually found to be m

the deionized, doubly distilled water that had been used for

all of the solutions.

Tn 0f experiments the water was boiled for 30


S n S l s an/pIureS, again with a nitrogen(Purge. into
a flask containing 4.0 grams of lime. After cooling
132

and filtration under a constant nitrogen purge the


calcium hydroxide solution was placed under vacuum
and agitated to help remove remaining dissolved
gases. Next the solution was bubbled with nitrogen
to displace as much of the remaining C02 as possible.
The flask was then sealed and stored. The rest of
the procedure was performed as previously described
except that water for the aluminum sulfate solution
was also boiled, subjected to vacuum and bubbled with
nitrogen.

When this precipitate was x-rayed it showed large, well

defined peaks for ettringite indicating both a high percentage

of ettringite and good crystal structure (Appendix 18).

Although calcite was still present, it appeared to be a

relatively minor component. Estimates based on x-ray

diffraction indicated that the purest ettringite synthesized

contained less than 1% calcite by weight. In this semi-

quantitative analysis an internal standard was used by adding

a known weight percentage of calcite to the ettringite and

comparing the effects on the area of the main calcite peak.

The analysis was run with -pure1' ettringite, ettringite plus

2% calcite, and ettringite plus 5% calcite. The weight percent

calcite in the "pure" ettringite was then estimated

graphically, as shown in Figure 27.

It was concluded that the synthetic ettringite from

experiment three was quite good and would be suitable for the

purpose intended. Efforts to obtain a more pure product would

be time consuming and not really justified. The ettringite

synthesis experiments are summarized in Tabl


133

in synthetic ettringite.
P E R C E N T C A L C IT E CALCULATED FROM PEAK AREA
134

Table 9

Summary of Ettrinaite Synthesis Experiments

Saturated
Ca(0H)2 Solution A 1 2( S 0 J 2*18H20 Mixing n2
Experiment f0.024 m( Solution (0.120 mi Water Kate Purge Results

1000 ml 36 ml Double 2-3 Ho Ettringite with


ES-1
distilled; ml/min much calcite;
deionized minor gypsum.

1000 ml 36 ml Double 2-3 Yes Ettringite with


ES-2
distilled; ml/min less calcite than
deionized ES-1. No gypsum
detected.

36 ml Double 2-3 Yes Excellent


ES-3 1000 ml
distilled; ml/min crystalline
deionized; ettringite with
boiled; trace calcite
bubbled with an^ no gypsum
nitrogen after detected
mixing with CaO
and cooling.

Synthesis of Thaumasite

Experiment One

Attempts to synthesize thaumasite began in July 1983

From the procedures described by K o l l n a n n (1978)'


stoichiometric mixtures of Ca(OH)2/ a 3, 4 ^
and Si02 were added to purified w a t e r * ° off s i l i S
aqueous solution. Four different sources^of sil
were used in the initial experiments, as shown
Table 10.
135

Table 10

Thaumasite Synthesis
Initial Chemistry of Experiment One Group

Silica Source

Group

TS1-1 #325 mesh Si02 glass spheres


TS1-2 -200 mesh Ca2Si06 (amorphous)
TS1-3 -100 mesh Type I Portland cement
TS1-4 -200 mesh fly ash

Initial Chemistrv1— Grams

H 20 H 20
Sample CaS04-2H20 CaC03 Ca(0H)2 Si02 Initial Final2

TS1-1A 34.4 20.0 14.2 24.0 100 290


34.4 20.0 14.2 24.0 75 75
TS1-1B
. 34.4 20.0 14.2 24.0 50 550
TS1-1C
34.4 20.0 14.2 21.2 125 125
TS1-2A
34.4 20.0 14.2 21.2 150 200
TS1-2B
34.4 20.0 14.2 21.2 175 375
TS1-2C
20.0 14.2 60.0 150 250
TS1-3A 34.4
20.0 14.2 17.0 150 250
TS1-4A 34.0

1Stoicionetric for 1/10 mole of thaumasite


"Included water added after approximately thirty days reaction
time.

The solutions were mixed in 1000 ml erlenmeyer flasks


and placed in a Haake Model NK-22 water bath
maintained at a constant temperature of 30 1 C.
Various initial water contents were tried in an
effort to find the optimum consistency. More water
was added to some of the samples after 30 days m
order to maintain a slurry-like consistency.

After a period of three months each of the samples was

analyzed by x-ray diffraction. The powder method employed is

only capable of detecting crystalline solids present m

quantities greater than 3 to 5%, by volume, of the

sample. Thaumasite was detected only in samples '

TS1-2B, TS1-3A (Appendix 19), and TS1-4A.


136

Portland cement has been commonly been used as the silica

source in previous research (Kollmann, 1978) since the silica

in cement is particularly reactive at high pH. Type I cement

typically contains approximately 20% silica, by weight. Fly

ash is derived from burning of coal in coal-fired power plants.

The resulting ash contains a high percentage of high

temperature, amorphous silica that is unstable at low

temperature and high pH. Typically, fly ash contains 62 to 82

weight percent Si02 + Al203 + Fe203, with silica and aluminum

being the dominant components. Fly ash is frequently used in

concrete as a substitute for up to 30% of the Portland cement.

The fly ash increases the concrete's resistance to

attack (M e h t a , 1986).

Of the three successful experiments, the Portland cement

produced the highest quantity of thaumasite (most complete

reaction) as well as the best orystal structure. In all of


CaS04-H20 were still
these experiments Ca(OH)2, CaC03, and
was amorphous, could
present. The silica, which in all cases
It is difficult to
not be detected by
x-ray diffraction.
, . „ v,., v-r-av diffraction, without
quantify mineral proportion Y
, using this standard
using an internal standard. Even usi g
quantitative determinations are difficult in multi-component

systems. Since one of the components in the thaumasrte

synthesis is amorphous only crude approximations are possible

internal standard was used. Based simply on the


anyway so no
137

area of the main peaks, estimates were obtained for the

percentage of thaumasite present (relative to crystalline

compounds) in the four samples:

Table 11

Estimated Percentage of Thaumasite


In Experiment One Samples

Sample No. % of Thaumasite

TS1-2A <5
TS1-2B <5
TS1-3A 20
TS1-4A 15-20

Following the initial x-ray analysis samples TS1-1A and

TS1-2B were "seeded" with thaumasite from sample 0-2B from

Owens Avenue. The portion of the sample used in the seeding

consisted of a minus 50 mesh plus 60 mesh fraction that had

been elutriated to concentrate the low density mineral,

thaumasite. Sufficient air pressure was applied to the bottom

of the elutriation tube so that particles just reached the top

of the tube before falling back down. The sample was allowed

to mix in this manner for approximately 5 minutes. At that

time the air pressure was slowly increased until a tiny amount

of the soil was blown out of the top the tube. A filter bag

was then placed over the top of the air tube. Another sl'g

increase in air pressure raised a portion of the particles int

the filter bag. Experimentally it was found that separating

approximately 20% of the soil in this manner produced the

highest concentration of thaumasite. Although far from pure,

the product was suitable for seeding purposes. The elutriation


138

technique, however, was shown by this experiment to be

incapable of separating even reasonably pure thaumasite from

the Stewart/Owens Avenue soils.

Experiment Two

Samples TS1-1B, TS1-1C, TS1-2A, and TS1-2C were abandoned

following the initial x-ray analysis.

New samples were prepared using Portland cement


(TS2-3B and TS2-3C) and fly ash (TS2-4B and TS2-4C)
(Table 12). Samples used stoichiometric proportions
of the component minerals with an excess of Portlan
cement or fly ash. All four samples were seeded with
0 5 gram of thaumasite concentrate prepared by
elutriation of sample S-7B from Stewart Avenue.

Table 12

Thaumasite. Synthesis
al Cheini strv of Pxneriment Two Group
Source Initial Thaumasite
H20 Seeding*
Sample CaS04'2H20 CaC03 Ca(0H)2 Si02
initial Ohemistrv--Grams

34.4 20.0 14.2 60 1000 0.5


TS2-3B
10 7.1 30 1000 0.5
TS2-3C 17.2
10 7.1 9 500 0.5
TS2-4B 17.2
10 7.1 9 1000 0.5
TS2-4C 17.2

*Sample S-7B, Stewart Street.

The progress of the reactions of Experiment Two and

remaining groups from Phase One were monitored by x-ray

diffraction for over three years. The results are summarized

in Table 13. Again, quantitative interpretation from x-ray

patterns containing multiple minerals is, at best, approximate

and impossible when amorphous compounds are present.


139

addition, a variety of sample preparation techniques were

employed, further complicating the data. The important point

shown by the results is that changes in the areas of the

primary x-ray peaks, and presumably in the proportions of the

minerals, did occur and that the percentage of thaumasite

increased with time.

Table 13

Estimated Percentages of Thaumasite in R emaining


Samples of Experiment One and Experiment Two**

Sample TS1-3A

Thaumasite Gvosum Calcite Portlandite


Date

20 25 30 25
09/24/83
25 10 15 50
11/06/83
15 15 40
12/03/83 30
10 15 40
01/08/84 35
Sample TS2-3B

15 20 40
01/08/84 20
15 40 5
03/04/84 40
0 5-10 50
40-45 10
* 80 0 10

Sample TS2-3C

40 30 20
11/06/83 10 15
03/04/84 40 5 40

Sample TS2-4A

25 55 0
09/24/83 20
Sample TS2-4C

30 15 40
11/06/86 15

* Relative to crystalline compounds.


** Petroleum jelly slide preparation.
140

Fvpftriment Three

In October 20, 1985, six new experimental specimens were

prepared using silica gel as the source of Si02. Samples were

prepared using the following proportions:

Table 14

Thaumasite Svnthesis-Initial Chemistry of


Experiment Three

Initial Chemistry1 (Grams)

Sample CaC03 Ca(0H)2 i4’1/2H20 Si02 gel H 20

4.00 2.96 5.80 2.40 50


TS3-A
8.00 2.96 5.80 2.40 50
TS3-B
4.00 5.92 5.80 2.40 50
TS3-C
4.00 5.92 11.60 2.40 50
TS3-D
4.00 5.92 5.80 4.80 50
TS3-E
2.96 5.80 2.40 50
TS3-F2 4.00

^ to ic i o m e t r i c proportions for 1/50 mole


20pen to atmosphere

X-ray diffraction analysis of each sample was performed

after 30 days but no thaumasite was detected in any sample.

The reason for this was thought to be the relatively large

particle size (#16 mesh) of the silica gel.

New batches of Samples TS3-A through TS3-G for


Experiment Three were prepared on Novembe ' tg
This time the silica gel and all other
were pulverized to -200 mesh. Component proportions
were as previously shown (Table 14).

X-ray diffraction analysis on January 18, 1987, showed the

following estimated percentages of crystalline compounds.


141

Table 15

Results of Thaumasite Synthesis


ExDeriment Three

Estimated Percentaae of Crystalline Compound

Sample Thaumasite Gypsum Portlandite Calcite

TS3-A 40 35 0 25
TS3-B 30 15 0 55
TS3-C 45 20 0 35
TS3-D 30 50 0 15
TS3-E 15 55 0 30
TS3-F 25 20 0 55

The x-ray pattern for Sample TS3-C is shown in Appendix 20.

Experiment Four— Two Solutions

Since ettringite can be readily precipitated by mixing two

appropriate solutions it was hoped that a similar procedure

might work for thaumasite.

One thousand milliliters of a solution saturated with


Ca(OH)2/ Si02 and Ca(S04) •1/2H20 was prepared by
placing 10 grams of each in a flask with purifie
water. The solution was allowed to equilibrate for
two weeks prior to slow addition of a second solution
of pure water saturated with C02.

The two solutions were stored and mixed at a


temperature of 3.5°C. The C02 solution was added to
the more complex solution with a buret at a rate of
one ml per minute. A white precipitate formed
immediately on contact between the two solutions.

X-ray diffraction analysis showed the precipitate to

consist of an estimated 70% calcite and oO-s gypsum. No

thaumasite was detected and the concept was pursued no further.


142

Experiment Five— Separation/Osmosis

The hypothesis was made that thaumasite might

preferentially nucleate on a single compound. If this was

true, pure thaumasite might be synthesized by separating the

nucleation sites from the remaining compounds by a permeable

membrane.

In an attempt to ascertain which, if any, compound


is the preferred thaumasite nucleus, each was
separated from the remaining components by a filter
membrane. The compound being studied was placed in
the top of a Nalgene 125 ml membrane filter unit with
125 ml of C02 free, purified water. The remaining
compounds were placed in the bottom of the cham er
with 150 ml of purified, C02 free water. The
components were of stoichiometric proportion f /
mole except that the compound isolatedL in the top
compartment was only sufficient for 1/100 mole. This
was to promote as complete a reaction as possible y
keeping the compound under investigation in relative
short supply. The experiment was allowed to stand
at 3.5°C for 6 months.

Samples from each of the two chambers were analyzed by

x-ray diffraction. The results obtained are summarized in

Table 16.
143

Table 16

Results of Thaumasite Synthesis


in Experiment Five

Amorphous
Sample # Thaumasite CaCO. CaSO.'l /2H20 carom. sio,__
TS5-At + +*
TS5-Ab + -
TS5-Bt +* -
TS5-Bb + +
+ - -* -
TS5-Ct +
TS5-Cb + +
_ -*
TS5-Dt
TS5-Db + + +
TS5-Et + — — +*
TS5-Eb + +

= top compartment, b = bottom compartment


* ft

component isolated in top compartment


+ = detected by x-ray diffraction
- = not detected by x-ray diffraction

On the basis of these results it appears that thaumasite

does nucleate preferentially as Ca(OH)2. However, there did

not appear to be any great potential for synthesis of pure

thaumasite. The reactions were incomplete and only a small

amount of thaumasite formed.

Experiment Six— Kinetic Exchange

German research (Kollmann, et. al., 1977, Kollmann, 1978)

has demonstrated that a high pH solution containing calcium,

aluminum, sulfate, silica and carbonate will grow e t t n n g i t e

above a temperature of 15 °C but not thaumasite. Below this

temperature, and particularly in the range oj.

ettringite grows first and then slowly changes to thaumasite


144

through a gradual transition. This suggests that, while

ettringite may be kinetically favored over thaumasite,

thaumasite may be thermodynamically more stable below 15 C.

This hypothesis opens the possibility of converting reasonably

pure ettringite to thaumasite.

In experiment six ettringite was prepared by the best


methods previously developed and described in this
research using vacuum filtration and a nitrogen
purge. The samples were not allowed to dry and were
not aged. It was thought that both aging and drying
could "tighten" the crystal structure and make the
transition more difficult.

The ettringite was placed in the top compartment of


a Nalgene 125 ml membrane filter unit. The bottom
compartment contained the other components necessary
for the conversion. Both compartments were filled
with de-ionized, distilled water. The following
experiments were set up:

Table 17

Initial Chemistry of Thaumasite Synthesis


Experiment Six

initial Chemistry-Gms

Top Chamber Bottom Chamber

Sample Ettringite CaCO. Ca(0H)2 Si02 gel Kaolinite**

5.0 1.8 * 0.5 0.0


TS6-EA
TS6-EB 5.0 1.8 0.5 0.5 0.0
TS6-EC 5.0 1.8 0.5 0.5 0.0
5.0 1.8 0. 5 0. 0 2. 0
TS6-ED

* pH brought to 8.0
** API standard {9

The ettringite from the isolated (top) chambers of Samples

TS6-EA and TS6-EB was x-rayed after 20 days. Although each

sample showed some changes in its x-ray pattern,


145

clear what these changes meant. The samples were allowed to

react for over 30 months before they were checked again. The

later x-ray patterns of Samples TS 6 -EA, TS 6 -TB, and TS 6 -EC were

distinctive for ettringite. In fact, the ettringite seemed

more pure and more crystalline than when the experiment was

started. Sample TS 6 -Ed, however, showed definite double peaks

in the areas of d = 9.7J and d = 5.0J indicating some reaction.

The two peaks that really differentiate thaumasite from

ettringite (d = 4.58J and d = 3.792) were not present and the

approach was deemed unsuccessful. The x-ray pattern for sample

TS 6 -ED is included as Appendix 21.

SUMMARY

Six major experiments were conducted to synthesize

thaumasite pure enough for solubility/thermodynamic

determinations. Although most of the experiments did yield

significant thaumasite, none produced thaumasite sufficiently

isolated from other components that it could be separated and

purified. The effort was abandoned following experiment six.


146

CHAPTER 7

THE THERMODYNAMICS OF ETTRINGITE DISSOCIATION

experimental methodology

The solubility of ettringite was determined at 10 C, 25 C,

and 40°C. For these experiments the purest ettringite

previously synthesized was placed in each of two polyethylene

erlenmeyer flasks. A sample containing gypsum was also

prepared. Since its solubility is well established, gypsum

was used as a check on the experimental techniques. Finally,

a dilute solution containing calcium chloride was prepared.

A calcium chloride solution would turn milky when exposed to

carbon dioxide and serve as a visual indication of C0 2

contamination.

A magnetic stirring rod and doubly d ^ t i l l e ^


deionized water were added to each f_lask. 'r*e_ ^ e r
had been boiled for 30 minutes and placed in a se<al
polyethylene bottle to cool m order to minimize the
amount of dissolved C0 2 present. The composition
solutions are shown in Table 18.
147

Table 18

Composition of Solutions used in Determininq


Ettrinaite Solubility

Sample ET-B ET-C** G CaCl2

Grams of Components

Synthetic
Ettringite* 0.400 0.400 0.400 0.00 0.0

Gypsum 0.0 0.0 0.0 2.00 0.0

Calcium
0.0 0.0 0.0 0.0 8.800
Chloride

800.0 800.0 800.0 800.0 800.0


H 20

* Synthesized in ettringite synthesis Experiment Three.


** Added after measurements at 10° and 40 C.

Air in the flask was displaced with a nitrogen purge prior

to and during the time the water was added. The flasks were

then sealed with a rubber stopper lightly coated with petroleum

jelly. Since the flasks were polyethylene excessive stirring

would have worn a hole through the bottom. The magnetic

stirrers were set with an electronic control that mixed each

flask for a twenty-minute interval every eight hours,

addition the solutions were stirred for ten minutes prior to

any measurements or sampling. Several variations

stirring and settling times prior to conductivity measurement

were tried. No significant variation was noted, as long as the

solutions were stirred briefly for a few minutes prior to the

measurement. Even if the solution was not stirred the


. „ oi irjhtlv lower than when the
conductivity readings were only si g Y

sample was stirred.


148

It was assumed that ettringite would have normal, as


opposed to retrograde, solubility. Since it is
kinetically easier to dissolve a mineral than it is
to precipitate it from a slightly oversaturated
solution, it was decided to proceed from the lowest
to the highest temperature. The temperature of the
solutions was maintained constant using a Haake Model
NK-22 constant temperature bath. Since the unit
itself would not hold more than one flask, an
auxiliary, insulated and covered bath was connected
to the main controller. All sample flasks were
placed in the larger bath, along with the standard
solutions for pH and conductivity.

Unfortunately, the temperature controller in the bath was

malfunctioning such that when set at 10 C, temperatures

fluctuated between 3°C and 20°C. Occasionally the temperature

would rise to 25°C. A phone call to the distributor suggested

a faulty temperature controller. A new controller resulted m

a false sense of improvement. The temperature would seem to

remain nearly constant for a period of several days,

fluctuating only one to two degrees. This resulted m a

considerable waste of time as it continually appeared that "one

more adjustment" should produce the desired stability.

Finally, after a frustrating period of trial and error, the

constant temperature unit was returned to the Desert Research

Institute for repair. Approximately eight weeks later the unit

was returned and operated perfectly from that time on

temperature could be easily set and maintained within ±-0.1°C.

The temperature controller was a d j u r e d to the


desired temperature and the samples a l l o w e d t o
equilibrate. Initial “Samples!
a
cno nduPctivTtye w“ dL a ° s ureTwitnfa Barnstead Model PM
149

70CB conductivity bridge. Measurements of pH were


taken with a Beckman Model 4500 digital pH meter
using a Corning 476051 electrode. The pH meter was
standardized using NBS traceable buffer solutions of
pH 7.00 for the gypsum and calcium chloride solutions
and pH=10.00 and 11.00 for the ettringite solutions.
Conductivity was standardized for the 10 C
measurements with a 140 millimho solution of KCl.
A 280 millimho KCl solution was used for the
experiments at 25"C and 40 C.

Progress toward equilibrium was monitored


periodically by measurement of conductivity.
Measurements of pH were minimized in order to
minimize contamination of the solution from K, Na,
and Cl in the electrode. All measurements of
conductivity and pH were made under a constant
nitrogen purge. Results are summarized below for
the three selected temperatures.

Table 19

Conductivity finKicromhosl and pH of


Ettringite Solubility Experiment at lO'C

ET-10A ET-10B G-10 CaCl^lO


Sample. Conductivity pH
Conductivity pH Conductivity pH Conductivity pH
Date
10.38 825micromhos 6.27 819 7.51
08/22/87 91 10.36 91
269 _____
1,450 — 1,450 . . .

09/02/87 269 —
6.15
407 11.12 1,920 6.13 1,920
01/21/88 404 11.12
— 403 — - 1,980 — 1,760 . . .

02/01/88 406 ___ 1,860 —


413 1,980 6.26
02/06/88 415 —
1,860 6.17
11.20 414 11.20 1,980 6.26
02/07/88* 416

* Samples collected for analysis


150

Table ?0

Conductivity finKicromhosl and cH of


Fttrinaite Solubility Element at 25*C

ET-25A ET-25B G-25 CaCl525


Conductivity pH Conductivity pH Conductivity pH Conductivity pH
Date

481 ___ 476 — 1.94 — 1.93 —


02/15/88 ...
464 — 466 — 2.07 — 1.93
03/01/88 ...
444 — 449 — 2.07 — 1.95
...
429 ... 433 — 2.08 — 1.94
03/15/88
■Raisedtemperature to 60*ConMarch 16, 1988. Droppedto 40'ConMarch 20, 1988 andbackto 25ConHay 10, 1988,

431 ... 421 — 2,090 ... 1,890
05/12/88 ...
433 421 — 2,090 — 1,930
05/24/88 5.40
427 10.85 421 10.80 2,070 6.11 1,920
06/05/88 — ... ...

418 — 399
... 397 — 2,060 — 1,930
06/26/88 417
10.68 390 10.61 2,070 — 1,910
06/28/88 416 — ... ...

07/01/88 415 — - 383
... 374 — 2,050 — 1,920
07/10/88 404 — ...
07/13/88 ... — — 2,070
— —
07/17/88 403 365 — 2,030
... 361 — 2,040 — 1,890
07/24/88 401 — 1,910
07/31/88 398 ... 355 — 2,070
10.90 348 10.71 2,040 6.11 1,900
08/06/88* 390

New Sample - ET-25C

conductivity pH

395 10.93
08/06/88
08/08/88 412
416 10.75
08/09/88 — -----------
08/10/88 415
— — —

08/12/88 415
414 10.60
08/13/88*

* Samples collected analysis


151

Table 21

Conductivity finKicromhos) and dH of


Ettrinaite Solubility

Exoerinentat 40*C

ET-40A ET-40B G-40 CaCl£'


Conductivity Conductivity dH Conductivity B2 Conductivity $
Date $

446 ___ 441 2060 ... 1940


03/22/88
436 ____ 429 2080 — 1900
03/29/88
04/11/88 448 ____ 438 2090 — 1900
453 ____
438 2120 — 1980
04/23/88
449 — 436 2200 — 1920
05/03/88
10.20 438 10.12 2110 — 1920 5.92
05/10/88* 450

* samples collected for analysis

A preliminary ettringite solubility sample at approxi

mately 22"c (room temperature) had been analyzed for component

ions in an effort to establish the anticipated ionic concen­

trations. This was necessary to allow the laboratory to

establish appropriate dilution factors and test procedures for

a mineral with an unknown solubility. The sample was prepared

by simply placing 0.3 grams of low grade ettringite (containing

more than 5 % calcite) into a polyethylene container with 1000

ml of doubly distilled, deionized water. The sample was purged

with nitrogen, sealed, shaken frequently, and allowed

equilibrate for approximately four weeks at 22-25°C. Results

of the analysis are presented in Table 22.


152

Table 22

Chemical Analysis of Preliminary Sample for


Fttrincrite Solubility

Description Ca A1 S04 C0 3 OH

Mg/liter 33.0 8.2 41.0 15.9 9.8

All solutions for analysis were collected under a


constant nitrogen purge both in the experiment flask
and the sample container. In order to keep the
liquid to solid ratio nearly constant in the
experiment flasks, the samples were kept very small.
Based on the preliminary analysis (Table 22) a
dilution of 1 0 : 1 was selected as the most appropriate
for the anticipated concentrations and the
requirement for a small sample size. _ .Fifty
milliliters (50 ± 0 . 1 ml) of boiled, doubly distilled,
deionized water were placed into each of five 1 2 5 mi
polyethylene sample bottles _using 3 pipette ®
water had been brought to 20 C(±l), the calibration
temperature for the pipette. Five miHiiite:rs (5.00
+ 0 1 ) of water were then removed from the sampl
container with a 5 milliliter pipette. Two. five ml
( 5 0 0 + .0 1 ) solution samples were collected from each
of the ettringite experiment flasks with the same
pipette . A d d i t i o n of the samples to the Previously
prepared sample bottles resulted m e esi ^
dilution. In one sample from each of the
ettringite flasks the water had previously Totsen
acidified by five drops of nitric acid in ord«-r
keep the aluminum in solution. It was assumed
the dilution would keep the other ion
precipitating, in spite of the temperature change
between the bath and the sample bottle; Th® '
in the bottle was at the, temperature of the
uncontrolled environment, typically 20-25 o.

gypsum sample was prepared for each


Only a single
procedure was identical to that of
temperature. The

ettringite, however, no acidification was necessary.

All solution analyses were performed at the Desert

Research Institute, Water Resources Center laboratory in Reno,


153

Nevada. The ettringite samples were analyzed for calcium,

aluminum, carbonate, and sulfate which were known to be the

only major ions present. The gypsum sample was analyzed for

calcium and sulfate and the calcium chloride solution was not

analyzed at all. Because some contamination would be expected

from the pH electrode, samples were also checked for sodium,

potassium and chloride.

After the solutions had approached equilibrium at 10 C (as

shown by constant conductivity) the temperature of the bath was

raised to 25 +0.1°C. The conductivities were monitored for 30

days. During this period the conductivity dropped steadily

giving the impression that ettringite was precipitating rather

than dissolving. In addition, comparison of analytical results

from the preliminary sample (22°C) and the experiment at 10°C

also suggested that ettringite had retrograde solubility. As

a consequence it was decided to raise the temperature of the

bath to 60°C, let the solutions equilibrate for several weeks

and then lower the temperature to 40 C. In this way the

ettringite, if it indeed was of retrograde solubility, should

precipitate slowly at 60'C and begin to redissolve at the 40 “C.

Again the kinetics of dissolution and nucleation/precipitation

favor this approach.

The solutions were maintained at 40 ‘ ±0. l ‘C for six weeks,

easonable stabilisation of the conductivity measurements was

ctually reached in only 3 weeks. After sampling, the tempera-


154

ture was lowered to 25° ±0.1°C and maintained for 8 weeks.

This time the conductivities for ettringite declined slowly and

then seemingly leveled off. However, in the 9 days between the

final two readings, the conductivity for both ettringite

samples dropped markedly. At that time the 25 C samples were

collected and submitted to the Desert Research Institute, Water

Resources laboratory. Because the samples at 25"c are critical

to the analysis of the ettringite dissociation constant and

because they had shown behavior inconsistent with equilibrium,

a third sample was prepared. The conductivity and pH

measurements for ET-25C were also shown in Table 20.

EVALUATION OR THE ANALYTICAT, DATA FROM THE SOLUBILITY

EXPERIMENTS

The results of the water sample analyses performed during

the ettringite solubility experiments are presented m

Table 23.

Table 23

Analytical Results of Ettringite Solubility Samples

jnalvsis in Mg/Liter

Ca Na K Cl SO, n C03
Samcle No.

4.1 2.5 7.0 70.0 13.0 63.0


ET-lOa 57.8
3.4 2.1 6.0 73.0 13.0 63.0
ET-lOb 58.9
1 2 86.0 17.0 36.4
ET-25a 64.9 1
1 2 87.0 16.0 20.6
ET-26b 62.3 1
1.5 1.7 2 70.0 13.0 57.0
ET-25C 61.6
1.2 5 97.0 18.0 88.5
ET-40a 76.8 3.4
1.8 6 98.0 18.0 85.5
ET-40b 74.5 3.4

2.3 7.0 1230 — <15


G-10 59.4 5.5
<1 <2 1450 — <15
G-25 60.9 <1
1.4 5.0 1480 — <15
G-40 60.5 12.1
155

These data were entered into the WATEQDR computer program

(Bohm and Jacobson, 1981). WATEQDR is an ion-paring chemical

model for determining the distribution of species in waters.

At the present time the program does not consider either

ettringite or thaumasite since their thermodynamic properties

were not known. The program, however, is extremely useful in

calculating the individual ion activities in complex aqueous

solutions. WATEQDR also provides the analytical and calculated

milligram-equivalents per kilogram of the input species. In

entering the data, the total analytical carbonate was split

into carbonate and bicarbonate, as appropriate, on the basis

of solution pH.

The milligram-equivalents per kilogram (epm) of the 10

solutions is summarized in Table 24.


156

Table 24

EPM Balance of Analytical Data

Milligram-Equivalents per Kilogram

Net 1 Net 1

Na K Cl SO, CO, HCO 3 Analytical Error Computed Error


A1 Ca

Sample

'1.458 ' 2 .1 0 0 - .818 9.8 '.646 11.3


ET-lOa 1.446 2.885 .178 .064 M 9 7
.148 .054 -.167 '1.520 ' 2 .1 0 0 -- .797 9.5 .665 1 1 .6
ET-lOb 1.446 2.940
'1.791 ‘1.213 - 2.138 25.9 2.156 32.7
ET-25a 1.891 3.239 .044 .026 '.056
'1.812 '.687 -- 2.403 32.0 1.063 18.0
ET-25b 1.779 3.109 .044 .026 '.056
'1.458 "1.900 -- 1.215 15.1 .244 4.5
ET-25C 1.446 3.074 .065 .044 '.056
1.615 15.5 .407 6 .1
ET-40a 2 .0 0 2 3.833 .148 .031 ".141 " 2 .0 2 0 "1.500 ".738
'2.041 '1.434 '.721 1.55 15.1 .465 7 .2
ET-40b 2 .0 0 2 3.719 .148 .046 ".169

-- '.123 4.017 7.2 3.948 1 0 .1


G-10 29.696 .240 .059 ’ .198 '25.656
__ - '.123 .205 0.3 .098 0 .2
G-25 30.453 .436 .256 '.565 '30.253
-- '.123 ".326 0.5 .427 1 .0
G-40 — 30.253 .527 .036 ".141 "30.879

Unfortunately, there is considerable error in some of the

analyses. The computed balance is off by as much as 33% for

ET-25a and 18% for ET-25b. Fortunately, sample ET-25C was much

better with an error of 4.5%. This is within the 5% error

normally assigned to the analysis of each of the ions (Jim

Heideker, 1988; personal communication).

The reason for the discrepancies in the epm balance is not

readily apparent; however, a comparison of the mole ratios

provides considerable insight. Table 25 compares the

stoichiometric mole ratios of ettringite to those calculated

from the analytical data.


157

Table 25

Stoichiometric and Measured Hole Ratios

Ettringite Samples

Stoichiometric Calculated Ratio:1


Mole
Katio ET-lOa ET-lOb ET-25a ET-25b ET-25C ET-40a ET-40b
Ions

3:1 2.99 3.03 2.57 2.62 3.19 2.87 2.79


Ca: Al
2:1 1.98 1.93 1.81 1.72 2.11 1.89 1.82
Ca:S04
1.5:1 1.51 1.58 1.42 1.53 1.51 1.51 1.53
S04:A1

Gypsum Sample

Stoichiometric Calculated Ratio:!


Hole
Ions 'Rati o G-10 G-25 (HO

Ca:S04 1:1 1:1.16 1:1.007 1:0.98

For samples ET-lOa and ET-lOb the calculated ratios are very

close to the stoichiometric ratios. The calcium to aluminum

ratio show virtually no calcium beyond that necessary for

ettringite. The calcium to sulfate ratio indicates that either

sulfate is slightly high or calcium is slightly low. The

sulfate to aluminum ratio is slightly high which,

combination with the calcium aluminum and calcium:sulfate

ratios, suggests that, in reality, the measured sulfate is

probably a little low. The data for samples ET-lOa and ET-lOb

support the x-ray analysis that indicates very little calcite

contamination in the ettringite. Most of the carbonate

probably came from either the boiled, purified water or from

air entering the flask in spite of the nitrogen purge. Since

the experiment lasted for over a year it is possible that some


158

carbon dioxide entered the polyethylene flasks by diffusion

through flask walls. In any case, the mole ratios and the

computed negative epm imbalance suggest that most of the error

for ET-lOa and ET-lOb is in the measurement of the carbonate.

The calcium to aluminum and calcium to sulfate mole ratios

for samples ET-25a and ET-25b are very low. Since the ratio

of sulfate to aluminum is fairly close, either both sulfate and

aluminum are low, in which case the sulfate to aluminum ratio

is fortuitous, or the measured calcium is too low. Even if the

calcium is raised to the ideal mole ratio, a negative balance

remains. It seems likely that errors in measurement of both

calcium and carbonate have compounded to yield the large, net

negative imbalance. Again, however, the analysis supports the

initial purity of the ettringite.

Sample ET-25C was mixed over a year after Samples ET 25a

and ET-25b using a split of the same synthetic ettringite.

Sample collection and analysis of ET-25c was performed only a

month later than for the original two solutions. In this case,

both the epm balance and the mole ratios are pretty good. The

calcium to aluminum and the calcium to sulfate ratios both show

an excess of calcium beyond that necessary for stoichiometric

ettringite. Once again, the sulfate to aluminum ratio is quite

close and unless this ratio is fortuitous and both sulfate and

aluminum are low, the calcium value would appear high. In

fact, if the calcium figure is correct, it would be more


159

double the amount that would be derived from complete

dissolution of the calcite indicated by the x-ray analysis.

The small negative epm imbalance again seems to involve both

calcium and carbonate. In this case a high reading in calcium

may have fortuitously helped to balance an ever higher

carbonate measurement.

The mole ratios for Samples ET-40a and ET-40b are similar

to those of ET-25a and ET-25b, however the errors are much

smaller. The epm imbalance is, again, negative suggesting the

problems are similar to those discussed for ET-25a and ET-25b.

The epm balance for gypsum samples G-25 and G-40 are

excellent. These good quality data are verified by the mole

ratios shown in Table 25. Sample G-10 has a positive imbalance

of about 10%. The calcium to sulfate ratio is also high,

however, it is not possible to deduce whether calcium is too

high or sulfate too low.

Summary

Although the analytical data for ettnngite are not

outstanding, it was judged suitable for the intended purpose.

The data for samples ET-25C, ET-40a, and ET-40b are actually

quite good. The two samples from 10 C are tolerable,

evaluation of the data is complicated by the presence of

under ideal conditions no carbonate would be


carbonate.
160

present. The x-ray pattern for the ettringite, however, showed

approximately 1% calcite. Furthermore, although water for the

solutions was treated to remove as much of the carbon dioxide

as practical, some carbon dioxide surely remained. Finally,

additional carbon dioxide must have been introduced into the

system during measurement of conductivity and collection of

water samples. This latter contamination was not indicated by

the calcium chloride solution which was apparently too weak to

be effective. Upon completion of the experiment the calcium

chloride solution was opened to the atmosphere and did not turn

cloudy.

The gypsum samples did not show very much carbonate. This

is most likely because carbon dioxide dissolves in aqueous

solutions as carbonic acid. Certainly much more acid can be

dissolved in a strongly basic solution than in one that is

slightly acid. If no carbon dioxide was present in the gypsum

solution, the pH would be 7 at 25”C. The pH values measured

in the solubility experiments for gypsum were around 6,

indicating the presence of some C02.

THE SOI.TTBTT.ITY OF GYPSUM

The solubility of gypsum was measured in this research

primarily as a check on the experimental techniques. All of

the procedures performed on the ettringite samples, incl ' g

the nitrogen purge, were performed on the gypsum samples. In


161

retrospect, a better choice might have been calcium hydroxide.

Calcium hydroxide, while still a simple compound like gypsum,

produces a high pH solution like ettringite. Since reagent

grade Ca(OH ) 2 is available, any carbonate in the solution would

have to have come from the air or water. This would have

provided a little better understanding of the extent of

carbonate contamination.

The individual ion activities for calcium and sulfate were

determined from WATEQDR, as previously described. An example

of the computer printout for 25°C is presented in Appendix 22.

The dissociation reaction for gypsum is:

CaS04*2H20 = Ca2+ + S042' + 2H20.

The dissociation (equilibrium) constant for the reaction is:

Kg = a[Ca2+] a [SO2'].

Using the above relationships and the WATEQDR output, the

equilibrium constant was calculated for temperatures of 10 C,

25°C, and 4 0 °C. The results are presented graphically as

Figure 28. The least squares method was employed to fit a

curve tothe points. Since only three data points were

available, the fit was perfect. The equation describing the

curve, and thus the change of the equilibrium constant with


TEMPERATURE IN

pK GYPSUM
Figure 28 The negative logarithm of ^he dissociation
constant of gypsum as a function of temperatu .
163

temperature (degrees centigrade) is: log K* = -4.6766 +

.005866T - 1.204 x 10 - 4 T2. For comparison the equation used

in WATEQDR is:

log Kt = -4.6535 + .004545T - 1.01 X 10-* T2.

The curves described by both of these equations are shown on

Figure 28 as are the error bars calculated from the data of

this study. In this case the bars are deceptively large due

to the small scale of the p K ^ axis. The error bars are based

on an assumed 5% analytical error in the water analysis. For

samples G-25 and G-40 this is certainly a valid and

conservative assumption. For sample G-10, the error probably

exceeds 5%.

At 25°C the log of the equilibrium constant for gypsum is

calculated to be -4.605 from the data of this study. For

comparison, the value calculated from the equation used in

WATEQDR is -4.603. Published values range from

(Krauskopf, 1979) to -4.61 (Drever, 1981). Above 25 C the

WATEQDR curve and the curve of this study are virtua y

identical. Both give a value of -4.634 at 40 C. Below 25

the curves diverge slightly. WATEQDR yields a value of -4.619

at 10 “C while the K 40 measured by this study is -4.630. This

discrepancy is actually quite small and most likely the result

of the analytical errors indicated by the 10 * epm imbalance

Sample G-10. In conclusion, the equilibrium constant for


gypsum at 25°C and the curve of the change in the equilibrium

constant with temperature, as determined by this study, are in

excellent agreement with published figures.

THE SOLtTBTT.TTY OF E T T R IN G IT E

Individual ion activities were, again, calculated with

WATEQDR. A typical computer printout (25 C) is included in

Appendix 23. The dissociation reaction for ettringite is:

Ca 6 [Al (O H )6]2(S04)3•2 4 H20 =

6 Ca2+ + 2A1 (O H )4” + 3S042' + 40H' + 20 H20.

The dissociation (equilibrium) constant for the reaction is.

Ket = a [ C a 2+] 6 a [ A l (O H )4' ] 2 a [ S 0 2' ] 3 a [ 0 H - ] 4

Using this equation and the ion activities obtained from

WATEQDR, the equilibrium constant was calculated for

temperatures of 10°C, 25°C, and 40°C. These values are

summarized in Table 26.


165

Table 26

Calculated Equilibrium Constants for Ettrinqite

Samples Temperature log K«t

ET-lOa 10°C -48.30


ET-lOb 10°C -48.21
ET-25a 25 °C -47.21
ET-25b 25 °C -47.78
ET-25C 25 °C -48.49
ET-40a 40 °c -47.22
ET-40b 40 °C -47.67

As shown by the equation for the dissociation constant,

equilibrium for this reaction is strongly affected by minor

changes in the individual ion activities. This is particularly

true of calcium and hydroxide which are raised to powers of six

and four, respectively. As a consequence, any errors, in

analytical data are amplified in calculating the equilibrium

constants. For this reason, the data from ET-25a and ET-25b

have not been included in the evaluation of the equilibrium

constant and its change with temperature. The remaining five

data points have been plotted as Figure 29. Using the

squares method on these points results in a best-fit curve

described by the equation:

log ket = 196.429 - 1.6699T + 0.002846T2

This curve is also shown on Figure 29.

The equation gives the log equilibrium constant for

ettringite as -49.49 at 25"C (298.16'K). Since there is only

one acceptable datam point at this temperature, the value


TEMPERATURE
166
167

calculated from the curve is coincident with that calculated

from the ion activities.

The analytical values of the various ions should be within

5 %. Given the epm imbalance of some of the data, errors are

probably a little worse than that, particularly for the

experiments at 10°C. The smallest epm imbalance occurs with

Sample ET-25C. Since 25°C (298.16 K) is the standard state

temperature of chemical thermodynamics, this is most fortunate.

An assumed analytical error of plus or minus 5% results in the

error bars for the calculated equilibrium constants shown on

Figure 29. The error in the calculated equilibrium constants

is probably within one-half a pK unit and, for sample ET-25c,

probably within one-quarter of a pK unit.

If the sole purpose of this research had been to evaluate

the ettringite solubility as precisely as possible, it would

have been necessary to run multiple samples (3 to 5 minimum)

at numerous temperatures between 5 'C and 50"C. Given the other

uncertainties in the geochemical/geohydrologic mechanism of

lim e -in d u c e d heave, th e v a lu e s o b ta in e d w ill be show n to b e

s u ita b le fo r th e ir in te n d e d p u rp o s e in C h a p te r 8.

THE THEnMOnyNAMTC.S OF ETTfTNGXTF. SOTIIBILITY.

The equation of the curve in Figure 29 fits the mean log

KC values to within ± 0 . 2 units and the value at 25'C exactly.


168

In natural logarithms the equation is equivalent to

lnKat = 452.2981 - 3.8450T + 0.006552T2.

From this expression and the relation AGr° = -RTlnk, we may

compute

AGr° = 7.6400T2 - 0.01302T3 - 898.7163T.

Thus at 298.16 °K (25° C) AGr° = 66.14 ±0.26 kcal/mole

Using the relationship

AHr0 = RT2(01nKet/0T) ; AHr° = 0 .02604T3- 7 .6400T2

At 298.16° K this expression yields A H r = 11.0+ 0.3

kcal/mole. The heat capacity of the reaction computed from

ACP° = (aHr°/9T) gives ACP° = 0.07812T2 - 15.28T or, at 298.16 K,

Acp° = 2.4 ± 0.1 kcal/mole. Finally, the expression

- ASr° = AGr°/ T yields Sr° = 15.28T - 0.03906T2 - 898.7163 and 19

cal/deg mole at 298.16 K.

Using the values of Gr°, Hr°, and Sr° obtained above for

the dissociation reaction, the specific values can be

calculated for formation of synthetic ettringite. For example,


, o=
AGr AGporoducts - ._oeactants, or AGF
AGr Ar»et = AG °ro,
a^p *a - AGr°.
duot From standard

tables of thermodynamic data (Krauskopf, 1979), this yields a


169

value of AGF°et = -3641 kcal/mole. Enthalpy and entropy for

ettringite were calculated in a similar manner. The

thermodynamic values obtained are summarized in Table 27.

Table 27

Summary of Thermodynamic Values for Ettringite a_t


25 °C f298.16 °IQ

Parameter Calculated Value

Log Ket -48.49 ± 0.3

AGf° -3641 ± 0 . 3 Kcal/mole

ah; -4128 + Kcal/mole

as; -647 ± 52 Cal/deg mole


170

CHAPTER 8

ANALYSIS AND EVALUATION OF EXPERIMENTAL AND FIELD DATA

t.&RORATORV SIMULATION OF LIME-INDUCED HEAVE

Expansive Pressure

The test to measure the maximum potential pressure from

growth of ettringite/thaumasite in a controlled environment met

with immediate success. The lime and sulfate apparently did

not react with the aluminum and steel in the apparatus m a

manner that inhibited crystal growth. The results of the

experiment have been presented as Figure 24. The maximum

expansive pressure was found to be 7535 pounds per sguare foot

for the soil tested. This pressure is certainly enough to

heave pavements and also foundations and floor slabs of lightly

loaded structures. The measured value is well within the range

commonly measured for expansive clays.

volume Change

The methodology for duplicating the heave mechanism in the

laboratory required the most time consuming and frustrating

experiments of the entire project. The greatest stumbling

block was the unexpected ability of the brass volume change

meter to react with the solutions and inhibit the formation of


171

ettringite. Once this problem was recognized, it was easier

to duplicate the swelling soil in the laboratory.

The experiments provided several important keys to the

understanding of lime-induced heave. It is hoped that the

comprehensive experiments in progress at the University of

California, Berkeley, (Appendix 11), will provide additional

details on some of the reactions.

The results of volume change experiment five were

presented graphically in Figure 25. The diagram shows that the

maximum volume change from kaolinite soils treated witn 6*

hydrated lime (4.5% equivalent quicklime) is, roughly, 50*.

In other words, a one-half-inch thick sample will expand to

approximately three-fourths-inch in height. In these

experiments the reactions were provided with an unlimited

source of sulfate by immersion in a solution either saturated

with calcium sulfate or containing 5500 ppm sulfate from sodium

sulfate. in addition, both silica and aluminum were present

in great excess due to the use of pure kaolinite clay, rather

than natural soil.

The initial volume change rates of samples VC-5A and VC-5B

were nearly identical, even though the sodium sulfate solution

provided more than three times the sulfate per liter than

gypsum solution. Sample VC-5C, which contained calcite,

expanded at a much slower rate than the other two.


172

The similar growth rates for VC-5A and VC-5B indicate that

sulfate is not the rate limiting step in the reaction process.

The fact that the sodium sulfate solution leveled off at a

slightly lower percent volume change than the calcium sulfate

solution, suggests that the reaction ran out of calcium before

it ran out of aluminum. In other words, the pH of the system

may still have been high enough to break down clay minerals.

The stoichiometry of ettringite, Ca6[Al(OH)6]2* (S04)3-26H20,

indicates the need for abundant calcium. The same holds true

for thaumasite, though it is unlikely to have occurred given

the temperature range of the experiment. It is concluded that

clay bearing soils containing either gypsum, or the much less

common sodium sulfate minerals thenardite and mirabilite, can

react with lime to produce heave. Of the two, gypsum probably

results in the higher heave magnitudes due to the additional

supply of calcium. Magnesium sulfate minerals, such as

epsomite, would probably behave in a manner similar to sodium

sulfate.

Sample VC-5C also provides interesting information. The

lower rate of growth is important, especially since expansion

continued beyond the time allotted for the experiment.

Presumably, the ultimate volume change would have approached

that of the other two samples had the test been run for

years. The slower rate may be the result of complex factors

not yet understood. A simplistic explanation, however, is that


173

the reduced rate is the result of the common ion effect. The

presence of calcium carbonate should reduce the solubility of

gypsum. The equilibrium reaction for the dissolution of gypsum

is:

CaS04*2H20 = Ca2+ + SO/" + 2H20

Any increase in the chemical potential of calcium, caused

by dissolution of calcite, should drive the reaction back to

the left. Since calcite was present in the native soil at both

Stewart Avenue and Owens Avenue, the effect on the reaction

rate is of considerable interest. Had calcite not been

present, the heave might have occurred much more quickly. In

fact, the expansion potential might even have been detected in

the laboratory during R-value testing and mix design. The

lime-induced heave reported by the highway departments of Utah

and Texas occurred prior to paving. Although this is entirely

speculation, the rapid response suggests that minimal carbonate

was present in the system.

The sixth phase of the volume change testing used native

soils from areas of major distress along Stewart Avenue. The

results of these experiments have been presented graphically

in Figure 26. In this case, the volume change of the two

samples reached approximately 26% for VC-6A and 21% for VC-6B.

The experiment was terminated after 6 months, however, the

growth curves showed no signs of leveling off. Since sulfate


174

and a high pH were being provided from outside the soil, the

expansion would be limited only by the amount of available clay

(aluminum). Although both soil samples contained abundant

soluble sulfate, neither specimen reacted significantly when

the apparatus was submerged in purified water. After several

weeks the water was replaced with a solution saturated with

gypsum and calcium hydroxide. The response was sluggish but

the two samples soon began to expand at a nearly uniform rate

of approximately 1% per week. The rates were considerably

slower than for the pure kaolinite "soils" but exactly in line

with the 100% volume increase shown on Stewart Avenue over a

period of two years. The slower rate, relative to kaolinite,

suggests that the dissolution of clay is the rate controlling

step in the reactions, even at the maximum pH (12.3) for lime

saturated water. The presence of carbonate, however, also

slows the reaction. The fact that expansion did not occur

until the samples were immersed in a sulfate solution is

important. It suggests that the heave might not have occurred

or would have been very slow if only pure water had been added

to the system at Stewart Avenue. Perhaps the calcium would

have become tied up by normal pozzolanic reactions before

sufficient sulfate could be generated to induce heave.

Subsequent, slower, reactions would have attacked the subbase

through the same mechanism by which sulfate deteriorates

hardened concrete. In hardened concrete ettringite growth

produces crumbling and strength loss but only a very small

volume increase.
175

The movement of ions through the porewater system of soils and

sediments is very slow (Berner, 1971). It may be that a great

deal of time is required for the minerals to ionize in the pure

porewater and then migrate to favorable nucleation sites. It

has been shown by experiment that the nucleation sites are

probably on or around the lime particles where the clay

minerals are releasing aluminum and silicon. Much more rapid

reaction seems to occur when the water entering the soil/lime

system is already rich in sulfate. Under this condition, the

necessity of mobilizing the gypsum ions within the soil is

avoided. Water entering most of the damaged areas of Stewart

Avenue and Owens Avenue appears to have reached the nucleation

sites through capillarity, presumably after moving through a

significant thickness of soil. The sulfate leaching experiment

performed during this research showed that pure water flowing

through a 24 inch (61 cm) long column of typical Stewart Avenue

soil carried over 5500 ppm sulfate. Thus, the use of a sulfate

solution in the Phase 6 experiments is not only the correct

model, but may also be necessary to stimulate reaction.

The addition of lime to the external solution probably was not

necessary for the reaction since little calcium hydroxide would

be present in the capillary water on Stewart Avenue. The lime

was added in the experiment in hopes of speeding up the

reactions so that the expansion could be measured m a


176

reasonable time. Even so, the experiment required over six

months and the benefit of the external lime is not really

known.

FTTRINGTTF, /THAUMASITE GROWTH IN OTHER SOILS

Although lime-induced heave has not been a common problem,

this experiment demonstrated the potential for ettringite/

thaumasite growth in soils other than those of Stewart and

Owens Avenues. Of the fourteen soils tested, eleven were from

areas well outside the Las Vegas valley. As shown by Table 8,

lime-treated soils from the Bonneville Salt Flat, Utah, and

Carson Sink, Pyramid Lake, White Lake, and northwest Reno,

Nevada, showed significant ettringite/thaumasite. Three other

samples, Black Rock Desert, Nevada, Snake River Floodplain,

Idaho, and Yolo Causeway, California, have a somewhat weaker

potential for heave. Additional cases of distress, as

described in Chapter 4, show that the problem is world-wide and

not even be limited to arid environments or soils containing

sulfate minerals.

The mechanical analysis data (Table 7) show that the seven

reactive samples all contained a relatively high percentage

particles in the clay and colloid size range. Those that

not appear to have reacted, generally have relatively low

clay/colloid levels. No chemical analyses were performed on

any of the samples, other than those from Stewart Avenue.


177

Previous research (Kollmann, 1978; Kollmann, et. al., 1981) has

demonstrated that ettringite will form when clay minerals are

mixed in a slurry with lime in the presence of sulfate. It

follows that soils that do not form significant ettringite

either lack sulfate, clay, or sufficient water. Since the

samples of this experiment had abundant water, those that did

not react were deficient either in sulfate or clay. Based on

the hydrometer analyses, all of the samples likely had enough

clay. Therefore, those that did not react probably lacked in

sulfate. Only four of the soils showed sulfate bearing

minerals in the x-ray diffraction analysis of the elutriation

concentrate (Table 8).

The soils of this experiment generally required that the

low density minerals be concentrated in order to positively

identify the ettringite/thaumasite. The amount of ettringite/

thaumasite present was, therefore, quite small. The evaluation

of the Stewart/Owens distress revealed that sulfate and other

ions migrated to the nucleation sites. As a consequence, lime-

induced heave is dependent upon the initial soil chemistry only

to the extent that ettringite/thaumasite nucleation must be

thermodynamically and kinetically favored. The magnitude of

the distress, where it even occurs, is more dependent upon the

ability of the geohydrologic/geochemical system to supply

component ions to the reactions. These factors will be

considered in detail in later sections of this chapter.


178

SYNTHESIS OF THAUMASITE

Synthesis of pure thaumasite has been performed by a

number of workers, as discussed in Chapters 5 and 6.

Unfortunately no one is known to have isolated this pure

thaumasite from the other components involved in the reactions.

Several methods to isolate pure thaumasite were attempted

during this research. These were described in detail in

Chapter 6 but included rinsing of the mineral aggregate, use

of heavy liquids, elutriation, preferential nucleation, and

kinetic exchange. These experiments consumed a considerable

amount of time and, unfortunately, were unsuccessful. Although

several methods were found to increase the concentration of

synthetic thaumasite, none could isolate the mineral to tne

extent that it would be suitable for solubility measurements.

As a consequence, the solubility of thaumasite was not

determined.

SYNTHESIS OF ETTRINGITE

As described in Chapter 6, the synthesis of reasonably

pure ettringite was successful. Gypsum can effectively be kept

from precipitating with the ettringite by mixing the solutions

at a slow enough rate to prevent localized gypsum super­

saturation. in contrast, however, calcite presents a much more

challenging task. If all of the carbon dioxide could be

removed from the purified water and all aspects of the


179

synthesis performed in a C02 free environment, it would be

possible to synthesize pure ettringite. Unfortunately, these

conditions are very difficult to achieve.

M A SS—VOTiIIME R E L A T IO N S H IP S

Profiles across severely damaged areas of Stewart Avenue

showed maximum heaves exceeding 12 inches (30.5 cm). This

correlates well with observations in the trenches where the

lime-treated subbase has expanded from an approximate initial

thickness of 12 inches (30.5 cm) to as much as 27 inches

(68.6 cm). Because ettringite/thaumasite is hydrated with an

extraordinary number of attached water molecules, the mineral

takes up a much higher volume than its constituent ions. The

volume change experiments, even using pure kaolinite and

unlimited sulfate, suggest that the maximum heave that could

be expected from 4.5% quicklime would be on the order of 50«.

The trilinear (Figures 18, 19, and 22) and stiff (Figure 20)

diagrams, however, indicate that there has been considerable

influx of ions, beyond those present in the native soil. These

additional ions could account for the observed discrepancy

between the volume change measured in the field and that

measured in the laboratory.

As a check on the mass-volume relationships in the field,

the approximate volume change expected from a given soil

chemistry can be calculated. The mass-volume relationships in


180

the field can be estimated from either the change in density

or from the change in specific gravity. The latter is thought

to be the most reliable since the initial (end of construction)

density would have to be assumed from construction records.

Although this density must still be used in calculating the

total sulfate available per cubic foot of soil, that particular

calculation is not overly sensitive to density. Test pit S-8

was selected for the evaluation because all of the needed data

had been specifically measured in that location and S-8 is

adjacent to a control test pit, S-9. The following assumptions

were made:

1. Thaumasite was the only secondary mineral grown m

the system.

2. Infinite silica was available from dissolution of

clay and infinite calcium and carbonate from

dissolution of calcite. That is, sulfate was the

limiting component.

3. L im e -tre a te d s u b b a s e was o r i g i n a l l y 12 i n c h e s thick,

as per d e s ig n .

4. Sample S-9B represents original subgrade soil

chemistry for S-8.


181

5. All of the sulfate available in the native soil

reacted.

The initial data needed for the calculations were obtained

from a combination of construction field reports, field

measurements taken during the distress analysis, laboratory

testing of the soil properties and chemistry, and the

Encyclopedia of Minerals. These data are summarized in

Table 28.

Table 28

nata for Mass-Vo]ume Calculations

Maximum dry density of native soil 106.9 pcf*


Specific gravity of native soil 2.73 gm/cc
Specific gravity of damaged^subbase 2.37 gm/cc
Specific gravity of ettringite 1.77 gm/cc
Formula weight of ettringite 1254.0 gm/mole
Molar volume of ettringite 708.5 cc/mole
Specific gravity of thaumasite 1.91 gm/cc
Formula weight of thaumasite 1220.0 gm/mole
Molar volume of thaumasite 639.0 cc/mole
Initial dry density of
lime-treated subbase 96.2 pcf*
Final dry density of
damaged lime-treated subbase 59.3 pcf
Final moisture of
damaged lime—treated subbase 59.0%
Initial volume of 1 ft2 of
lime-treated subbase 1.0 ft3
Final volume of 1 ft2 of
lime-treated subbase 2.1 ft3
42.000 mg per kg
Available sulfate (S-9B)
of soil
16.000 mg per kg
Available calcium (S-9B)
of soil

construction (Converse Davis and


* Measured at time of
Associates, 1975).
182

Chemical calculations

a. Assume all 42,000 mg-sulfate reacted to form


thaumasite. ----------
kg-soil

b. One ft3 soil at 96.3 pcf = 4 3 . 8 kg/ft3

c. (43.8 kg/ft3) x (42,000 mg/kg) = 1840 grams of


available sulfate per cubic foot of soil

d. 1840 gm sulfate = 19.2 moles of sulfate per cubic


---------------— foot of soil
96 gm/mole

e. From stoichiometry of thaumasite, 2 moles sulfate


yields 1 mole of thaumasite:

19.2 moles sulfate = 9.6 moles of thaumasite

(9.6 moles) x (1220 gm/mole) = 2 5 . 8 lbs of thaumasite

454 gm/lb

Mechanical Calculations

a. To achieve an average specific gravity of 2.37


(damaged subbase) from weighted average of 2.73 and
l.gi reguires 56% soil and 44% thaumasite.

(.56) X (2.73) + (.44) X (1.91) - 2.37

b. (59.3 pcf) x (.44) = 2 6 . 1 pcf thaumasite

c. (26.1 pcf) x (454 gm/lb)


(1.91 gm/cc) x (28317~cc/ft3) = 0.219 ft3 thaumasite.

d. (.56) x (59.3 lbs) x (454 gm/lb)

(2.73) X (28317 cc/ft3) = 0.194 ft3of soil

1- 59.3 lb
(2.37 gm/cc) x 62.4 pcf) = 0.599 - volume of
voids
So for each cubic foot of damaged subbase there is
approximately:

0.19 ft3 soil


0.22 ft3 thaumasite
0.59 ft3 voids.
jssm

183

g. There are 2.1 ft3 (per square foot) of damaged


subbase that grew from 1 ft3 of treated soil. Thus
the volume of thaumasite required is (2.1) x (0.22
ft3) = 0.44 ft3.

h. Mass of required thaumasite is


(0.44 ft3) X (1.91) X (62.4 pcf) = 54.8 lbs.

i. From step lg, the mass of thaumasite that can be


grown from 42,000 mg-sulfate/kg soil is 25.8 pounds.

j. On the basis of this analysis, the closed system is


deficient by 54.8 lbs - 25.8 lbs = 29.0 lbs of
thaumasite.

k. Thus the available sulfate is only sufficient for


approximately:

25.8 lbs x 100 = 47%

54.8 lbs
of the required thaumasite.

In reality, ettringite very likely grew before thaumasite.

This is well supported by laboratory research of previous

workers (Kollmann, 1978; Kollmann et a l , 1977; Kollmann and

Strubel, 1981). Ettringite requires more sulfate than

thaumasite indicating an even larger deficiency. However,

since the specific gravity of ettringite is lower than that of

thaumasite, the effect would not be as large as might be

anticipated from stoiciometry. Clearly, however, in both cases

the available, closed system, sulfate is less than one half of

that necessary to satisfy the mass-volume requirements.

The laboratory tests using the kaolinite "soil" provided

unlimited sulfate and aluminum to the reactions. In VC-5A the

reaction appears to have been limited by calcium since the


184

expansion ceased before that of VC-5B (Figure 25). Sample

VC-5B was supplied unlimited calcium from the external calcium

sulfate solution. For comparison, the potential volume change

was calculated using the data of test pits S-8 and S-9 but with

calcium as the limiting component. The procedure was exactly

like that used for sulfate. However, in this case there are

two sources of calcium, neither of which is entirely

predictable. The first source is the native minerals, calcite

and gypsum found in the untreated soil. Calcite was probably

not completely dissolved by the 50:1 extraction ratio such that

there is somewhat more available in situ calcium than the

analysis shows. The second calcium source involves soil

treatment with 4.5 weight percent quicklime. Since the

compacted unit weight of the soil varies along the alignment,

the mass of lime added per cubic foot of soil varies. In

addition, the entire construction process of adding and mixing

lime is difficult to control. In spite of these limitations

an estimate is possible:

1. Assume initial unit weight of 96.2 pcf as per


construction records.

2. (96.2 lbs) X (1.045) - 96.2 lbs = 4.335 lbs (1967 gm)


of CaO.
(1967 gm) x 40 (atomic weight of Ca)
56 (formula weight of Ca(OH)2) = 1405
gms of calcium)

4. 96.3 lbs = 43.7 kg treated soil


185

5. (43.7 kg) x (16000 mg/kg)

1000 mg/kg = 699 grams Ca.

6. 699 gm + 1405 gm = 2104 gm available calcium.

7. 2104 gm

40 gm/mole = 52.6 moles Ca

8. from stoiciometry, 6 moles Ca yields 1 mole


thaumasite:

52.6 moles

6 = 8 . 8 moles thaumasite.

9. (8.8 moles) x (1220 gm/mole)


454 gm/lb = 2 3 . 6 lbs thaumasite.

10. From step 2h of previous calculation, 54.8 lbs of


thaumasite are required. The available calcium is
thus only sufficient for approximately

(23.6 lbs) x 100

54.8 lbs = 43% of the required


thaumasite.

Since not all of the calcite dissolved during the


extraction, the actual figure may be closer to 50«.
In most cases the 4.5% quicklime probably supplies
over half the calcium.

The calculated maximum volume change is in good agreement

with the 50% volume change measured in the laboratory for the

kaolinite ''soil.*' The experiments appeared to be first limited

by lack of calcium when sulfate was kept in supply by sodium

sulfate. When both sulfate and calcium are unlimited, the

reaction should continue until aluminum is no longer available.

This could occur either when the pH drops below 10.5 or the

available clay has been exhausted. On Stewart Avenue, east of

Lamb Boulevard, the latter was probably the case. The pH of


186

lime-treated subbase, either damaged or undamaged, was nearly

always above 10.5 on both Stewart and Owens Avenues.

The remaining ions necessary for the measured volume

change would have to be added to the system through the pore

water. In fact, the calculated maximum volume changes probably

greatly exceed that which could have occurred in the field in

a closed system. In most areas, some of which exhibited 100%

expansion, the sulfate content of the native soil was far less

than 42,000 mg/kg. Additionally, x-ray diffraction showed that

much of the sulfate did not react to form ettringite/thaumasite

as it was tied up in gypsum. It is concluded, therefore, that

an influx of sulfate, and presumably, calcium and carbonate was

necessary to generate the observed heave magnitudes. This

conclusion is well supported by the trilinear and stiff

diagrams, as previously discussed.

One very interesting point brought out by the mass volume

analyses is just how little of the heave (volume increase) is

the direct result of the volume of the ettringite/thaumasite

mineral mass. In the reactions, the total volume of the soil

solids per cubic foot decreased from, roughly 56% to 44%.

Presumably, this reflects the change in original soil volume

as soil ions were consumed by ettringite/thaumasite. The

actual volume of thaumasite is only about 0.22 ft out of each

cubic foot of damaged subbase. Consequently, in the expansion

from 1 ft3 to 2.1 ft3, the thaumasite accounts for,


187

approximately, (2.1 x .22) 0.46 ft3 of the 1.1 cubic foot

increase in volume. Correcting for the loss of soil volume due

to reactive consumption, leaves a net volume increase of 24%

due directly to thaumasite. The remaining volume change, which

is 74% of the 1.1 cubic feet of expansion, is the result of

increased voids (loss of density) associated with the mechanics

of dissolving and growing minerals within a highly compacted

soil. The increased voids, decreased original soil volume and

volume of ettringite/thaumasite are depicted graphically in

Figure 30.

At Stewart/Owens, near-end member member thaumasite was

the final reaction product. Only silica* and, perhaps, a small

quantity of aluminum and hydroxide, were consumed from the clay

dissolution. Although calcite and gypsum were, doubtless,

consumed, the x-ray patterns often show calcite and gypsum at

levels that indicate secondary mineralization. Large crystal

aggregates of gypsum and calcite were observed, at Stewart

Avenue, along the upper and lower contacts of the lime treated

subbase. Thus, the secondary growth was verified in the field.

The reactions that account for secondary calcite and gypsum

will be discussed with the geochemical reaction model m a

later section.
188

HEAVED CONDITION

.. 24" UNFILLED VOIDS «


6%

VOIDS/WATER 53%

INITIAL CONDITION
-12"
15% UNFILLED VOIDS

28% VOIDS/WATER ETTRINGITE/


20%
THAUMASITE

57% SOIL SOLIDS SOIL SOLIDS 21%

-0"

Figure 30 Changes in soil volume due to growth of


ettringite/thaumasite.
189

tfigs-FT.UX CALCULATIONS

The evidence seems to require an influx of outside ions

into the lime soil system. Mass-flux calculations were

performed in an effort to see if the necessary ions could reach

the nucleation sites in a reasonable amount of time. At

Stewart Avenue and Owens Avenue, the maximum vertical

expansions were reached within two years of construction. In

addition, Las Vegas has an arid climate with rainfall of less

than five inches per year.

The data used in the mass-flux analysis are summarized in

Table 29.

Table 29

Initial permeability of lime-treated 0.2 gpd/ft2


Final^permeability of damaged lime-treated 3.6 gpd/ft2
subbase1 , 5700 ppm
Sulfate concentration in porewater 3.3
Average hydraulic gradient

"Measured in laboratory _ street middle of LTSB


2Assumed one foot of storm water in flow distance of
1.5 feet below top of pavement and a v e ^ r a g ^ e ^ Qf LTSB; 0 .5
0.75 feet. Flow distance is °ne If ^ LTS B _ The average
foot for undamaged and / ? / ° ° l L t i n g the hydraulic gradient,
of these values was used in estimating

The following calculations were performed:

KI
q = KIA o r f o r A - 1 ft / Q
_L . From Darcy's Law:
r i. w in .
ahig itv for undamaged and damaged
2. Use average permeability
condition.
190

q = ( i . g gpm/ft2) x (3.3) = 6.3 gpd/ft2 = 23.8 liters per


day/ft2
Using the first mass-volume calculations, system is short
29 lbs of thaumasite or:
29 X 454
= 1 0 . 8 moles thaumasite or (2) x (10.8)
-- 21.6 moles sulfate.
1220 gm/mole
(21.6 moles) x (96 gm/mole) x (1000 mg/gm) - 2,073,600 mg
sulfate needed
From sulfate leaching experiment, soil porewater contains
5700 mg/1 sulfate.

2,073,600 mg = 364 liters

5700 mg/liter

8, 364 liters = 15 days

23.8 liters/day

on the basis of the above analysis, it is estimated that

15 days of flooded streets would be required to supply a

sufficient quantity of ions to the nucleation sites. This

appears to be a little high for the Las Vegas climate.

However, much of the water was probably supplied by capillarity


, t +- no not clear exactly how this
from the utility trenches. It
. . than it would likoly
would affect the rate estimate
increase the amount of saturated time necessary to supply the

ions. It is also not Known just how high the water level could

rise in utility trench bacKfill during thunderstorms. Ponding

of water in utility trenches would supply capillary water

after the street flooding had disappeared. Also, water would

pond in trenches from storms far too small to cause street


191

flooding. Overall, the calculations suggest that the required

ions could probably be supplied within the observed time frame.

The common growth of large crystals along the upper and

lower contacts of the lime-treated subbase is physical evidence

of water having entered the system at the horizontal boundaries

(refer to Figure 15). The larger crystal aggregates indicate

that more ions reached the boundaries than were carried by

capillarity into the center of the subbase. This suggests that

water reached the boundaries more often than it reached the

middle of the layer and/or the ion supply was increasingly

depleted as water migrated inward from the subbase boundaries.

CTOCHEMICAT, REACTION MODEL

The geochemical reactions involved in the growth of

ettringite, and its subsequent transition to thaumasite, are

doubtless quite complex in lime-treated soils. The initial

Stewart/Owens soil chemistry, primary and secondary mineralogy,

experiments performed during this research, and the published

solid solution series between ettringite and thaumasite suggest

a series of stepped equilibrium reactions as a general model

of lime-induced heave (Hunter, 1988). These reactions are as

follows:

1. CaO + H20 = Ca (O H )2
(hydration of quicklime)

2. Ca(OH) 2 = Ca2+ + 2 (OH)' .


(ionization of calcium hydroxide, pH rises to 12.3)
192

3. Al4Si4O10(OH) 8 + 4 (O H )' + 10H20 = 4Al(OH)4"+ 4H4Si04


(dissolution of clay mineral, at pH > 10.5)

4. 2H4Si04 = H2Si042~ + 2H+


(dissociation of silicic acid at pH > 11.7)

5. 5Ca2+ + 6H22~Si04 = Ca5(Si601BH2) •4H20 + 60H"


(formation of calcium-silicate-hydrate gel)

6. MxS04-nH20 = XMY+ + S042" + nH20 _


(dissolution of sulfate minerals; x = l , y=2 or x-2, y-1)

7. 6Ca2+ + 2A1 (O H )4" + 4OH" + 3(S04)2" + 26H20 =


Ca6[Al(OH)6]2* (S04)3•26H20
(formation of ettringite)

8. C02 + H20 = H2C03


(formation of carbonic acid)

9. CaC03 + H2C03 = Ca2+ + 2H+ + 2C032"


(dissolution of calcite in carbonic acid)

10. Ca r A l f O H ) , ] , * (S04)3-26 H20 + 2H2Si042" + 2C032 + 4H


S;[sI(OH)I];-(SO:)l-(CO,)V24H,0 + 2Al(OH)'t + SO/' + 40R- +
2H20 , ,
(isostructural substitution as ettringite changes to
thaumasite)

11. Ca2+ + S042" + 2H20 = CaS04•2H20


(formation of secondary gypsum)

12. Ca2+ + C032" = CaC03


(formation of secondary calcite)

Reactions 1 through 5 are "normal reactions resulting m

siliceous cementation of lime-treated soils. Reaction 3 is

shown for kaolinite since it was used successfully m the

laboratory. In fact, any clay mineral could be used in the

model since the sole function of the clay is to provide a

source of silicon and aluminum. In the presence of excessive

sulfate, the rate of the forward reaction in equilibrium pair

number 5 approaches zero. Sulfate from any evaporite (Reaction

6) reacts with aluminum generated by dissolution of any clay


193

mineral to form ettringite (Reaction 7). Based on the

Tennessee Department of Transportation study, sources of

sulfate other than evaporate minerals are possible.

Once ettringite has nucleated, it continues to grow as the

nearly pure end member until the temperature of the system

drops below approximately 15°C. Below 15 C ettringite is

transformed, through a series of intermediate solid solutions,

to thaumasite. The change occurs by isostructural substitution

of silica for aluminum and carbonate for sulfate. Based on the

reaction, there may also be some pH control on thaumasite

formation. Regardless of temperature, the transformation may

not begin until ettringite has consumed sufficient OH to drop

the pH below some threshold level. Secondary gypsum (Reaction

11) likely forms from calcium released by dissolution of

calcite (Reaction 9) and sulfate released from dissolution of

ettringite. Reactions 7 and 10 stop as pH drops below 10.5,

thereby cutting off the supply of aluminum and silicon

(Reaction 3). In the presence of excess calcium, Reaction 12

occurs at that time as carbonate ions are no longer being

consumed by thaumasite.

The temperature control of the ettringite to thaumasite

transition must be, at least, partially a function of the

retrograde solubility of caloite. However, the strong pH

dependence of both carbonate and silica solubility adds

additional complexity to the problem. At high pH, silica is


194

quite soluble. Calcite solubility, however, decreases rapidly

at higher pH. Thaumasite growth, therefore, must be, partially

controlled by a balance between two nearly mutually exclusive

reactions. A diagrammatic representation of the lime-induced

heave mechanism is presented as Figure 31.

One unanswered question is whether or not thaumasite is

even necessary to induce heave. In the laboratory, ettringite

always forms first at any temperature when both aluminum and

silicon are present. The transition to thaumasite may well

have occurred after the heave. At Stewart/Owens, the

transition was complete; no identifiable ettringite was found.

In Kansas (Mitchell, 1986), ettringite was blamed. This may

represent a case where the damage occurred and was studied

prior to the transition or it may be another indication of the

difficulty in distinguishing ettringite from thaumasite by x-

ray diffraction. The Tennessee Department of Transportation

case was not analyzed to determine if the specific distress-

related mineral was ettringite or thaumasite. The position of

the primary x-ray diffraction peak is strongly suggestive of

ettringite but is not conclusive. Since no carbonate minerals

were found in the untreated soils, it is suspected that the

solid solution series would favor ettringite. However, the

extent that carbon dioxide (either dissolved in the porewater

or biogenic) participated on the reactions is not known. Where

C02 can be constantly supplied, laboratory research has shown

that thaumasite can form without the need for carbonate

minerals.
195
196
W ATER FROM RESIDENTIAL
L A W N IRRIGATION, CAR
WASH, POOL DRAINING ETC.

ETTRINGITE + CALCITE+ SILICA“^-THAUMASITE + GYPSUM WHEN TEMP. DROPS BELOW 15 C

CAPILLARY MOVEMENT

NATIVE CLAY PONDED WATER REMAINS


LONG AFTER STORM

Figure 31c Diagram m atic re p re se n ta tio n of the mechanism


of lime induced heave.
198

CHAPTER 9

CONCLUSIONS AND APPLICATIONS

Up to this point the research has been primarily concerned

with developing a detailed understanding of lime-clay-sulfate

reactions and the resultant volume increase. The underlying

goal of the project, however, is to put this knowledge to

practical use in predicting the conditions under which lime-

induced heave is likely to occur. From a thermodynamic

standpoint this is not difficult. The free energy of

ettringite is such that it will be the stable phase whenever

lime (CaO or Ca(OH),), clay, sulfate, and water are present,

in reality, some ettringite can be tolerated in the soil

structure without producing disruptive swelling. On the basis

of this research sulfate levels as high as 5000 mg sulfate per

1000 grams of soil should be acceptable for flexible pavements.

For rigid structures such as concrete slabs or foundations a

more appropriate limit would be in the range of

mg/kg. The real problem is controlling drainage and

capillarity sufficiently to maintain an essentially

system and prevent additional sulfate from reaching

ettringite/thaumasite nucleation centers. This is not

trivial chore, as will be discussed, and may void any economic

advantages of using lime to stabilize sulfate bearing soils.


199

L A B I L I T Y DIAGRAMS

The stability fields, relative to ettringite, of many

minerals and compounds can be derived from the equilibrium and

thermodynamic values obtained during this research. The goal,

however, was to develop diagrams that would be helpful in

evaluating the potential for ettringite/thaumasite growth in

lime-treated soils. For this purpose three ettringite

stability diagrams were selected, as discussed below.

Portland! te fCafOH),!

The stability relations between portlandite (hydrated

lime) and ettringite can be evaluated from the equilibrium

reaction:

Ca6[A1 (OH) 6]2* (S0 4)3*26H20 + 20H' = 3Ca(OH)2 + 3Ca2+ +

3 (S04)2' + 2A1(0H); + 26H20 .

From the expression Gr’ = products - reactants, the

value of g; for ettringite determined in this study, and

published values for the other species (KrausKopf, 1979) G° for

this reaction is 44.49 kcal/mole. The relationship between the

free energy and the equilibrium constant is log K„ = ___


-1.364
From the equilibrium reaction, log Keq

log (a[Ca2+]2 a[Al(OH)4']2 a[S042']3) = -32.64

log a [OH']2
200

J
201

R e a rra n g in g term s re s u lts in

log a[Ca2+]3 a[S02"]3 = -32.64 + 2 log a[0H"] - 2 log a[Al(0H);]

from which the stability diagram of Figure 32 has been

calculated. The stability fields of the two compounds have

been determined with aluminum activities ranging from 10"1 to

10"6. The pOH of lime-saturated soil and the minimum pOH for

dissolution of clay have also been plotted. The diagram is

useful mostly in that it shows the expected increase in the

stability field of ettringite with increase in the activity of

aluminum. An increase in the activity of calcium and/or

sulfate drives the reaction into the stability field of

ettringite. Conversely, an increase in hydroxide activity

drives the reaction toward portlandite. The pOH boundaries of

lime-saturated soil and dissolution of clay greatly limit the

actual stability fields.

Cypsum (CaSCj •2HzOl

The stability relationship between gypsum and ettringite

can be represented by the equilibrium reaction.

3Ca2+ + 2A1 (OH) 4 + 4OH" +


Ca6[Al (O H )6]2' (S04)3*2 6H20

3CaS04•2H20 + 20H20

Using the same procedures as for portlandite, Gr 48.48

log Ket = -35.54. The expression for the equilibrium constant


202

is log Keq = log (a[Ca2+]3 a[Al(0H);]2 a[0H‘]4) = -35.54 which can

be rearranged to _/

log (a[Ca2+]3 a(0H')4) = -35.54 - 2 log a[Al(0H);]

and used to generate the stability diagram shown in Figure 33.

This diagram, while interesting, is only marginally useful

since the activity of aluminum is very difficult to measure.

In addition, the diagram does not consider sulfate which, of

course, is critical to lime-induced heave. Nonetheless, the

figure shows that a very slight increase in the activity of

calcium, a rise in the pH, or an increase in the activity of

aluminum drives the system toward the ettringite stability

field. All three of these changes occur when a clay soil is

mixed with lime.

Bvn-T-inaite a n d t h e S a t u r a t i o n Index

The potential for ettringite precipitation can probably

best be determined from the saturation index. The saturation

index is defined as:

log (ion activity product)


p ( C a r (OH)
203
204

In this case SI -

lo g ( a [ C a ‘* ] ‘ a [ A l ( 0 H ) ; ] ‘ a [ S o ; - ] ‘ a [ 0 H - l ‘ a [ H ;0 ] ” )

-48.5

when the log of the ion activity product is greater than the

log of the equilibrium constant (SI>0) the system is

oversaturated with respect to ettringite and precipitation

could occur. Conversely, when the saturation index is

negative, the system is undersaturated with respect to

ettringite and no precipitation should occur. In reality, even

though the thermodynamics clearly favor this relationship, the

reaction kinetics may not. Often a high activation energy

precludes nucleation, even in solutions grossly oversaturated

with a specific mineral. In the case of ettringite, nucleation

and precipitation probably do occur when the solution is

saturated or only slightly oversaturated. This is evidenced

by the ease with which ettringite can be synthesized from a

variety of components as well as its formation when sulfate

contacts hardened concrete.

Figure 34 is a graphical presentation of the saturation

index relationship for ettringite and its component species.

The limits of pOH for lime-saturated soil and that required for

dissolution of clay minerals are shown, as are the effects of


. • x- “ The calculated locations
two different activities for A1(0H)4 . The cai
for untreated and lime treated soils from both Stewart and

Owens Avenues have also been plotted. These points are based

on the chemical analyses of these soil samples as presented in


p(Car(S04)
205
206

Appendix 5. The individual ion activities were calculated

using WATEQDR and are based on a fairly complete soluble ion

chemistry of the soil.

The diagram shows a distinctive grouping of treated and

untreated soils. This would be expected from the addition of

lime and the consequent rise in calcium activity, aluminum

activity and pH (decrease in pOH) . Once again, the most

difficult variable in the equation is determining the

appropriate range for the activity of aluminum. In general,

it is well known that the activity of dissolved aluminum

increases rapidly above a pH of approximately seven (refer to

Figure 7). Although the presence of sulfate has been shown to

affect aluminum solubility in acid solutions, there does not

seem to be much change for basic solutions (Nordstrom, 1982).

On the basis of Figure 7, the activity of Al(OH)“ can be

assumed to be approximately 10'* for the untreated Stewart/Owens

soils. This has been shown on Figure 34 and is for soils with

a pH in the range of 7 to 8. The estimation of Kl(OH):

activity for lime-treated soils is more difficult. Based again

on Figure 7, the activity could exceed iO- for lime saturated

soil (pH=12.3). However, a more realistic value is probably

in the range of 10-» since this is the value measured in the

determination of ettringite solubility. An activity of 10'3 has

been shown on Figure 34 as a slightly conservative value. In

fact, if the activity actually does reach 10'3 at lime

saturation, 10 3 is not conservative at all.


207

The majority of the samples of untreated soils plot above

the a(Al(0H)„‘) = 10'6 boundary, beyond the stability field for

ettringite. Samples S-4A, S-12B, S-19C, and S-21C, however,

plot in the ettringite stability field. Because the pOH is

higher than the maximum 3.5 required to supply aluminum,

"native" ettringite should not actually be present in this

environment. The data simply indicate unusually high activity

levels for calcium and/or sulfate. In addition, the data

suggest that the real aluminum activity for untreated soils

may be more on the order of 10'7 or 10‘B.

For lime-treated Stewart/Owens soils, the ettringite

stability is defined by the estimated boundary for a[Al(OH)t ]

= 10-3. in this case, all of the lime-treated data points fall

within the appropriate stability field.

FVAT.TTATINr; SUSCEPTIBILITY TP T,TMF,-INDUCED_JiEAVE

Figure 34 has been reproduced at a larger scale without

the data points and presented as Figure 35. The diagram is

interesting as it demonstrates the profound chemical changes

incurred by lime treatment of sulfate bearing soil. For

example, the calculated activities for several Stewart Avenue

soils have been plotted along with the calculated changes due

to addition of 4.5 weight percent CaO. The diagram shows a

pronounced shift well into the ettringite stability field. In


mixing soil with 4.5% CaO.

■NJ
209

fact, unless the native soil has a p[Ca]6 [S04]3 of greater

than, roughly 36, ettringite will always be the stable phase.

Assuming a gypsum source, this corresponds to an extremely low

sulfate level of less than 15 mg/kg of soil at pH = 12.3. The

figure is simply an indication of the very strong potential for

reaction at the high pH. In reality, there must be some

threshold value of sulfate below which reaction kinetics will

inhibit ettringite nucleation. It is not known what that value

might be, however, it is likely to be considerably more than

15 mg/kg. Distress occurred on Stewart Avenue in areas with

initial soil sulfate levels as low as 700 mg/kg, however, the

situation is complicated by porewater chemistry and ion

mobility. A great deal of additional, long-term

experimentation would be necessary to determine the threshold

sulfate value. One difficulty would be in detecting the

earliest, minute quantities of ettringite. Again, due to ion

mobility, the threshold level may not even be important from

a practical standpoint. As has been demonstrated at

Stewart/Owens, a great deal of sulfate can be derived from

outside the system.

in practice, a certain amount of sulfate can be tolerated,

even if it does cause minor heave. This is particularly

for flexible pavements. Figure 36 shows the calculated percent

expansion as a function of the initial sulfate content of the

soil. The diagram uses an assumed density of 100 pcf and

represents the probable maximum volume increase if all of the


PER CEN T VOLUME IN CREASE

Mg/Kg SU LFA TE
Figure 36 Calculated maximum volume inrease induced by complete reaction
of sulfate in lime treated clay soil. Dry density assumed to be 100 pcf.
211

sulfate reacts. Some, perhaps half, of the sulfate, will not

react so the figure is properly conservative in that regard.

The relationships shown are based on laboratory

experiments, assumptions that may not be entirely valid, and

site specific empirical data obtained from soils on

Stewart/Owens Avenue in Las Vegas. In addition, other factors

such as the availability of water and ion mobility have

tremendous bearing on whether or not nucleation of e t t n n g i t e

will even result in distress. On Stewart/Owens even the rock

hard, undamaged, subbase contained minor ettringite. A

with no measurable sulfate could, conceivably, react if sulfate

were brought in from an outside source. This could occur from

surface water, capillary porewater or from biogenic sources,

as may have happened in Tennessee. Other sources of sulfate

from outside the soil system may be possible. The writer, the

University of Nevada, the Desert Research Institute and SEA

INC. make no warranty, expressed or implied, that the use of

Figures 35 and 36, and other data in this research, will not

result in evaluations that are either overly or insufficiently

conservative.

engineering CONSIDERATIONS

The potential for lime-induced heave can be predicted for

the two extremes. If the geochemical environment contains less

than 15 mg sulfate per 1000 gm soil (50:1 extraction ratio)


212

sulfates, no lime-induced distress should occur. At the other

end of scale, where soluble sulfates approach 1% (5000 mg/kg),

clay minerals account for at least 10% of the soil, and the

environment is frequently saturated, significant (> 1/2 inch)

lime—induced heave will occur. The difficulty, then, is

determining heave potential between these extremes and

establishing a safe set of guidelines that can be used by the

practicing geoscientist.

It would be easiest to exclude all soils with measurable

sulfate, and given the state of the art, this might be the

appropriate conservative approach. The risks involved are

high. With Stewart Avenue, lime treatment might have saved

$500,000 in initial construction costs over some of the other

alternatives. Repair of lime-induced distress, along only 2/3

of the roadway, however, cost $2,700,000. In addition,

liability to the engineer for lime-induced heave has increased

with awareness of the problem. Unfortunately, excluding all

sulfate soils from lime stabilization is overly conservative

and may incur excessive construction costs.

A limited amount of swelling produced by low sulfate

levels can readily be tolerated, provided the R-value does not

decrease significantly. At Stewart Avenue, however, lime-

induced heave occurred in areas where native soils may have

an initial sulfate content as low as 700 mg/kg but were

adjacent to a major water source. In addition, sulfate levels


as high as 20,500 mg/kg were recorded in native soils from

undamaged areas where subbase had to be removed with a

jackhammer. The majority of these areas were east of Lamb

Boulevard where clay content was generally low. Sulfate levels

as high as 43,500 mg/kg were recorded in undamaged subbase.

In these cases, sulfate was tied up in gypsum rather than

thaumasite. The particular value of 43,500 mg/kg was found in

the same trench as, and only a few feet away from, an area of

severe heave. The heave followed a linear trend adjacent to

a water main.

In conclusion, the results of this study indicate that

lime treatment of sulfate-bearing clay soils is risky, even at

relatively low sulfate concentrations. It does appear that

approximately 10% by weight, or more, clay-sized particles are

needed for adverse reactions. While both sulfate and clay

minerals are required, their threshold concentrations are

complicated by ion mobility in soil pore water. Excessive

water may mobilize and concentrate necessary ions in areas

where distress might otherwise not occur. Since diffusion

through saturated fine-grained soils is extremely slow, lime-

induced heave occurs preferentially in areas where water flows

more freely. This includes, but is certainly not limited to,

utility trench backfill and construction joints.

If utility trenches use impermeable synthetic liners and

are sloped to drain and construction joints are adequately


sealed, it should be possible to successfully use lime

stabilization in soils with high concentrations of both sulfate

and clay minerals. Excellent surface drainage would also be

mandatory to prevent ponding of water. The ultimate technical

solution would include an impermeable membrane or capillary

break beneath the lime stabilized soil. This should be quite

effective; however, removing and replacing the soil in order

to install the barrier would offset lime's major cost advantage

which is in-place stabilization.

Since cement-sulfate-clay reactions are similar to those

of lime-sulfate clay, similar risks are present for soil

cement. Heave in soil cements has also been documented where

sulfate is present (Lukas 1975).


215

REFERENCES

American Public Health Association, American Water Works


Association and Wastewater Pollution Control Federation,
1980, Standard Methods for Examination of Water ana
Wastewater, 15th Edition, 1193 p.
Barnes, H. L. , 1980, editor, Geochemistry of hydrothermal ore
deposits: New York, John Wiley and Sons.
Bell, J. w . , 1982, Subsidence in the Las Vegas Valley: Nevada
Bureau and Mines Geol., Bull. 95, p. 13“1^;
Bensted, J. , and Varma, Prakash, 1973, Studies of Ettringite
and Its Derivatives, Part IV: The low-sulphate form
calcium sulphoaluminate (monosulphate): Cement Technology,

l\
Berner, R. A., 1971, Principles of Chemical Sedimentology.

Bohm!r~ r d anS J°acobson! L 1981, WATEQDR an updated

C a U ? c r n I a VaDepartm ent of Transportation 1976


stabilization study: A selected lltera\ur® rJ}T£rt
Materials and Research Department, Research Repor

C a l i f o r n i a ^ e ^ r t m e n t of Transportation, 1 9 7 4 , Highway Research


Report-Lime Soil Stabilization Study, P h a s e I I , Laboratory
Investigation of California Soils for ^ m e Reactivity.
Transportation Laboratory Research Report CA-DOT TL 2812

California6 d i v i s i o n of H i g h w a y s “ ? select^
Department, 1967, Lime-soil stabilization,
literature review, 94 P; 1974, subgrade Soil

coS g a £ r
Street to Nellis Boulevard, Las Vegas, Nevada. v
Consultant Report 40 p. 1Q7r proiect No.
converse, Davis, and Associates June 13 ^ V a / e ^ting,
N-75-808-D, Final Report on Lime Tr « unpublished
Stewart Avenue Improvements, Las Vegas,
consultant inspection Report 7 p. cement exposed to

sulfates: Highway R e s . ^ r d , descriptive

Mineralogy": John Wiley and Sons, New^York, l” 1 . 1 ‘3 ^

Dapulveration M n d S E ' & r & o T « ^ l i « stabilization:


Highway Res. Rec., 92, p. 103 126.
216

Dinger, J. S., 1977, Relation between surficial geology and


near-surface groundwater guality, Las Vegas Valley, Nevada:
Unpublished PhD dissertation, Department of Geology,
University of Nevada, at Reno.
Drever, James I., 1982, The geochemistry of natural waters:
Prentice-Hall, Inc., Englewood Cliffs, N. J., 388 pp.
Eades, J. L. , and Grim, R. E. , 1966, A quick test to determine
lime requirement for lime stabilization: Highway Res. Rec . ,
139, p. 62-72. ^ ^ ^ , , ,.
Eades, J. L. , and Grim, R. E. , 1960, Reactions of hydrated lime
with pure clay minerals in clay stabilization: Highway
Research Board, Bulletin No. 262, p. 51-63.
Edge, R. A., and Taylor, H. F. W., 1971, Crystal structure of
thaumasite: Acta crystalography, Bulletin 27, 594-601.
Font-Altaba, M. , 1960, A thermal study of thaumasite:
Mineralogical Magazine, 32, 567-572.
Fournier, R. 0., and Marshall, W. L %, 1983, Calculation o
amorphous silica solubilities at 25 to 300 C and apparen
cation hydration numbers in aqueous salt solutions using t
concept of effective density of water: Geochemical
Cosmochimica Acta, V. 47, No. 3, p. 579-586.
Gaffron, Winston, 1988, Tennessee Department of Transportation,
Mochvilie Tennessee: Personal Communication.
Garrels R.' m ., and Christ, C. L . , 1965, Solutions, Minerals
and Equilibrium: San Francisco, Freeman, Cooper and Company.
Gouda, G. R., Roy., D. M . , and Sarkar, A., 1975, Thaumasite in
deteriorated soil cements: Cement Concr. Res., 5 / 519-522.
Graf, Edward, 1986, Pressure Grout Inc., San Francisc ,
California: Personal Communication. Mr,rraw_H n i
Grim, R. E., 1968, Clay mineralogy: New York, McGraw Hill,

Habib P 'p and Aversence, D., 1987, Gonflement d'un sol


H a oonienant des sulfates et traits S la chaux et au ciment:
Revue Francaise de Geotechnique, 42, 61 64. Nevada-
Heidker, James, 1988, Desert Research Institute, Reno, Nevada.

H i £ T 2 . ,
C° a " d ^ n , D. *., I960, Lime fixation in clayey

Hol?isS: BHlgr y \endear?awcett:tlnN.2 6 D., ^oratory


investigation of the use of mixtures of lime and pulverized
fuel ash for soil stabilization: Roads Road Constr.,
44(517), 3-6.
Lime-induced heave in sulfate-bearing clay
Hunter, Dal, 1988,
Geotechnical Engineering, ASCE, 114(2),
soils: Journal of
150-167. 1960, Ettringite from Franklin,
Hurlbut, C. S. and Baum, J. L. Mineralogist, v. 45, No. 11 and
New Jersey: The American
12, P. 1137-1143.
217

Kelley, Conard M. , 1977, A long-range durability study of lime


stabilized bases at military posts in the southwest:
National Lime Association Bull. 328, Washington, D. C. , 29

Kirov, G. N., and Poulieff, C. N., 1968, On the infra-red


spectrum and thermal decomposition products of thaumasite.
Mineralogical Magazine, No. 36, p. 1003-1011.
Kollmann H. , 1978, Minerlogische untersuchungen uber
ausbluhungs und treiberscheinungen and baustoffen durch
sulfate: Giessener Geologische Schriften, Giessen, Germany,

Kollmann’, H. , Strubel, G. , and Trost, F. (1976). _"Synthesen


von Ettringit und Thaumasit." Fortschritte der Mineralogie,
Beihefte, 54(1), p. 50—51. „„ Mlnpral-
Kollmann, H. , Strubel, G. , and Trost, F. , 1977, Mineral
synthetische untersuchungen zu treibursachen durch Ca
sulf at und Ca-Si-carbonat-sulf at hydrat." Tonmdu s t r i e
Zeitung, 101(3), Germany, 63-70. m q 77 ' i

K° S i o n s ^ h a , = ^ r
gips-putzen durch ettringie und thaumasit. Zement K

Kollmann"^0 #H ^ ,* and"2Strubel, G. 1981,


Mischkristalle von Brenk (Eifel): Chemical Erde, 40,
Giessen, Germany, 110-120. r^nnhpmistrv*
Krauskopf, Konrad B, 1979, introduction to Geochemistry.
McGraw-Hill, Inc., New York, N. Y. , 617 pp.
TaHd C C Moh. Z. C., and Lambe, T. W., 1960, Receive sox
u M e Mesiirch i t the Massachusetts Institute of Technology.
Highway Res. Board Bull. 262, 64-84.
^ o v ^ t - o f ^ ? 1^ S i S T u l i c;M p o U ;

L e h m S n ? Y JR !S ^ r i n g m “^laache?
Gebiet^iH^Jeifes1J^hbtich de? Mineralogie und Geologie, 6, 273-
275. -o u Rffl,rvpr Ben. and Roberts, R.
Longwell, C . p o s i t s of Clark County,
NiMadaf 'Nevada Bureau of Mines and Geology, Bulletin 62,

Lukas,^W., 1975, Betonserstorung SO,.angriff unter bildung


von thaumasit und woodfordit: Cement Concr.
^ * — •-i -1.!-mo ypc: T"Ch clt IOWcl St3.t0
Mateos, Manuul, 196 ' ° soil Mechanics and Foundations
Division^ASCE ^Proceedings ,Syol • ?1. *>• * . ^ u _ fly
M°ashe and ^ t h e f U m e ^ a c l f v e ^ a t e r i a l s : Highway Research
Board, Bulletin 231, p. 60-66.
218

McDowell, Chester, 1966, Evaluation of soil-lime stabilization


mixtures: Highway Research Record, 139. ,
McDowell, Chester, 1972, Flexible pavement design guide.
National Lime Association, Bulletin, 327.
Mehra, S. R. , Chadda, L. R . , and Kapur, R. N. , 1955, Role of
detrimental salts in soil stabilization with and without
cement I. The effect of sodium sulfate: Indian Concr. J . ,

Mehta,3p. K., 1986, Effect of fly ash composition on sulfate


M resistance of cement: American Concrete Institute Journal,
Mi“ i r ^ “ ^ 8D4r9^ r b|ci1luS = Thonindustrie-

M i H U r k NCi . “ ^ E “ « . M . , 1979. P l u v i a l l a . e s ana


e stim a te d p lu v ia l c lim a te s o f Nevada: N evada Bureau M in e s

Mitcheii?UJ^ K w iP
6
8
9
l
' ractical p r o b l e m s l r o m surprising soil
behavior: J. Geotech. Engrg. Div., A 3 CB, 112( 3 ), 259 289.
Mitchell, James K., 1988, Department of Clvl1.
University of California, Berkeley, California: Personal

„ o ™ ! CEa ^ ° a nn a T a y l o r , H . E . W 1970 C ry sta l stru ctu re of


ettringite: A c ta Crystalography, B u l le t i n 2 6 , 3^ . ^

n ? S i l J f r o r c r t at r e S,' SilifV: Geological society of


America, Bulletin 69, 1620-1621.
A., 1960, Ettringite
Murdoch, J. , and Chalmers, California: American
("Woodfordite") from Crestmore,
Mineralogist, 45, 1275-1278. mineral:
Nordenskiold, A. E., 1878, Thaumasite, ein neues m m .
Compt. rend. 87 p.. 3 1 4 . ffect of sulfate on aluminum
Nordstrom, D. K. , 1982, m e e n c ___ofahilitv Relations
concentrations in natural waters: B o n e Stahl ^
in the System Al203 " s03 “ H ^° ar
P e c r G e ? a l 1 r ai9 A7 t % e x i s ( D
42pa6S e n t o f T ra n s p o rta tio n , A u s tin ,

Ra“ , Lo?en:n i 9 8 3 M ” ;hCSep“ tment of Transportation, Salt


Lake City, Utah: Personal 197A Encyclopedia
R °ofMinerals; V a n ^ o s t r a n d Reinhol^Co. f New York, p.^197-198;

Ruff^ C . G. , and Clara, Ho. . 1966- '


Time--t-P-ature^trength-
reaction product relationships m 11 42-6 0 .
mixtures: Highway Research eco ' r city of Las Vegas,
SEA, Inc. , 1983, e T o fliis B W o Las Vegas,
Stewart Avenue 28th Street e
Nevada: Unpublished consultant repor .
219

SEA Inc., 1988, Evaluation of continuing pavement distress,


Owens Ave., Las Vegas, Nevada: Unpublished consultant

S h e r w o o d P . T. , 1958, Effect of sulfates on cement-stabilized


clav: Highway Res. Board Bull, 198, 45-54.
Sherwood, P. T. , 1962, Effect of sulfates on cement and lime
treated soils: Highway Res. Board Bull. 353, 98"107* .
Taylor, H. F. W. , 1964, The chemistry of cements: Academic
pvpqc New York, Vol. 1, 460 p.
T a y l o r / W. H. , and Arman, Ara, 1960, Lime stabilization using
precondition soils: Highway Research Board, Bulletin 262,

Thompson, M. R., 1966, Lime reactivity of Illinois soils:


J. Soil Mech. Found. Div., ASCE, 92(5), 67 92. survey
United States Department of Agriculture 1972, lltl
laboratory methods and procedures for collecting soil
samples: Soil Survey Investigation Report No 1, 31

- ? o S : = -
related substances in the presence^ of sulphates, cemenr ana
Concrete Research, Vol. 5, p. 225 232' ,t . Arkj_v for
Welin, E., 1956, Crystal structure of thaumasite.
Mineralogi och Geologi, Sverige^ 2 1 • California's
Zube, E., Gates, , and Hatano, FL, ^ ' construction:
experience with lime trearme
California Highways and Public Works.
Sample Number S-2A S-2B S-2C S-3A
Sieve Size Percent By Weight Passing

1-1/2 Inch 100 100 100


1 Inch 100 100 94 96 90 100
96
3/4 Inch 97 83 87 91 81 100 100
1/2 Inch 89 64 77 80 100 74 97 93 92
3/8 Inch 78 56 100 66 75 99 68 91 86 89
No. 4 62 46 98 54 63 98 58 81 72 85
No. 8 52 44 47 74 • 57
No. 10 35 97 54 98 81
No. 16 45 36 36 67 48
No. 20 - - - - - - “ “
No. 30 41 - - 28 - - 25 59 45 -
No. 40 _ 25 96 - 46 97 - - - 74
No. 50 36 _ - 22 - - 17 50 43 -
No. 60 _ 22 94 - ’ 42 97 - - - 70
No. 100 32 19 72 21 40 95 13 42 41 61
No. 200 19 12 52 17 29 91 6 38 38 46

Hydrometer Percent Fines of Total Sampl e


Silt T.U74-.005mm) _ 11 20 - 22 44 35 15 ~

Cl ay (.005-.001mm) - 1 20 - 5 32 1 21
Colloids (<.001mm) - 1 12 2 15 ” 2 2

Liquid Limit . _ 42 _ 33 36 - 59 57 30
Plastic Index NP NP 19 NP 4 14 NP 20 43 9
Moisture Content (%) 8 25 25 8 35 17 20 37 21 7

FIELD DENSITY TESTS

Depth (feet) 0.46 1.25 2.00 5.75 1.08 2.50 0.50 1.00 2.50 0.0

In-Place Dry D e nsit y-( pcf) 125.1 89.9 100.2 122.3 78.1 99.5 112.3 96.7 110.8 98.4

Maximum Dry D e nsit y-( pcf) 141.4 120.0* 125.9 129.5 120.0* 118.1 129.0 119.0* 114.9 123.1

Relative Compaction (%) 88.5 74.9 79.6 94.4 65.1 84.2 86.8 81.3 96.4 79.9

11.3 9.4 13.0* 14.6 9.0 12.5* 16.8 9.0


Optimum Moisture (%) 6.0 13.0*
& Assoc
Lime Treated Subgrade Testing, Stewart Avenue", 7 - 1 - 7 5 :, Converse Davis
nhtainpd from "Final Reoort on

220
Appendix 1 a
MECHANICAL ANALYSIS
S-6B S-6C S-7A S-7B S-7B1 S-7C S-8A
Sample Number S-4B S-4C S-4D
Percent By Weight Passing
Sieve Size
92 100 100 100 100
1-1/2 Inch 69 94 97 96
1 Inch
_
87 -

_ 85 100 69 92 93 95
3/4 Inch 63 83 89 89
1/2 Inch 100 100 80 99
74 99 59 78 87 100 82
3/8 Inch 99 97
57 90 49 75 84 98 67
No. 4 100 97 89 54
_ 42 82 39 -
No. 8 71 80 95
No. 10 99 95 82 - - -
42
34 76 29 -
No. 16' _ _ - 69 77 94
No. 20 21 32
No. 30 29 71 -

72 _ - - 65 75 93
No. 40 97 82 25
_ 25 67 15 -
No. 50 61 73 93
No. 60 95 73 68 - . - -
22
57 57 22 65 13 SB 72 92
No. 100 85 9 46 67 86 18
No. 200 59 32 39 18 59

Hydrometer Percent Fines of Total Sample


_ 17 23 - 42 31
Silt (,074-.005ram) 10 15
Clay (. 005- .001mm) 19 _ 1 30 - 3
1 6 1 40
Co lloi ds (<.001mm) 3 -
33 51 24 - 80 52
Liquid Limit 32 24 27 NP
17 3 NP 22 NP
Plastic Index
Mo istu re Content (%)
17
13
6
7
14
15 36 9 54 21 30 7

FIELD DENSITY TESTS


3.00 0.43 1.1 1.00 2.67 0.50
Depth (feet) 2.00 2.50 4.17 1.00
73.8 102.2 129.3 61.1 90.5 87.9 131.1
In-Place Dry D e n s i t y - ( p c f ) 87.1 85.8 102.2
106.0* 122.9 141.0 106. 106.0* 106.9 141.0
Maximum Dry D e n s i t y - (p c f ) 119.0* 123.1 123.1
83.2 90-95 57. 85.4 82.2 90-95
Relative Co mpac tio n (%) 73.2 69.7 83.0 69.6
18.5* 12.2 6.0 18. 18.5* 18.5 6.0
Optimum Mo istu re (%) 12.5* 10.8 10.8
Assoc.
Lime Treated Subgrade Testing, Stewart Avenue", 7-
*Data o b tain ed from "Final Report on

Appendix 1 b
MECHANICAL ANALYSIS

S-8B S-8C S-8D S-9A S-9B S-9C S-9D S-10A S-10B S-10C
Sample Number
Sieve Size Percent By Weight Passing
_ 100 100 -
1-1/2 Inch _ _ 96 97 -
1 Inch _ _ 95
100 100 94 -
3/4 Inch _ 88 92 100
1/2 Inch 92 99
_ _ 98 84 88 99
3/8 Inch 86
78 100 100 100 100 - 96 71 84 98
No. 4 _ _ - 56 “
No. 8 74 99 - -

99 99 98 100 95 - 81 96
No. 10 43 -
No. 16 71 98 - - - -
98 _ - - - “ - -
No. 20 _ - - 32 - -
No. 30 68 97
96 97 82 94 82 - 73 94
No. 40 _ _ - - 24 " “
No. 50 65 96
94 93 71 92 71 - 68 90
No. 60 49 18 62 90
No. 100 64 91 92 88 55 84
87 84 66 55 71 30 13 53 49
No. 200 54

Hydrometer Percent Fines of Total Sample


Silt (TU74-. 005mra) 50 74 _ 19 51 - - 48 14
Clay (.005-.001mm) 2 8 _ _ 29 17 - 1 7
7 3 " 4 28
Colloids (<.001mm) 2 5 - -

56 31 ■ 30 52 39 24 - 82 38
Liquid Limit 70 18
20 33 18 13 31 23 10 NP 25
Plastic Index 5 57 15
Moi sture Content (%) 59 32 23 11 10 17 15

FIELD DENSITY TESTS


4.50 0.00 1.50 2.50 4.00 0.46 1.33 3.33
Depth (feet) 1.25 3.00
87.4 95.5 93.7 74.7 73.0 69.3 126.5 61.5 88.2
In-Place Dry De ns i t y - ( p c f ) 59.3
106.9 108.8 105.0* 118.5 112.9 141.0 108.0* 108.0
Maximum Dry Densit.y-(pcf) 105.0* 106.9
89.3 86.1 71.1 61.6 61.4 88-92 56.9 81.7
Relative Compaction (%) 56.5 81.8
19.2* 10.0 12.9 0 18.5* 18.5
Optimum Moisture (%) 19.2* 18.5 18.5 13.8
, Converse Davis & Assoc.
Lime Treated Subqrade Testing, Stewart Avenue ", 7-1-75
obtained from "Final ReDort on
ME C H A N I C A L AN ALYS IS
S-11B S-11C _
S-11D S-12A S-12B S-12C S-12D S-13A
Sample Number S-10D S-11A
"Percent By Weight Passing
Sieve Size
100 100
1-1/2 Inch _ 100 - - -
_ - 92 88
1 Inch 92 100 -
73
- 100 86
3/4 Inch 100 85 99 64
96 - 98 83
1/2 Inch 99 75 80 59
98 68 93 100 100 95
3/8 Inch 99 92 76 47
No. 4 98 56 85 99
_ - 89 39
No. 8 44 - -
73
78 99 99 100 100
No. 10 94 -
30
34 86
No. 16 _ - - - " ~ “
No. 20 83 22
No. 30 26 68
70 97 96 97 97
No. 40 83 16
18 80
No. 50 92 94 64
No. 60 76
63 11
64
57
96
92
95
91 73 85 87 63 10
No. 100 75 59 69 67 34 5
No. 200 43 4 47 77
Percent Fines of Total Sample
Hydrometer - -- W 59 -
Silt (.074-.005mm) 44 33 -

2 39 - - 9 5 -
Clay (. 005- .001mm) 5 3 -
Colloids (<.001mm) - - 1 5 -

47 37 41 37 58 42 57
Liquid Limit 29 36 22 30 NP
14 NP 14 22 19 18
Plastic Index 19 7 12 14 13 3
Moisture Content (%) 14 21 39 29
FIELD DENSITY TESTS
4.33 0.00 1.17 2.33 3 0.12
4.00 2.50 18.0 3.00
Depth (feet)
82.5 103.4 92.1 62.8 77 135.8
In-Place Dry Densit,y-(pcf) 89.7 107.4 65.3 73.9
108.0 108.0 108.0 108.0 - 141.0
Maximum Dry D e n s i t y - ( p c f ) 125.6 130.0 108.0* 108.0
76.4 92-96 85.3 58.1 - 95-97
Relative Compaction (%) 71.4 80-85 60.5 68.4

18.5-* 17.6 ■ 17.6 14.3 14.3 14.3 6.0


Optimum Moisture (%) 11.7 9.0
& Assoc.
Lime Tr eated Subgrade Testing, Stewart Avenue", 7-1-75, Converse Davis
I *nata obtained from "Final Report on
MECHANICAL ANALYSIS
S-14A S-14B S-14C S- 15A S-15B S-15C S-16A S-16B
Sample Number S-13B S-13C
Percent By Weight Passing
Sieve Size
100 . 92 - - 100
1-1/2 Inch 85 81 100
1 Inch 89 100 -

87 94 100 80 100 74 95
3/4 Inch 100 88
100 83 83 97 75 95 68
1/2 Inch 98 63 83
94 99 76 74 95 71 90 -
3/8 Inch 59 81 50 68
No. 4 83 99 63 54 83
No. 8 51 _ _ 47 - 38
38 77 68 100 58
No. 10 71 98 -

40 - - 34 “ 28
No. 16
No. 20 _ 23 - - 21
No. 30 31 50
55 97 22 70 - 49 98
No. 40 _ _ 15 - 16
No. 50 24 49
50 96 18 69 - 44 96
No. 60 66 10 38 92 12 43
No. 100 45 88 18 15
37 70 14 11 54 5 29 72 8 35
No. 200
Hydrometer Percent Fines of Total Sample
8 25 20 42 30
Si U""{T074-. 005mm) 23 58 3
7 2 6 - 8 26 -
Clay (. 005- .001mm) 8 - 4 2
6 5 1 23 1
Colloids (<.001mm)
46 52 27
Liquid Limit 30 NP NP 5 NP NP
Plastic Index NP 24 NP NP 18
22 30 7 55 21 6
Moisture Content (%) 41 25 7
FIELD DENSITY TESTS
2.00 • 0 .48 1.00 1.50 0.46 1.00
Depth (feet) 1.50 3.00 0,.46 1.17
88.4 123 .8 54. 4 73. 6 131.5 86.3
In-Place Dry D e n s i t y - ( p c f ) 61. 4 64.5 118 .5 76.7
107.0* 103.4 130 .0 107.,0* 98. 6 141.0 107.0*
Maximum Dry D e n s i t y - ( p c f ) 106.,0* 101.3 130 .0
71.7 85.5 90- 95 50..8 74.,6 90-95 80.6
Relative Compaction (%) 57..9 63.7 90- 95
19.8 11.0 19..4* 21..5 7.0 19.4*
20.518,.5*11.0 19.4*
Optimum Moisture (%) & Assoc.
Treated Subgrade Testing. Stewart Avenue". 7-1-75, Converse Davis
*Data obtained from "Final Report on Lime

224
A p p en d ix 1e
225
i
MECHANICAL ANALYSIS
S-21B S-21C S-22A S-22C S-22D S-23A S-23B
Sample Number S-19C S-19D S- 21A
Percent By Weight Passing
Sieve Size
100 _ - 100 -
1-1/2 Inch 100 92 97
1 Inch 100 100 100 100
94 87 89 87 100
3/4 Inch 78 90 96 75 88
73 81 89 77 68 81
1/2 Inch 68 60 76 - 67 83
3/8 Inch 69 79 83 72
50 45 68 100 55
No. 4 62 71 69 46
58 54 39 48
No. 8 99 100 62
No. 10 64 45
35 35 38
No. 16 54 42
No. 20 32 24 31
No. 30 49 32 46
35 98 97
No. 40 52 24
43 23 28 16
No. 50 97 93 42
45 30 38
No. 60 26 21 10 96 91 18
No. 100 37 35 16 31
23 12 4 86 86 12
No. 200 28 21 9
Percent Fi nes of Total Sampl e
Hydrometer 10 46 ay
Silt (.074-.005mm) 24 15
5 1 33 - 1
Clay (. 005-.001mm)
Colloids (<.001mm)
1
3
- ; 3 1 _ 7 - 1

35 29 73
Liquid Limit 52 14
17 NP NP 16 10 NP
Plastic Index NP NP NP 4 39
5 14 3 10 22 10
Moisture Content [%) 7 5
FIELD DE NSIT Y TESTS
3.00 0.50 2.00 3.50 0.21 1.00
2.50 3.50 0.50 1.08
Depth (feet)
121.1 128.2 84.2 84.4 132.2 65.0
In-Place Dry Densit,y-(pcf) 109.9 106.5 121.2 94.4
127.5 141.0 117.3 117.3 141.0 115.0*
Maximum Dry Densit.y-(pcf) 125.2 125.2 130.0 1 0 7.0**
95.0 90-95 71.8 72.0 90-95 56.5
Relative Compaction (%) 87.8 85.1 90-95 88.2
7.0 14.4 14.4 6.0 15.0*
8.0 8.0 9.0 19.4* 11.1
Optimum Moisture (%)
rse Davis & Assoc.
Lime Treated Subgrade Testing, Stewart Avenue", 7-1-75 , Conve
*Data obtained from "Final eport on

A ppendix 1g
MECHANICAL ANALYSIS

Sample Number S-23C S-24A S-24C S- 25A S-25B S-25C S-26A S-26B S-26C S-27A
Sieve Size ---- ---- Percent B.y Weight Passing
- - - - 100
1-1/2 Inch 100 94
1 Inch 82 _ 100 - - 100 100
_ 99 100 - 95 98 88
3/4 Inch 75
_ 93 98 - 78 90 80
1/2 Inch 72 75
3/8 Inch 100 69 - 85 97 68 81
100 71 84 - 51 64 60
No. 4 98 56 46
t No. 8 45 - 57 - 39
97 _ 99 - 59 " 53 100
No. 10 30 34
No. 16 35 - 44 -
No. 20 “ _ - 23
26 32 22
No. 30 _ - 100 39 98
No. 40 96 83 26
_ 23 - - 17 16
No. 50 19
95 73 - • 20 99 " 36 96
No. 60 16 99 12 33 93 11
No. 100 95 12 64 16
6 58 9 11 94 8 27 78 6
No. 200 93
Percent Fines of Total Sample
Hydrometer
82 _ 54 - --------8 ------58------ - 24 73
Si It (.074-.005mm)
Clay (.005-.001mm) 7 _ 2 - 1 2 “ 2 1
Colloids (<.001mm) 4 - 2 - 2 4 _ 1 4

. 30 39 _ - 25 - 48
Liquid Limit
19 NP NP 8 NP NP 26 NP
Plastic Index
Moisture Content (%)
11
16
NP
6 20 6 14 16 6 39 18 6

FIELD DENSITY TESTS


0.50
Depth (feet) 2.67 0.50 2.00 0.54 1.50 2.50 0.48 1.00 2.17

90.4 126.4 109.3 92.6 96.7 62.7 96.7 131.1


In-Place Dry D e nsit y-( pcf) 106.6 119.1
130.0 113.0* 114.4 125.9 105.0* 117.2 141.0
Maximum Dry D e nsit y-( pcf) 118.0 125.9 115.2
78.5 95-98 96.7 80.9 76.8 59.7 82.5 90-95
Relative Compaction (%) 90.3 94.6
• 16.4* 16.6 9.3 19.2* 14.3 7.0
14.4 9.3 14.0 9.0
Optimum Moisture (%)
, Converse Davis & Assoc.
iimp Treated Subgrade Testing. Stewart Avenue", 7-1-75
♦nata nht.ain.ed_.from "Ftaal ReDort on

227
Appendix 1h

---------------------------------------------------------------------- .a
M E C H A N I C A L ANALYSIS

;-28B S-28C S-29A S-29C S-30A S-30B S-30C


Sample Number S-27B S-27C S-28A
-- Percent iTv Weight Passing
Sieve Size
- -
1-1/2 Inch
1 Inch 100 100
3/4 Inch 97 - 91
1/2 Inch 87 - 84 100
3/8 Inch 61 _ 69 99
No. 4 _ 55 -

No. 8 47 100 - 99 55 100


38. 10 38 100
- - 41 -
No. 16
No. 20 30 _

No. 30 32 99
No. 40 21 96 25 98
_
-

21
97
-

No. 50 21 94 - 94
No. 60 18 82 14 88
No. 100 10 61 6 76
No.' 200
Percent Fines of Total Samp!e
Hydrometer 56 69 a /H
10 60 “ 4

ram
— Silt "(7074-.005mm) 3 2 6
1 5 5 2
Clay (.005-.001mm) 1 3 _ 4 - 1 3
Colloids (<.001mm) 1 1
32 23
53 34 3
Liquid Limit 12 NP 13 NP NP
NP 27 NP NP 20 12
Plastic Index 37 26 6 18 5
Moisture content (%) 32 29 9
FIELD DENSITY TESTS
0.25 3.00 0.50 - 2.00
0.92 2.00 0.40 0.67 2.25
Depth (feet)
130.1 80.2 125.8 - 84.8
89.7 83.3 118.9 71.9 83.3
In-Place Dry D e n s i t y - ( p c f )
141.0 114.0 126.2 - 118.5
105.0 111.5 130.0 108.0 116.0
Maximum Dry D e n s i t y - ( p c f )
90-95 70.4 99.7 - 77.1
85.4 74.7 90-95 66.6 71.8
Relative Compaction (%)
7.0 15.3 9.8 - 10.4
13.6 11.0 18.5 ■ 13.7
Optimum Moisture (%) 19.2
Treated Subgrade Testing. Stewart Avenue", 7-1-75, Converse Davis & Assoc.
*nata ohtaijied from "Final Report on Lime
to
to
Appendix 1i co
MECHANICAL ANALYSIS

Sample Number S-31A S-31B S-33A S-33C S-34A S-34B S-34C S-B2 S-B3
Sieve Size Percent By Weight Passing

1-1/2 Inch 65
1 Inch _ _ 100 100 62 - - -
3/4 Inch 100 _ 81 87 50
1/2 Inch 89 _ 73 67 38
3/8 Inch 81 _ 67 54 34
No. 4 64 _ 53 34 24
No. 8 48 _ 40 21 19
No. 10 _ _ 100 - - 100 100 100
No. 16 34 - 29 14 15
No. 20
No. 30 23 _ 19 9 13
No. 40 _ _ 99 - - 98 97 94
No. 50 15 _ 13 6 11
No. 60 _ _ 98 - - 96 95 86
No. 100 10 _ 8 97 4 7 93 90 79
No. 200 5 - 4 80 2 1 89 88 77

Hydrometer Percent Fines of Total Sample


Si It (.074-.005mm) _ _ 68 - 57 58 69
Clay (.005-.001mm) _ _ 7 - - 23 19 5
Colloids (<.001mm) - - 5 9 11 -

Liquid Limit 34 26 _ _ 32 _ _

Plastic Index NP 9 NP 5 NP NP 10 - -

Moisture Content (%) 6 21 5 14 5 5 30 23 16

FIELD DENSITY TESTS

Depth (feet) 0.50 1.08 0.50 2.50 0.50 1.50 3.00

In-Place Dry De nsit y-( pcf) 119.3 114.5 122.2 86.9 126.6 117.8 79.2

Maximum Dry De nsit y-( pcf) 130.0 120.0** 130.0 117.8 130.0 - 116.2

Relative Compaction (%) 90-95 90-95 90-95 73.8 95-98 - 68.2

Optimum Moisture (%) 11.0 13.0 11.0 12.6 •11.0 - 13.8


, 7-1-75, Converse Davis & Assoc.
*Data obtained from "Final teport on Lime Tredted Subgrade Testing, Stewart Avenue
MECHANICAL ANALYSIS
0-2BU 0-2C 0-3B 0-3C
0-1B 0-1C 0-2B
Sample Number
Percent By Weight Passing
Sieve Size_
1-1/2 Inch 100 100
92 100
1 Inch 94 100 82
3/4 Inch 82 61
76 76 87
1/2 Inch 67 79 100 55 '
3/8 Inch 69 98 45
61 60 58
No. 4 43 96 32 100
54 100 38 97
No. 10 24 29 95 32
No. 40 44 99 30 95
97 21 93
No. 60 42 91 27 92
35 94 13 12
No. 100 5 82 22 85
26 88 8
No. 200
Percent Fines of Total Sample
Hydrometer
6 75 20
Silt (.074-.005mm) 20 42 1
42 1 - 3
Clay (.005-.001mm) 4 4 1
Colloids ( .001mm) 2 4 1 41
35 19
Liquid Limit 42 NP 26
23 NP NP 20
Plastic Index NP 17 28
28 23 5 10
Moisture Content

2.33 2.92 0.85 2.25


0.85 1.67 0.85
Depth (feet)
104.8 84.5 81.8 93,3
123.8 82.7 89.7
In-Place Dry De nsity-(pcf)
122.6 116.5 123.0* 112.0
135.0* 112.1 123.0*
Maximum Dry De nsity-(pcf)
85.5 72.9 62-67 83.3
90-95 73.8 70-75
Relative Compaction (%)
11.4 12.7 13.0* 15.0
8.1* 15.2 13.0*
Optimum Moisture (%)
Stewart Avenue", 7-1-75, Converse Davis & Assoc.
*Data obtained from "Final Report of Lime Treated Subgrade Testing,
MECHANICAL ANALYSIS

0-4BU 0-4C 0-5B 0-5C


Sample Number
Percent By Height Passing
Sieve Size
1-1/2 Inch 100 100
90 100 100 100 100
1 Inch 89 84
82 80 89 98 94
3/4 Inch 76 68 77
1/2 Inch 69 70 72
66 61 66 57 66
3/8 Inch 57 49 100
41 59 47 51 42
No. 4 37 140 29 37 99
No. 10 28 54 95
45 30 31 20 22
No. 40 10 17 92
9 43 26 28 17
No. 60 23 22 11 11 90
No. 100 8 40 87
34 19 14 10 6
No. 200 7

Hydrometer Percent Fines of Total Salts


9 - 4 77
Silt (.074-.005mm) 5 27 7
5 - 4 1
Clay (.005-.001mm) 1 1 3
Colloids ( .001mm) 1 2 1

41 - 56
-
Liquid Limit NP NP 34
Plastic Index NP 24 NP
21 20 3 26 27
Moisture Content 4 20

FIELD DENSITY TESTS

0.95 1.60 0.50 0.80 2.5


Depth (feet) 0.75 1.85
95.6 109.3 123.8 92.7 89.7
In-Place Dry Densit y-( pcf) 112.1 80.1
123.0* 119.0 143.7 123.0 117.3
Maximum Dry Density-(pcf) 123.0* 109.6
75-80 91.8 86.2 65-70 76.6
Relative Compaction (%) 87-92* 73.1
13.0* 12.5 4.9 13.0* 13.2
ODtimum Moisture (%) 13.0 15.2
& Assoc
Testing, Stewart Avenue", 7-1-75, Converse Davis
*Data obtained from "Final Report of Lime Treated Subgrade
Group
Major Divisions Symbol* Typical Names Laboratory Classification Criteria

0 60 _ ( D jo l’
W ell-graded gravels, gravel-sand » C , -------- g reater th a n 4 ; Q *
lu re s , lit t le or no fines 0 ,0 0 ,0 X 060

P o o rly graded gravels, gravel-sand n N o t m e e tin g a ll g ra d a tio n re q u ire m e n ts fo r G W


lu re s , l it t l e o r n o fines

A tte rb e rg lim its b e lo w " A " A b o v e “ A " lin e w ith P .l.


S ilty gravels, gravel-sand-silt m ix tu lin e o r P .l. less th a n 4 b e tw e e n 4 and 7 are border-

C la y e y gravels, gravel-sand-clay i A tte rb e rg lim its b e lo w " A ” d u a l s ym b o ls


lin e w ith P .l. gre a te r th a n 7
lures

0 60 , _ _ ( O jo l 1
W elt-graded sands, g ra v e lly sands, little Q j -------- g reater th a n 6 ;
o r n o fines 0 10 0 , 0 X 0 60

P o o rly graded sands, g ra v e lly sands, N o t m e e tin g a ll g ra d a tio n re q u ire m e n ts fo r SW

SI lit t le o r n o fines

A tte rb e rg l i m i t l above " A " L im its p lo t t in g in h atched


S ilt y sands, sand-silt m ix tu re s lin e o r P .l. less th a n 4 zone w ith P .l. b e tw e e n 4

A tte rb e rg lim its above " A ” re q u irin g use o f dual s y m ­


C la y e y sands, sand-clay m ix tu re s
lin e w ith P .l. g reater th a n 7 bols ,

In o rg a n ic silts and ve ry fin e sands,


ro c k flo u r , s ilty o r c la y e y fin e sands,
o r c la y e y silts w ith S light p la s tic ity
Plasticity Chart
In o rg a n ic clays o f lo w to m e d iu m
p la s tic ity , g ra ve lly c la ys, sandy clays,
s illy c la ys, lean clays

O rg a n ic s ilts a nd o rg a n ic s ilty clays o f


lo w p la s tic ity
1

In o rg a n ic silts, m ica ceous o r d ia to m a -


ceous fin e sandy o r s ilty soils, elastic
silts

In o rg a n ic clays o f h ig h p la s tic ity , <at


clays _____________ _

-) - O rg a n ic clays o f m e d iu m
p la s tic ity , o rg a n ic silts
to high

1
Pt Peat a nd o th e r h ig h ly o rg a n ic soils

Hi
, C M .„ d SM „ „ up . into tubdivitiont o. d ,nd u . . . .o , -o .d . ...............id , oniy. Sttbdivi.ion i, o .,.d on "• «« « " h’ "
L .L . is 28 o r less end th e P .l. is 6 o r less, the s u ffix u used w h e n L .L . is greater th a n 28 c o m b in a tio n s o f g ro u p s y m b o ls . F o r e x a m p le :
i^ B o rd e rlin , c la s s ific a tio n s , used fo r soils possessing c h a r.c ta ris t.e s o f tw o g ro u p s, are des.gnatad b y c o m b .n a t.o
GW -GC, < ra il-g ra ded gravel-sand m ix tu r e w ith c la y b in d e r.

Appendix 2 Uni fi ed S o i l s C l a s s i f i c a t i o n S y s t e m .
( S o u r c e : H u n t , 1 9 8 3 , p.
P 353.)
EXPANSIVE PRESSURE (PSF) VOLUME CHANGE (%)
233
Appendix 4 T rial dilution ratios
To+a I
S o lu b le S o lu tio n
Mn2+ Na+ K+ S0„2- CO,2- HC0-.” c r S I0 9 S a lts* pH
a i 3+ Ca2+ Mq2+

Mq/1000 am S o i l * * (ml 11l e q u l v a l e n t s / 1 I t e r )

900 275 9,600 0 2,0 00 600 1,000 17,575 •


S—13 <50 4,000 125 (10.15) 7; 5
( 0 .7 8 ) (0.15) (4.00) (0) ( 0 .6 8 ) ( 0 .3 4 )
Lime Treated « o .in ( 3 .9 9 ) (0.21) -

1,500 900 1,100 0 3,170 650 1,150 8,5 55 7 .4


S-1C <50 310 125 (3.75)
(0.12) (1.00) ( 0 .4 6 ) (0.46) (0) (1.07) ( 0 .3 7 )
N at iv e (< 0 .1 1 ) ( 0 .3 1 ) -

850 300 20,000 0 1,800 350 31 ,4 10


S-2B 8,000 no - (18.3) 7 .5
( 0 .1 8 ) (0.87) ( 0 .1 5 ) (8.33) (0 ) ( 0 .7 2 ) (0 .2 0 )
Lime Treated - (7 .9 8 )

1,000 900 800 0 3,0 00 400 6 ,5 0 0


S-2C 320 80 - (3.54) 7 .2
(0.32) (0.13) (0.87) ( 0 .4 6 ) (0.33) (0) (1.20) ( 0 .2 3 ) -
Nat 1ve -

1,400 420 6,750 1,300 0 1,200 1,350 15,245


S-3B <50 2,800 20 <1 9.7
(1.22) ( 0 .2 2 ) ( 2 .8 1 ) ( ( . )7) 0 ( 0 .6 8 ) (8.63)
Lime Treated « o .in ( 2 .7 9 ) ( 0 .0 3 ) « 0.1 )

1,800 850 24,500 0 2,865 1,150 1,350 41, 6 15


S-3C <50 9,000 100 <1 (22.94) 7 .7
( 0 .1 6 ) ( < 0 .1 ) (1.57) ( 0 .4 3 ) (1 0 . 2 0 ) (0) (0.94) ( 0 .6 5 )
N a t iv e (0.11) (8 .9 8 )

850 1,250 34,255


S-3D <50 8,000 210 <1 1,350 600 20,500 0 2,7 45
(19.73)
7.4
( 0 .3 5 ) «0.1 ) ( 1 .1 7 ) (0.31) ( 8 .5 4 ) (0) ( 0 .9 0 ) ( 0 .4 8 )
N at iv e ( 0 .1 1 ) ( 7 .9 8 )

60,000 500 192,000 0 2,100 7,850 - 299,990


S-4A 37,500 40 - (1 7 5 . 6 2 ) 8.1
( 0 .6 6 ) (52.2) ( 0 .2 6 ) (7 2 .9 5 ) (0) ( 0 .6 9 ) ( 4 .4 3 )
N at iv e - (3 7. 4 3)

1,750 35,000 0 2,1 00 2,150 1,450 55,675


S-4B 50 11,500 275 - 2,900 •7.7
( 2 .5 2 ) ( 0 .8 9 ) (1 4. 5 6) (0) (0.84) ( 1 .2 1 ) (31.96)
N at iv e ( < 0 . 11) (1 1. 4 8) ( 0 .4 5 )

5,500 21,000 0 2,650 800 - 39,750


S-4C - 8,500 10 1,300 (22.52) 7 .7
( 0 .0 2 ) ( 1 .1 3 ) ( 2 .8 1 ) (8.74) (0) ( 0 .8 9 ) ( 0 .4 5 )
N a t iv e ( 8 .4 8 ) -

• E x c l u d e s A l 3 + , Mn2 + , S I 0 2

• • L i m e T r e a t e d and N a t i v e S o l i s e x t r a c t e d a t 50 : 1 d i l u t i o n ra tio . Base Sample e x t r a c t e d 5:1 d i l u t i o n ra tio

A p p en d ix 5a
Tota I
S o lu b le S o lu tio n

A 13+ Ca2+ Mq2+ Mn2+ Na+ K+ S0a2" C 0,2" hco -t cr S107 S a lts* pH

Mq/1000 qm S o i l * * (ml 11l e q u l v a l e n t s / l I t e r )

850 250 6,500 0 1,850 450 12,510


S-63 2,5 00 110 - (7.25) 7.7
( 0 .1 8 ) (0.74) (0.13) ( 2 .7 1 ) (0) ( 0 .7 4 ) (0.25) -
Lime Treated - ( 2 .5 0 )

200 325 700 500 650 300 3,195


S-6C - 440 80 - (1.96) 8 .9
(0.17) (0.29) (0.33) (0)33 ) (0.26) (0.17) -
N a t lv e ( 0 .4 4 ) ( 0 .1 3 )

70 280 0 372 70 135 1,302


S-7A <50 100 105 - 170 7 .8
(0.36) ( 1 .1 7 ) (0) ( 1 .2 2 ) (0.39) (5.79)
Base (0 .1 1 ) ( 1 .0 0 ) (0 .1 7 ) (1.48)

400 24,500 0 3,230 550 1,600 40,355


S-7B1 <50 10,500 75 1,100 7.8
(0.20) ( 9 .9 9 ) (0) ( 1 .0 6 ) (0.31) (23.12)
Lime Treated ( 0 .1 1 ) (1 0 . 4 8 ) ( 0 .1 2 ) - (0.96)

800 2,750 0 3,720 700 1,400 10,020


S-7C <50 600 50 1,400 8 .0
(0.41) ( 1 .1 5 ) (0) ( 1 .2 2 ) (0.39) ( 5 .0 7 )
N a t iv e ( 0 .1 1 ) ( 0 .6 0 ) ( 0 .0 8 ) - ( 1 .2 2 )

750 11,500 0 1,650 1,000 - 20,600


S-8B 3,800 200 1,700 7 .8
(0.38) ( 4 .7 9 ) (0) ( 0 .6 6 ) (0.56) (11,99)
Lime Treated - ( 3 .7 9 ) ( 0 .3 3 ) - ( 1 .4 8 )

1,050 6,000 0 1,150 450 “ 11,280


S-8C - 1,150 180 - 1,300 8.1
(0.54) ( 2 .5 0 ) (0) ( 0 .4 6 ) (0.25) ( 6 .9 3 )
N a t iv e ( 1 .1 5 ) ( 0 .3 0 ) (1.13)

1,150 58,500 0 550 700 “ 84,500


S-9A - 21,000 950 - 1,650 7.0
(0.59) (2 4. 3 5) (0) (0 .1 8 ) (0.39) (4 9 .4 7 )
(2 0 .9 6 ) ( 1 .6 5 ) ( 1 .4 4 )
N a t iv e

650 42,000 0 600 250 60,850


S-9B 16,000 700 650 7.3
- (1 7. 4 9) (0) ( 0 .2 0 ) (0.14) - (3 5 .8 5 )
(1 5. 9 7) ( 1 .1 5 ) - ( 0 .5 7 ) (0.33)
N at iv e

950 6,500 0 1,300 550 - 11,355


S-9C 1,200 205 - 1,150 ( 6 .4 7 ) 7.9
- (0.48) ( 2 .7 1 ) (0) ( 0 .4 3 ) (0.31)
N a t iv e ( 1 .2 0 ) ( 0 .3 4 ) ( 1 .0 0 )

• E x c l u d e s M 3\ Mn2 + , S I 0 2

'• L i m e T r e a t e d and N a t i v e S o l i s e x t r a c t e d a t 50:1 d ilu tio n ra tio . Base Samp le e x t r a c t e d 5:1 d i l u t i o n ra tio ,
Total
S o lu b le S o lu tion
pH

N>
cr S107 S a lts * •

1
K+ HCOV
soa2- .

0
Na +

0
Mq2+ Mn2+

1
Ca2+
Mn/1000 am S o i l * * (ml 1 1l e q u 1v a le n ts /lIte r)
700 1,150 39,415
9,500 4 50 - 1,300 650 24, 500 0 2,315
(22.82)
8.2
S-10B <50 (0.33) (9.99) (0) ( 0 .7 6 ) (0.39)
( 9 .4 8 ) ( 0 .7 4 ) ( 1 .1 3 )
Lime Treated « 0.1 1)

3,700 650 1,500 11,950


<50 900 375 1,400 375 3,0 50 0
(0.37) (5.85) 8.0
S-10C (1.22) (0.19) (1.27) (0) ( 1 .2 8 )
N a t iv e (< 0 .1 1 ) ( 0 .9 0 ) ( 0 .6 2 ) -

800 1,450 60,400


2,2 00 1,350 40, 500 0 8 .3
S-11B 14,000 100 - ( 0 .3 2 ) (0.82) - (34.73)
: (1.91 ) (0.69) (16.86) (0)
Lime Tre ate d (1 3. 9 7) ( 0 .1 6 )

700 450 51,075


1,000 1 ,050 34, 500 0 (29.91) 7.3
S-11C 13,000 375 (14.37) (0) ( 0 .2 8 ) (0.26)
_ (1 2. 9 7) ( 0 .6 2 ) - (0.87) (0.54)
N a t iv e

2,4 00 1,000 37, 500 0 1,700 850 — 55,950


7.5
S-12A 11,500 1,000 - (15.62) (0) (0 .68 ) ( 0 .4 8 ) (31.01)
N a t iv e
_ (1 1. 4 8) ( 0 .1 6 ) ( 2 .0 8 ) (0.51)

1,850 500 51,750


1,250 750 24, 500 0 7.2
S-12B - 12,000 900 ( 0 .6 1 ) ( 0 .2 8 ) (30.19)
- (1.09) (0.38) (14.37) (0)
N a t iv e (1 1. 9 8) ( 1 .4 8 )

2,400 4 50 17,600
1,200 700 9,5 00 0 7 .5
S-12C - 2,600 750 (0) (0.7 9) ( 0 .2 5 ) - (11.01)
- (1.04) (0.36) (3.96)
N a t iv e ( 2 .5 9 ) ( 1 .2 3 )

650 500 — 4,2 75


400 225 1,750 0 (2.64) 8 .3
S-12F - 440 310 (0) ( 0 .2 1 ) ( 0 .2 8 )
( 0 .5 1 ) - (0.35) ( 0 .1 2 ) (0.73)
N a t iv e ( 0 .4 4 )

1,900 3,5 00 74,945


4,500 850 46,000 0 8.3
18,000 195 - ( 0 .7 6 ) ( 1 .9 7 ) - (44.51)
S-13B - (3.92) (0.43) (19.15) (0)
Lime Treated (17. 96) ( 0 .3 2 )

3,475 1,200 1,550 31,375


2,8 00 800 17,000 0 (1 8 . 1 4 ) 7.8
S-13C - 5,500 550 (0) ( 1 .1 4 ) ( 0 .6 8 )
( 0 .9 0 ) - ( 2 .4 4 ) (0.41) (7.08)
N a t iv e ( 5 .4 9 )

• E x c l u d e s A l 3 + , Mn2 + , S I 0 2
Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o
••Lim e Treated and N a t i v e S o i l s e x t r a c t e d a t 50 : 1 d i l u t i o n ra tio
Tota 1
Soluble S o lu tion
a i 3+ Ca2+ Mg2+ Mn2+ Na+ K+ SOi.2- CO,2- HC0^“ C l- S !0 ? S a lts* pH

Mq/1000 gm S o i l * * (ml 11I e q u l v a l e n t s / 1 I t e r )

9,500 40 1 ,0 0 0 475 23, 000 500 610 750 1,650 35,875


S-14B - _ (21.19) 8.7
( 9 .4 8 ) ( 0 .0 7 ) (0.87) (0.24) (9.58) ( 0 .3 3 ) ( 0 .2 0 ) ( 0 .4 2 )
Lime Tre ate d

195 2,4 50 900 43, 000 0 2,680 1,150 1,600 65,875


15,500 8 .3
( 0 .3 2 ) “ (2.13) (0.46) (17.91) (0) ( 0 .8 8 ) ( 0 .6 5 ) (37.82)
Lime Treated - (1 5. 4 7)

12,500 900 2,400 900 40, 500 0 855 1 ,1 0 0 1,550 59,155


(1 2. 4 8) ( 1 .4 8 ) (2.09) (0.46) (16.87) (0) ( 0 .2 8 ) ( 0 .6 2 ) (34.28)
N a t lv e -

120 2,150 700 48, 000 0 2,150 1,000 - 72,620


S-1 5B 18,500 8.1
( 0 .2 0 ) (1.80) ( 0 .3 6 ) (19.99) (0) ( 0 .8 5 ) ( 0 .5 6 ) (42.29)
Lime Treated - (1 8. 4 6) “

445 600 475 49,000 0 700 200 - 68,920


17,500 (39.74) 7 .0
N a t iv e (1 7. 4 6) ( 0 .7 3 ) “ (0.52) ( 0 .2 4 ) (2 0 . 4 0 ) (0) ( 0 .0 8 ) ( 0 .1 1 )
-

425 700 500 4,0 50 0 1,200 350 - 7,825


600 8.3
(0 .7 0 ) (0.61) (0.26) (1.69) (0) ( 0 .4 8 ) ( 0 .4 8 ) ( 4 .5 4 )
N a t iv e - ( 0 .6 0 ) “

20 3,900 1,150 24,000 1,600 0 1,400 1,700 39,070


S-168 7,000 (2 2 . 8 3 )
(6.9 9) ( 0 .0 3 ) ( 3 .3 9 ) ( 0 .5 9 ) ( 9 .9 7 ) ( 1 .0 7 ) 0 ( 0 .7 9 )
Lime Treated -

1,050 4,100 1,600 18,500 0 2,745 1,000 1,450 32,795


S-16C 3,800 (1 9 .0 7 ) 7.5
(3'. 79) ( 1 .7 3 ) - ( 3 .5 7 ) ( 6 .8 2 ) ( 7 .7 0 ) (0) (0 .90 ) ( 0 .5 6 )
N a t iv e -

1,950 450 6,500 0 2,745 750 1,450 14,170


S-1 60 1,050 700 ( 8 .1 7 ) 7. 5
(1.0 5) ( 1 .1 5 ) - ( 1 .7 0 ) ( 0 .2 4 ) (2.71) (0) (0 .90 ) (0 .4 2 )
N a t iv e -

900 175 15,500 0 1,700 450 24,925


S-17B - 6,000 200 - (1 4. 5 7) 7.9
( 5 .9 9 ) (0 .33 ) ( 0 .7 8 ) (0 .0 9 ) ( 6 .4 5 ) (0) (0 .68 ) ( 0 .2 5 ) -
Lime Treated

♦ Ex c lu d e s A l ^ + , Mn2+, SIO2
**Ll me Treated and N a t iv e S o i l s e x t r a c t e d a t 50:1 d i l u t i o n r a t i o . Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o .
Total
S oluble S olu tion
HCO-s- C l" SIO? S a lts * pH
Mq2+ Mn2+ Na+ K+ SO*2- C0x2*
a i 3+ Ca2+
Ma/1000 am S o l i * * (ml 11l e q u l v a l e n t s / l I t e r )

5,500 0 1,600 600 - 11,480


230 200 2 ,5 0 0 800 (6.58) 8 .5
S-17C - ( 0 .4 1 ) ( 2 .2 9 ) (0) ( 0 .6 4 ) (0.34)
( 0 .2 3 ) ( 0 .4 5 ) - (2.22)
N ative

0 330 70 150 1,081


10.5 <1 160 90 320 7.7
S-18A <50 100 (0) (1.32) (0.39) (6.07)
( 0 .1 7 ) (< 0 .0 1 ) (1.39) (0 .4 6 ) (1.33)
Base (< 0 .1 1 ) ( 1 .0 0 )

17,500 1,200 0 500 1,750 27,603


7,500 3 <1 700 200 (16.63) 9.5
S-18B <50 (7.29) ( 0 .8 0 ) 0 (0.28)
( <0 .1 1 ) ( 7 .4 9 ) ( 0 .0 5 ) (< 0 .0 1 ) (0.61) ( 0 .1 0 )
Lime Tre a te d

5,000 0 3,4 00 1,000 1,550 14,050


600 300 <1 2 ,9 0 0 850 (8.04) 8.3
S-18C <50 ( 2 .0 8 ) (0) ( 1 .3 6 ) ( 0 .5 6 )
( 0 .6 0 ) ( 0 .4 9 ) « 0.01) (2.52) ( 0 .4 3 )
N a t iv e ( < 0 . 11)

1,300 0 1,150 " 43,610


10 1,000 650 28,000 10.4
S-19B 11,500 ( 0 .6 5 ) (3 5 . 8 6 )
Lime Tre a te d _ (1 1. 4 8) (0 .0 2 ) - (0.87) ( 0 .3 3 ) (11.66) ( 0 .8 7 ) 0

5,000 0 75 55 7,2 05
1,850 35 150 85 (4 1 . 8 1 ) 7.3
S-19C ( 0 .3 1 ) -
B a c k f l 11 _ (1 8. 4 6) ( 0 .5 8 ) - (0.91) ( 0 .4 3 ) (2 0 . 8 2 ) (0) ( 0 .3 0 )

25,000 700 350 800 1,650 38,100


9,500 300 950 500 (2 2 . 5 3 ) 9 .0
S-213 <50 (1 0 . 4 1 ) ( 0 .4 7 ) (0 .1 4 ) ( 0 .4 5 )
( 9 .4 8 ) (0 .4 9 ) - (0.83) ( 0 .2 6 )
Lime Tre a te d (<0 .11 )

0 2,200 1,100 1,350 101,750


1,000 - 2 ,1 0 0 1,350 69,000 7 .5
S-21CN <50 25,000 (0) ( 0 .8 8 ) ( 0 .6 2 ) (59.30)
( 1 .6 0 ) (1.83) ( 0 .6 9 ) (28.73)
N a t iv e « 0.1 1) (2 4 . 9 5 )

1,000 50 350 23,560


10 - 650 900 16,000 9.4
S-22B - 4,600 ( 0 .6 7 ) ( 0 .0 2 ) ( 0 .2 0 ) - (13.15)
(0 .0 2 ) (0.57) ( 0 .4 6 ) ( 6 .6 2 )
Lime Tre a te d ( 4 .5 9 )

0 1,300 300 4,650


95 350 650 1,500 8.2
S-22C - 455 - ( 0 .5 2 ) ( 0 .1 7 ) - (2.55)
(0.30) ( 0 .3 3 ) ( 0 .6 2 ) (0)
N a t iv e ( 0 .4 5 ) ( 0 .1 6 )

• E x c l u d e s A l 3 + , Mn2 + , S I 0 2
Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o .
• • L i m e T r e a t e d and N a t i v e S o l i s extra cted a t 50 :1 d i l u t i o n ra tio .

Appendix 5 e
Tota 1
Soluble S o lu tion
S0a2 - HCOt;" c r S !0? S a lts* pH
A, 3+ Mg2+ Mn2+ Na+ K+ CO, 2 -
Ca2+

Mq/1000 qm S o l ) * * (ml I I l e q u l v a l e n t s / l I t e r )

2,750 0 210 160 4,021


160 16 - 550 175 7 .7
S-24A ( 1 .0 4 ) (0) (0.84) ( 0 .9 0 ) - (1 6 .7 0 )
- ( 7 .9 8 ) ( 0 .2 6 ) (4.79) ( 0 .8 9 )
Base

31,000 400 750 2,200 1,550 51,015


6,500 15 7 ,5 0 0 2,650 8.7
S-24B <50 (12. 91) ( 0 .2 7 ) ( 0 .3 0 ) ( 1 .2 4 ) (29.11)
( 6 .4 9 ) ( 0 .0 2 ) - (6.53) ( 1 .3 5 )
Lime Tre a te d {<0.11>

0 800 950 22,990


3,200 240 - 2 ,5 5 0 1,250 14,000 •7.9
S-24C - ( 5 .8 3 ) (0) ( 0 .1 2 ) ( 0 .5 4 ) - (1 3 . 0 1 )
( 3 .1 9 ) ( 0 .3 9 ) (2.22) ( 0 .7 2 )
N a t iv e

2,100 0 2,650 21,805


2,850 5 3 ,4 5 0 750 10,000 10.0
S-25B - ( 1 .4 0 ) 0 ( 1 .5 0 ) - (1 3 . 2 9 )
(2.8 4) ( 0 .0 1 ) - (3.00) ( 0 .3 8 ) (4.1 6)
Lime Tre ate d

0 2,1 50 2,350 - 25,310


3,200 410 2 ,9 5 0 1,250 13,000 7 .8
S-25C - (0) ( 0 .8 6 ) (1 .3 3 ) (1 4 . 6 7 )
( 0 .6 7 ) - ( 2 .5 7 ) ( 0 .6 4 ) ( 5 .4 1 )
N a t iv e (3 .19 )

200 600 600 850 21,480


4,400 30 1,100 700 13,000 8.8
S-26B <50 ( 5 .4 1 ) ( 0 .1 3 ) ( 0 .2 4 ) ( 0 .3 4 ) (1 1 . 8 8 )
( 4 .3 9 ) ( 0 .0 5 ) - ( 0 .9 6 ) ( 0 .3 6 )
Lime Tre ate d {<0 .11)

0 2,850 700 1,450 21,290


2,750 240 1,800 1,500 10,000 7 .9
S-26C <50 (0) ( 1 .1 4 ) ( 0 .3 9 ) (1 1 .1 6 )
( 2 .7 4 ) ( 0 .3 9 ) - (1.57) (0 .7 7 ) ( 4 .1 6 )
N a t iv e ( < 0 . 11)

0 1,150 500 1,600 39,675


25 1,400 1,100 26,500 8.2
S-27B <50 9,000 ( 0 .4 6 ) ( 0 .2 8 ) (2 2. 5 7)
( 0 .0 4 ) - ( 1 .2 2 ) ( 0 .5 6 ) (11.03) (0)
Lime Treated ( < 0 . 1 1) ( 8 .9 8 )

0 2,4 00 800 1,450 33,590


290 1,100 1,500 20,500 7.7
S-27C <50 7,000 - (0) ( 0 .9 6 ) ( 0 .4 5 ) (1 9 .1 5 )
( 0 .4 8 ) ( 0 .9 6 ) ( 0 .7 7 ) ( 8 .5 4 )
N a t iv e ( < 0 . 11) (6.9 9)

950 600 15,210 7.4


260 950 1,350 9,000 0
S-27D - 2,100 (0) ( 0 .3 8 ) ( 0 .1 7 ) - ( 8 .3 5 )
( 0 .4 3 ) ( 0 .8 3 ) (0 .3 8 ) (3.7 5)
N a t iv e (2.1 0) -

• E x c l u d e s A l 3 + , Mn2 + , S I 0 2
50:1 d i l u t i o n r a t i o . Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o .
••Lim e Treated and N a t i v e S o l i s e x t r a c t e d a t

240
Tota 1
S oluble S o lu tion

(N
CO,2 - HC0n“ C l" S107 S a lts * pH

in
Mq2+

O
K+

1
A|3+ Ca2+ Mn2+ Na+

Mq/1000 am S o i l * * (ml I 1l e q u l v a l e n t s / 1 I t e r )

700 9 ,2 0 0

o o
S-288 - 1,350 10 1,000 1,250 3,800 1,100 10.0
Lime Treated ( 1 .3 5 ) ( 0 .0 2 ) ( 0 .8 7 ) (0.64) ( 1 .5 8 ) ( 0 .7 3 ) ( 0 .3 9 ) -

200 _ 33, 550


S-28C 1,000 8,000 500 1,500 23,500 0 650
( 0 .1 6 ) ( 7 .9 8 ) - ( 0 .4 4 ) (6.77) ( 9 .7 9 ) (0) ( 0 .2 6 ) ( 0 .1 1 ) (19.51)
N a t iv e -

S-29B 2,200 20 . 2,600 550 8,500 200 1,950 550 - 16,570


( 2 .2 0 ) ( 0 .0 3 ) - ( 2 .2 6 ) ( 0 .2 8 ) ( 3 .5 4 ) ( 0 .1 3 ) ( 0 .7 9 ) ( 0 .3 1 ) (9 .5 4 )
Lime Treated -

70 9 135 19 145 0 330 50 _ 758


( 0 .7 0 ) (0 .1 5 ) - ( 1 .1 7 ) ( 0 .1 7 ) ( 0 .6 0 ) (0) ( 1 .0 8 ) ( 0 .2 8 ) (4.15)
Base -

<50 3,750 25 <1 1,800 600 11,000 200 1,800 750 1,500 19,925
S-30B
(3.74 ( 0 .0 4 ) (<0 .01 ) ( 1 .5 7 ) (0.31) ( 4 .5 8 ) ( 0 .1 3 ) ( 1 .5 7 ) ( 0 .4 2 ) (11.51)
Lime Treated (<0.11>

1,150 175 <1 1,350 900 3,200 0 3,400 1,650 1,400 13,225
<50 (6.69) 8 .0
N a t lv e «o.n> ( 1 .1 5 ) ( 0 .2 9 ) (<0 .01 ) ( 1 .1 7 ) ( 0 .4 6 ) ( 1 .3 3 ) (0) ( 1 .3 6 ) ( 0 .9 3 )

2,000 10 2,150 600 7,000 1,200 250 1,350 - 14,560


S-33B ( 8 .7 7 )
Lime Treated ( 2 .0 0 ) ( 0 .0 2 ) - (1 .8 7 ) (0.31) ( 2 .9 1 ) (0.8 0) ( 0 .1 0 ) ( 0 .7 6 )
"

240 5,500 1,450 12,000 0 2,750 1,700 1,300 24,590


S-33C <50 950 8.1
(4 .7 9 ) ( 0 .7 4 ) ( 5 .0 0 ) (0) ( 1 .1 0 ) ( 0 .9 6 ) (13.93)
N a t lv e «o.m ( 0 .9 5 ) ( 0 .3 9 ) ”

125 350 375 1,100 0 2,150 400 - 5,400


900 •8.2
( 0 .9 0 ) ( 0 .2 1 ) ( 0 .3 0 ) (0.19) ( 0 .4 6 ) (0) ( 0 .8 6 ) ( 0 .2 3 ) (3.15)
N a t iv e - **

•Ex clud es A ) I +, Mn2+, S102

••L im e Treated and N at iv e S o i l s e x t r a c t e d a t 50:1 d i l u t i o n r a t i o . Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o .

A p p en d ix 5g
Tota I
Sample Number Ca2+ Mq2+ Na+ So lub le S o lu t I o n
K+ SO.2' CO,2" HC0t “ C l"
Mq/1000 qm S o i l * (ml 111eq u 1v a 1e n t s / l 1t e r )
0-1BU 3,400 75 1,050 800 9,000 1,200 0 700 16,225
Lime Treated ( 3 .3 9 ) (0.1 2) ( 0 .9 1 ) (0.4 1) ( 3 .7 5 ) ( 0 .8 0 ) (0) (0 .39 ) (9.7 7) 9.7
(undamaqed)

0-1B 11,500 475 1,250 1,350 32,000 1,200 100 200 48,075
N ati ve (11.47) ( 0 .7 8 ) ( 1 .0 9 ) ( 0 .6 9 ) (13. 32) ( 0 .8 0 ) ( 0 .0 4 ) ( 0 .1 1 ) (27. 50) 9 .3
0-1C 310 205 2,000 1,350 4,300 800 1,400 600 10,965
Nat ive ( 0 .3 1 ) ( 0 .3 4 ) (1.7 4) ( 0 .6 9 ) ( 1 .7 9 ) ( 0 .5 3 ) ( 0 .5 6 ) (0.3 4) ( 6 .3 0 ) 9.3
0-2B 4,250 25 ’ 650 600 11,500 600 400 250 18,275
Lime Treated ( 4 .2 4 ) ( 0 .0 4 ) ( 0 .5 7 ) ( 0 .3 1 ) (4.7 9) ( .4 0 ) ( 0 .1 6 ) (0.1 4) 9.5
0-2C 15,000 750 1,000 750 45,000
Nat 1ve 0 700 350 63,550
(14.97) ( 1 .2 3 ) ( 0 .8 7 ) (0.3 8) (18.74 (0) * ( 0 .2 8 ) (0.2 0) (36. 67) 7.4
0-3B 8,000 50 600 425
Lime Treated 19,500 400 400 750 30,125
(7.9 8) (0.0 8) ( 0 .5 2 ) ( 0 .2 2 ) ( 8 .1 2 ) ( 0 .2 7 ) 9.1
(0 .1 6 ) (0.4 2) (17.77)
0-3C 3,700 550 1,000 800 13,000 0 800 250 20,100
N ati ve (3.6 9) ( 0 .9 0 ) (0 .87 ) (0.4 1) (5.4 1) (0) ( 0 .3 2 ) (0.14) (11.74) 7.9
0-4BU 5,500 75 4 50 350 13,500
Lime Treated 1,000 0 850 21,725
(5 .49 ) (0 .12 ) (0.39) (0.18) ( 5 .6 2 ) (0.6 7) 9.9
. (undamaqed) (0) (0.4 8) (12.95)

0-43 7,000 55 700 4 50 17,000


Lime Treated 1,000 300 600 26,795
(6.99) ( 0 .0 9 ) ( 0 .6 1 ) (0.2 3) ( 7 .0 8 ) ( 0 .1 2 ) ( 0 .1 2 ) (0.34) 9.3
0-4C 10,000 750 3,100 1,200 37,000
Nat ive • 0 650 750 53,450
(9.98) (1.2 3) (2.7 0) (0.61) (15.41) (0) 7.9
( 0 .2 6 ) (0.42) (30.61)
0-6BU 270 5 950 650 6,500
Lime Treated 1,500 0 300 12,605
(2.6 9) ( 0 .0 1 ) (0.8 3) (0.33) ( 2 .7 1 ) ( 1 .0 0 ) 10.2
(undamaqed) (0) (0.1 7) (7 .74 )

0-6B 3,550 10 250 325 8,500 0 1,450 250 14,335


Lime Treated (3.54) (0.0 2) (0 .22 ) (0.1 7) ( 3 .5 4 ) (0) 7.7
( 0 .5 8 ) (0 .11 ) (8.2 1)
0-6C 7,500 55 2,150 1,600 26,500 0 750 200 38,705
Nat 1ve (7.49) ( 0 .9 0 ) 7.8
(1.8 7) (0.8 2) (11.03) (0) ( 0 .2 8 ) (0.1 1) (22.50)
•Lime Treated and N ati ve S o il s e x tr a c te d a t 50:1 d i l u t i o n r a t i o . Base Sample e x t r a c t e d 5:1 d i l u t i o n r a t i o .

Appendix 5h

242
Appendix 7a
X -R A Y DIFFRACTIO N SCAN
Sample: S-14B - Damaged Lime Treated Subbase; Mineral Aggregate
P olyhalite, Gypsum

Gypsum

Thaumasite
Thaumasite
Thaumasite
X -R A Y D IFFRA CTIO N SCAN
Sample: S-19B - Damaged Lime Treated Subbase; Single Crystal
X -R A Y D IFFR A CTIO N SCAN
Sample: S-30B - Undamaged Lime Treated Subbase; Mineral Aggregate
Appendix 7 d
X - R A Y D IF F R A C T IO N S C A N
Sample: 0-1B - Damaged Lime Treated Subbase; Mineral Agg re g at e
Appendix 7e
&K 4 M : SS5. ^ 1 « M 19811

Energy dispersal/scanning electron the lack


Th auma sit e, Ettringite and a mixed c r y s t a l s t r u b e l , 1981);
of a l u m i n u m in p u r e T h a u m a s i t e ( f > o m t o l l m a n

249
250

Ca

Sample : S-8B
Damaged Lime T re a te d Subbase
M in e r a l A ggregate

Appendix 9a Typical energy dispersal patterns from the


scanning electron microsope for Stewart Ave.
251

Sample: 0-1B
Damaged Lime Treated Subbase
M in e ra l Aggregate

Appendix 9b Typical energy dispersal patterns from the


scanning electron microsope for Owens Ave.
252
253
R8jfy?
Mr. Dal Hunter
March 8, 1988
Page 2

If you see any ways in which this program might be strengthened to


highlight the key factors (without running the number of test> out of
sight!), we would certainly appreciate it. Are there any main factors
that have been overlooked?

I look forward to hearing from you and also learning more about
your own research on this fascinating, but difficult problem.

Sincerely yours,

James K. Mitchell
Professor of
Civil Engineering

JKM/nh

A p p e n d i x 11b D e s c r i p t i o n ot l i m e / h e a v e e x p e r i m e n t s
in p r o g r e s s at U.C. Berkeley.
255
2 5 - 1 2 8 --------------------------------
3.41 9.56 [Ca3Si(Ol06-12H20](S0O(C03)
d 9.56 5.51

20 100 Calcium Carbonate Silicate Thaumasite


100 40
i/ i Sulfate Hydrate
dA 1/1 hlcl d A r/i hkl
Rad. CoKajX 1.7890 Filter Dia. 114.8mm 100 100 2.357 4 312
Cutoff I/I j Microphotometer 1/1 cor. 2 101 2.282 <1 204
Ref. Knill, Mineral. Mag., 32^416 (1960) 40 110 2.191 6 320
5 111 2.155 13b 402,223-t-
5 200 2.106 5 313
S.G. P63/m (176)
Sys. Hexagonal
co 10.40 A C 6 102 2.086 2 410
a<) 11.03 b0 Z 2 E)x [1.83] 201 2.045 4 411
4
a P y 3 322
16 112 2.019
Ref. Ibid. 210 1.934 3 412
1
6 202 1.911 10 500
Sign -
(a 1.470 n to/3 1.504 £y
Color Colorless 20 211 1.809 3 331
2V D 1.91 rap 16 300 1.778 3 421
Ref. Ibid. <1 301 1.733 3 332,006
<1 113 1.692 2 511
Specimen from Ballyalton, noruiei.. 2 220 1.626 4 414
Validated by calculated pattern 24-28. 14 302 1.592 4 600
To replace 13-156. 4 310 1.570 <1 430
<1 004 1.551 1 431
10 311 1.538 1 513
213 1.467 1 226,107
>.499 } 10

Appendix 1 2 X - r a y diffraction p e a k s for thaumasite.

to
<ji
U\

/'
m n tm m ....... ...........................M l.............II
Ca6Al2(SOH)3Ca0i2*25H2O
Calcium Aluminum Sulfate
Hydroxide Hydrate

Rad. CuK0l x 1.5405 Fillet Ni Dia. 2.616


Cui off 1/1, Diffractometer l/ l c o r. 2.564
Ref. Nat. Bur. Standards Circ. 539, £ 3-4 (1959) 2.524
2.487
S.G. Pbj/mmc (194) 2.434
5. Hexagonal
co 21.44 A c: 1.9092 2.422
11.23 b0
P y z : Dxl.754 2.401
2.347
Ilcf. Ibid.
2.230
2.209
n oj fi lY Sign 2.185
2V D ra p Color Colorless 2.154
Ref. Ibid. 2.130
2.124
Sample prepared at NBS. Spcct. anal, showed 0.1-1.0\ 2.081
Si, Sr; 0.01-0.lb Ag, Cr, Cu, Mr., No, Pb; 0.001-0.01% 2.062
B, On, Cs, Fc, Mn, Ni, Sn, 2n, Zr. 2.033
Pattern made at 25*C. 2.027
1.979
1.975

9 - H1 A
9.73 5.61 3.E 9.73 C a 6A l 2 (S 0 l| ) 3 ( 0 H ) l 2 * 2 5 ll20

100 Calcium Aluminum Sulfate
1/1 100 80 50
H y d r o x id e tty a rn
fEttrincitc)
d A 1/11 hkl d A 1/11
Filter L ) ia .
Had. A 1.946
Cut o f i/i i l/l c o r.
1.905
Ref. 1.875
1.853
S.G 1.845
Sys. C
co
A 1.829
ao bo Dx
y Z 1.812
n P 1.809
Ref.
1.786
1.768
Sign
n oj /3 y
D rap Color
2V
Ref.

See preceding card

Appendix 1 3 X - r a y diffraction p e a k s for ettringite.


259
261
262
diffraction pattern for sample
265
267
Appendix 22a WATEQDR ion activity calculations tor gypsum, samp---------- .
Z. ZZZZZZe-- ZZ Z.zZZZZZe*ZZ
0. 3283 ® 4 e -8 5
0.0000006+0® Z. 000000e-00
0 . lS4395e-05 0. 000000e-r00 z.ZZZZZZs+aZ
Z. 1 0 6775e-®5 0.000000e+00 a.zazzaze+zz
0.0000006+00 0 . i 5 *S 3 0 e — 09
0.6 l 96 l 4 e -0 6
0.000®00e+0®
0.35932£e-®6 0.0000806+0® „>j0 .0000006 +0® TST^ElEhSH^TUT^-
__^
0. 2083 6 7 e -0 6 hi
' 0 . ic;wJ6i9a-v5tr 0.0800006-00 0.000000e+0®
'0.'000000e-^0®i> jX•0i 3 a A 3 5 £ e - l 0
'0.696£S6e+07 0. 0.000006-*-®® a~*'^ '
0. 4©'*436b -07 -2rT7nzr?000e-<-02r

earnest
eomarj
•;:c,’ 'temperature 0.000000B+00
0. 1000006-^03
0. i00000e-*-03
0. 1000006 •*■03

tTS^zT

OH
CaOH

■K c v 5 u lT
Na£S 0 4
N aH C 03

‘ A&<KS 04
3 95 -
:k c i -..
6 3 HSOA

sample G -2 5
Appendix 22b WATEQDR ion activity calculations for gypsum
1water
r.x - W ' *; ■ :»'•; O - ' Y . #
W *■*-•-!•. ‘ -i* t •\V*v »it
ip .-.ro -.•••;-•V ■
t
• •n’
: ; - f? IP Ip p fp P
E--' •• •■•
,
_______• " '
- -------------------------- -
; " P : ll£
ir.it ;a. so* vil

P 3 I 1 1
- — — - — ---------------------- — y
V ^ n ^ t u r . I -8 ^ c.B r . ^ - C , f1
3. :
Ana Lytica. e3ncatfc =_ a fe^.3 Cr.alytical e rnian a a
-4. fc>u.s y
— “ — - ■ p -
« ♦* r e o o x - maui. «**
•. • ,• .-■ • .. . ..

'6 ■, oxy cert * 0.000 mg/ - __ T

l g § 1| J P ... 1 S § P § § § If f l I B H i l i
I
*** total co--cer.t rations cJ irout ssecies
a r * ’ " ’l 1 ” ” .. V i.-, a y --* - •., - Y .j
v ♦;» • .•-*£/. lo g t o t a l to
to ta l
m o la lit y mg/
rno 1 a 1 1 1 y
I
-£.e:33 s : . 6®00
3. 074 1?. 966 ca
0.1537£4e-0£ - 7 , 1 0 0 7 7 ”- - 0.
mg
- Y Y F f
“ iT T 55. i? .6 5 : ■ . '• 0 . 9 6 6 "•" c a + « 3 -Y
- 4 . :354 1 .5 0 0 0 0 ■
■ •••.•*. < n * j |
;'® . 6 5 £ 5 9 3 e - ® 4
1.7700® 0 . 044 • 0. ® 2l

IMMHHHMHHMHHHHHHHIji
®. 4 3 4 6 5 0 e - ® 4 —4 . 3 6 1 7 777T 5
i2i. 5 c,4 c.-ti^ e - i ;i-— •_. u Z .Z 7 . 0 34 r ,a * x
77.0002
SD* -1 i.. 72665I.-H2 0. 00.00 0.000 0.017 cl_ _ _ _
-cas -1 ____* • _ 570.0 0 0 ”
■ 0It.. 0 00 , . • 0
Mil
• . a '. e. ^e mi 1'
. 4 £ 7. .1-.. s o 4
J.-K-y.
* . flt-TlJ!
. r%
.• _

e a a ,•' f c o s * c o 3 ...
.'F-iif;}'!'
. •<•• ? : > . «
Biu£ tot wS- r T” , . *'• '■ '. 0 .0 0 0 0 ' :' e . * e a - a .* * 3 :---------:------------—
; - ■ !..■ • p a*, \< v -3 ■ 0.0 000
” z 77T 7 0 0
“ 0 . 9 d 004Q«—05
f
0.. ^ - 6 l S l . e - 2 3
:3.0000
“ 5 7 6 . t f f 3T

cor^vergence iterations «♦*

0 . : S 3696e - 0£ 7 . 6 6 7 3 0 7e 04
0. 00.0000^07
0;0.000Jie*0-: 0.0000000
ee-*-00
00: -i» ■ •
iaS'; _l 3 6 9 8 e -■«-»
-f?-' I.V6iTi*PA6e-03 03 - «
0 .- 3
TM1?
7.«3r.4
4 l3e®-04'
.. 0. 0.0.0000e-t-0® 0.000. 0e7.
-• V 5 - % : i : V t V ■* ' k r '

:0. 2 ® 2 0 71 a-0-4_
0. 000000e+-0‘0 0.000^00S’-v'^"” 7 . 3 6 toC0- ^ e -
-...■ a :':.:':.' > ■ -■■ ■ 0 . 0 0 0 0 0 /e '
0 . a S 0 8 £ c :e - © 3 - 0 . : 55560e - 0 ^
. 7607396-05 0..000000e7-00 0.000000S~0®
0. 00000'0a-n?,?i
•0. 60.£ ^ -0 7 <p-
0 . £ 38609® -
0.. g T A a ? ^ © - © ^ 0 . 0 0 .0 0 0 0 © * 0 0
0. 5£3SS6e7®4 3397e -

K)
* „ n ,n r i.« P3a W ATEQDR ion a c tiv ity ca lcu la tio n s fo r ettrin g ite , sam ple E T -2 5 C . 0>
VO

M m n n w n i—
7. 7 00.0*00«t— 07 0., 00-00.00 e -®0 0. £ 36f. ? f
0. 26631 7 e-05
•0. 0 0 00 0 0 C +0 0 0. 0 0 00 0 0 0 -0 0 -0. iw-l-O .
0 . 1 = 4 6 1 le-05
0 . e i l 6 S 4 e -05 0 . 6 6 9* 0 5 e -0 6 0. 00000.05+0-0 *. 000.000e-00
* . 5 * 71 6 2 e -0 6 0.00**00e-00 0 . 00 0 0 0 00 - 0 * 0. 23360:
' 0 . 4 3 59 6 7 e -0 5
0>. £674£fce—*6 0. 000.000 e + 0.0 0 . 00 0 0 0 00 - 0 0 - 0 .1 2 5 . 6
0.23422fce-05
0 . i6:99le-0S 0 . 0 0 00 0 0 e +0 * 0 . 00 0 0 0 00 - 0 0 - 0 . ic-sie'
13 0 . l £5870e-*5
•t - ;••.*+ li
14 0;67&6l0e-06 . 0 . 90 6 6 l la - * 7 0 . 00000tfe+00 . .0. 0 0*000e+00, ;• 0. 2 3 6 6 0 ^ e - l l : •
.‘ . '• J’X'/TVLL
... , ■- -l j:

- h

■ • '
■" : ' ■■■ 1V’: T-V hi

... ♦ c es c r : pt l o n of s o lutior **** F»



FT
EI ...... analytical comoutec • ■ on . ;T. :.*rV'i^K:
’U
.... . < eprncat -4.629 -2.555 ; 10.600 ' •T ' oc o£ =
ioa oco2 =
0. 14 9 2 4 2 e - 0 6
— 6 .8 2 6 1
V : -.

•/. e r ro r 15.1*3 4.546 te mp e r a tu r e 302 — 0 . 0 * 0 0 0 0 0 — 00


dci4 = 0 . 00 0 0 0 00 + 0 0
en = *■+•**** oe = 1 0 *. 00 * 25. 0* ceg c
coc tot = 0 . 5 4 2 7 2 0 0 —03
oe ca lc s = 0 . 1 00000e-03
d o -calc' d o x ? 0 . I®000i?e+03 ... ionic- s t renptn , •' • c e ns x t y 1-****. ‘ •* ^ •* f K
'• V & f e f 0.-47615Se-02' ■ ec -. .'0 . 0000000 + 0 0 ;^
^'SiiSSVV
pe s a te d o x B 0. l * 0000e+03
i •; ■ rt.-- •i’ ;• .v.-!.-.-. riVVl’ ,.'iS1 •• ■-

:
\ 'P a g e 1,
’ c i st r i p ut i o n of s o ec x e s
■fey
I species
i species

-2. 9140
a c ti v i t v

0,91£35e-83 ‘ “ 3. 0 396 “;-


act. coeff.

•0. 74S4£e+00r-^:~ - 0 . i£5T" :Na


-c»v
V fca" a. 4 6 & 4 9 e + 0 2 ■
:f 0 . l2i900-02.
0.565£9e-0A. -4 l £326. 0. 92 90 5 e + 00 -0.0320
-4.£007
4K
3 Na 0 . 14430e+01’ C. 62 9 9 9 e- 0 4
0.402070-04 -4. 3957 0. 9 2 7 7 1 e+ 00 "• -0.0326 K
0. l6543e+®l_ 0>. 43340-e— 04 -4.3631
- 1 0.5676 0..£51 19b -10 0 . 9 26 4 4 e+*® -0.0 3 2 2
0. £ 7£66e-07 0 . £/05 5 e - 10 -0.0326
5 Cl 0.£0000e+0i 0 . 5 £ 4£4e-0 4
0 . 6 3 9 8 9 e - 03
-4.£465
-3. 1339
0 .5£34£e-04
0 . 4 7 7 8 6 ° —03
0 . 9 6 7 7 . e+00
0.. 74674©+00 —0.1268
6 S04 0 . 6 l 456e+0£ ~471£2L5g ~ 0.9 3 0 £ 4 e + 0 0 -=i?rff3T7r-
0. 5’ 9 407e+01' 0 • 9 7 2 8 le— 04 -4.0 1 1 5 “ 0. 9 0 5 6 6 e - 04
7 MC 03 0 . 74SS 3 e + ®0 - -0. 1256
0 . 2 £ 5 4 9 e — 03 -3.6 4 6 9 . 0. 16 8 8 5 b — 03 -3. 77 25 •.
13 .C03 . 0 . l35£9e+02 0. 00 0 5
0.. 5 1 0 6 4 O - 0Q - 8 .2 9 1 9 0 . 100l£e+01
0 . 3l&23e-0.3 0.5l005e-06 -a. 2 9 2 4 _
86 H 2 C0 3
-3.3970 0. 3 2 o 4 4 e + 0® - 0 .0322
-1 0. 734i2e-i ”0*. 43 l7 4 e - 03 — 3.3646
£7 0-1 0. 9264 4 0 - 00 -0.0322
0.9 6 3 4 £ e - 05 - 5 . 004& 0 . 9 1 6 6 2e-®5 -5.0369
29 CaOr; 1
0
0 . 56472s
0 . I2075e- 0 .8 6 7 1 3 5 -0 4 - 4 .0520 0 . 6 6 6 1 1 e-04 — 4.0515 0.1001le+0'- 0. 0 0 0 5
ns 32 C aS04 ac 0. 92&4<+o+00 -0:83£cr
0 . 9 2 2 5 9 e — 06 ' - 6 .0350 0.. 8 5657 i -06 -6.0 6 7 2
Uj 30 CaHC03 1 0 . 93252o
-3.6602 0 . 2 1 8 9 2 e —03 -3.6597 0. i0 0 1 1 e + 0 1 0. 00 05
0. 2 l S 6 8 e - 03
- 31 CaC03 aa 0 0 . 21863e
0. 92644e-*-00 - 0 .0322
i
U6
44
£9
N aS04
Na2S04
-i
0
0.17971e
0 . 75492e— 05 0 . 5 3 1 5 9 e - 10 -10. £744 0. E,321 7 e - 10
0. i•-98l£e-0 6
-6.8 5 3 3
-10.2739
-6. 5256
0.100110+01
0.l001le-01
0.0005
0.0005
A3 Na nC G S 0 0 . 2 5 0 10e 0 . 2 5 7 6 3 e —06 -6. 5260
0. 926440-^?^ —0. 0322
— 01 0.. 197300-06 -6.70+9 0 . :,63l8e-05 — 6.7371
-7.3331 0. 1 0 0 1 le+01' 0 .0 0 0 5
*-02 0 . 46 36 6 e - 07 -7.3336 0. '
49 Na 2C 0 3 -10.1157 0 . 1 0 0 1 le+01 0. 0005 . NCC1 , £5
0 . 7 6 5 2 0 e — 10 -1 0. 1 1 6 2 0. 7 6 6 0 4 © - ;0
v KS C4 ~K
-
94 N O C 1.
46 KS04 -1 0. I 9 6 5 3 o - 0 1 0 .14542e _8373_ e. l3 5 0 2 e - 0 6 __
0 . 5 4 7 £ 4 © — •0
- 6 . e696
— 10. 2616
0. 92844©+0.0
~0~ 10 0 1 1e - 0 1
-0 .0 3 2 2
0. 0005 <C1
r,r 95 KC1 0
-1
0. 4 0 7 4 6 e - 0 5
*. l £ l 7 9 e - 0 6
0. 546&4e-
0.12549e
— 1*0. 2623
-11.9014 0 . 1 1 6 5 . e - 11 - 1 1.9336 0.92644©-f00 -0. * 3 2 2 -504
63 HS 0 4 01. 92 34 4O+0.0 -0. 0322 Pi(0-> c
0. 2 4 1 5te- - 1 1.6170 0. £ 2 4 £ 7 e-l 1 y>
0 .14731e-06
p

b
M
53
. ’ v :;iv;; 5 4
Pi <0-0 2 1
'PI (O h ) 4 - “ I .
0 . 4 5 7 770+02,
•••%■. •••; ; - p t i'i'.
0 . 46l9le -3. 3170 0 . 4 4 7 4 3 e —03 -3.3 4 9 3

e.i = 5 . 91 6
0 . 9284 4 © + 00

volts’ '

-0.0322
■ ..v,■
>4 ,
Pi <On) 4.; I

ecu ivalent
F cornou ino t i e C i s t r i p u t i on of so ec les.

You might also like