You are on page 1of 314

Expansion and Shrinkage of Early Age

Cementitious Materials Under Saturated


Conditions: The Role of Colloidal Eigenstresses A
by MASSkCHU TsINSTITUTE
Of TECHNOLOgY
Muhannad Abuhaikal
JUN 0 7 2016
B.S., Birzeit University (2007)
S.M., Massachusetts Institute of Technology (2011) LIBRARIES

Submitted to the Department of Civil and Environmental Engineering


in partial fulfillment of the requirements for the degree of

Ph.D. in Civil Engineering

at the

MASSACHUSETTS INSTITUTE OF TECHNOLOGY

June 2016

Massachusetts Institute of Technology 2016. All rights reserved.

I
Author ......
Signature redacted
D rtment of Cvil and Environmental Engineering
April 15, 2016
Certified by. Signature redacted 21,
.

Franz-Josef Ulh4"f
Professor of Civil and Environmental Enginee i'
Thesip Supervisor

Accepted by .. Signature redacted


( frieidi M. Nepf
Donald and Martha Harleman Professor of Ci vil and nvironmental
Engineering
2
Expansion and Shrinkage of Early Age Cementitious
Materials Under Saturated Conditions: The Role of
Colloidal Eigenstresses
by
Muhannad Abuhaikal

Submitted to the Department of Civil and Environmental Engineering


on April 15, 2016, in partial fulfillment of the
requirements for the degree of
Ph.D. in Civil Engineering

Abstract
Mixing water with anhydrous cement powder and other additives results in a viscid
cement slurry and triggers a set of complex exothermic reactions. As the cement
slurry transitions from a suspension to a gel and ultimately to a stone-like porous
solid, the material develops mechanical properties. This transition, however, is also
accompanied by bulk volume changes, which -if restrained- lead to premature crack-
ing of materials and structures. The main objective of this study is to relate bulk
volume changes as measured at a macroscopic scale to their finer colloidal origin under
controlled temperature and pressure conditions. To achieve this goal, an original set
of macroscopic scale experiments is designed and a multiscale microporomechanics
model is employed to rationalize the experimental results.
While bulk volume changes have been classically attributed to capillary pressure,
surface tension, and disjoining pressure that all relate to changes in relative humidity,
we herein argue that they are a consequence of eigenstresses that develop in the
solid phase of the hydrating matter due to attractive and repulsive colloidal forces at
mesoscale. To prove our hypothesis, we experimentally investigate volume changes
under saturated and drained conditions that eliminate any volume changes associated
with humidity and (effective) pressure changes. Under these conditions, we observe
first a volume expansion followed at later stages of hydration by volume shrinkage
that cannot be explained by classical theories.
By analyzing both expansion and shrinkage within the framework of incremen-
tal micro-poro-mechanics, we suggest that the expansion is caused by the relative
volume change between the reactive solids and hydration products in the hydration
reaction. After the solid percolates, this volume change is restrained by the per-
colated solid phase. This induces a compressive eigenstress in the solid phase that
entails a swelling of the material under overall stress-free conditions. In return, as the
material further densifies, the attractive forces between charged C-S-H grains prevail
causing the whole system to shrink. These attractive (tensile) forces compete with

3
the compressive solid eigenstress development, reversing the expansion into shrinkage.
By carrying out tests at different temperatures, we provide strong experimental evi-
dence that this tensile eigenstress development is an out-of-equilibrium phenomenon
that occurs close to jamming. Furthermore, the tensile eigenstresses calculated from
our shrinkage measurements agree qualitatively with those from meso-scale coarse-
grained simulations of C-S-H precipitation originating from the electrostatic coupling
between charged C-S-H particles mediated by the electrolyte pore solution.

Thesis Supervisor: Franz-Josef Ulm


Title: Professor of Civil and Environmental Engineering

4
Acknowledgments

I would like to express my gratitude to my advisor Prof. Franz-Josef Ulm for his

help and guidance throughout my masters and PhD studies and especially, I would

like to thank him for his patience. I would also like to thank Dr. John Germaine

for the great help in designing the main experimental setup in this study, Benjamin

Druecke, James Haug, Ivan Prestini, Tom Bell, Konrad Krakowiak, and Miranda

Amarante for the great help in the lab. I am also thankful to Enrico Masoero, Jeffrey

Thomas, Roland Pellenq, Elizabeth Dussan, and Henri Van Damme for the insightful

discussions. My thoughts go to my friends and colleagues who all made this a more

enjoyable experience and special thanks to Donna Hudson.

I gratefully acknowledge that this research was made possible by the funding from

Schlumberger as part of the collaboration on the X-Cem project with Schlumberger-

Doll Research Center.

The greatest thanks go to my parents and my siblings.

5
6
Contents

I Overview 12

1 Introduction 13
1.1 Industrial context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

1.2 Research question . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

1.3 M ethodology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
1.4 Outline of thesis. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17

2 Literature Review 19

2.1 Measurements of bulk volume changes . . . . . . . . . . . . . . . . . 19

2.2 Driving forces and models of bulk volume changes . . . . . . . . . . . 21

2.2.1 Capillary pressure . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.2.2 Disjoining/Joining pressure . . . . . . . . . . . . . . . . . . . 30


2.2.3 Surface energy . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

2.2.4 Crystallization pressure . . . . . . . . . . . . . . . . . . . . . . 32

2.3 Mitigation strategies of bulk volume changes . . . . . . . . . . . . . . 33

2.4 Colloidal forces in C-S-H gel . . . . . . . . . . . . . . . . . . . . . . . 35


2.5 Sum mary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

II Materials and Methods 38

3 Materials 39
3.1 Light-burn magnesium oxide . . . . . . . . . . . . . . . . . . . . . . . 39

3.2 C 2 S and C 3 S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

7
3.3 Class G cem ent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

.
3.4 Sum mary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

.
4 Experimental Setup 47

4.1 General Layout . . . . . . .. . . . . . . . . . . . . 49

.
4.1.1 Pressure cell.. . . . . . . . . . . . . . . . . . . . . . . . . . . 49

.
4.1.2 Syringe pumps . . . . . . . . . . . . . . . . . . . . . . . . . 50

.
4.1.3 Bladder accumulator . . . . . . . . . . . . . . . . . . . . . . 52

.
4.1.4 Flow meters . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

.
4.1.5 Check valve . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

.
4.1.6 F ilter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

.
4.1.7 Water bath . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

.
4.2 Configurations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

.
4.2.1 Permeability and Volume of the reaction Configuration (PVC) 58
4.2.2 Rubber Diaphragm Configuration (RDC) . . . . . . . . . . . 58

.
4.2.3 Liquid Barrier Configuration (LBC) . . . . . . . . . . . . . . 58

.
4.3 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

5 Analytical Framework and Boundary Value Problems .


5.1 Microporomechanics framework . . . . . . . . . . . . . . . . . . . .
5.1.1 Poroviscoelasticity . . . . . . . . . . . . . . . . . . . . . . . .
.

5.1.2 Useful relations and tools . . . . . . . . . . . . . . . . . . .


.

5.2 The Permeability and Volume of the reaction Configuration (PVC)


.

5.2.1 Stress and displacement boundary conditions . . . . . . . . .


.

5.2.2 Volume of the reaction Vrxn (chemical shrinkage) . . . . . .


.

5.2.3 PVC test with bottom surface connected to the pressure trans-
ducer only . . . . . . . . . . . . . . . . . . . . . . . . . . . .
.

5.3 The Rubber Diaphragm Configuration (RDC) . . . . . . . . . . . .


.

5.3.1 Boundary conditions . . . . . . . . . . . . . . . . . . . . . .


.

5.3.2 Field equations . . . . . . . . . . . . . . . . . . . . . . . . .


.

5.4 The Liquid Barrier Configuration (LBC) . . . . . . . . . . . . . . .


.

8
5.4.1 Boundary conditions . . . . . . . . . . . . . . . . . . . . . . . 96
5.4.2 Wetting, interfacial tension and viscous forces . . . . . . . . . 97
5.4.3 State of stress and strain in the liquid barrier test . . . . . . . 100
5.5 Sum mary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106

III Results 110

6 Results 111
6.1 Magnesium Oxide.. ......... .. ...... . ... .. .. ... . . 111
6.2 M onoclinic C3 S . . . . .. . . . . . . . . . . . .. . . . .. .. . 119
.125 6.3 f-C 2 S . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
6.4 Class G cement ... . .. .. .. . ...... . . . .......... .. . 128
6.4.1 Hydration kinetics . . . . . . . . . . . . . . . . . . . . . . . . 128
6.4.2 Bulk volume changes . . . . . . . . . . . . . . . . . . . . . . . 129

IV Discussion 140

7 Expansion During the Hydration of Magnesium Oxide 141


7.1 Hydration kinetics of MgO . . . . . . . . . . . . . . . . . . . . . . . . 142
7.2 Microstructure of MgO and Mg(OH) 2 . . . . . . . . . . . . . . . . . . 143
7.3 Volume fractions and relative volume changes . . . . . . . . . . . . . 152
7.3.1 Evolution of volume fractions during the hydration of MgO 153
7.3.2 Geometry and fractional conversion of a composite sphere 155
7.3.3 Relative volume change in the porous particles . . . . . . . . . 157
7.4 Bulk volume changes and multiscale microporomechanics for the hy-
dration of M gO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159
7.4.1 Multiscale Thought Model of MgO hydration . . . . . . . . . 159
7.4.2 LEVEL I: The composite sphere . . . . . . . . . . . . . . . . . 160
7.4.3 LEVEL II: The porous MgO-Mg(OH) 2 particle . . . . . . . . 167
7.4.4 LEVEL III: The continuum scale . . . . . . . . . . . . . . . . 172

9
7.5 Comparison with experimental measurements . . . . . 176

.
7.5.1 Hydrostatic state of stress (RD test) . . . . . . 176

.
7.5.2 Oedometric state of stress (LB test) . . . . . . . 178

.
7.6 Sum m ary . . . . . . . . . . . . . . . . . . . . . . . . . 180

.
8 Eigenstress Development During C3 S Hydration 181
8.1 Analysis of kinetics of C3 S hydration . . . . . . . . . . . . . . . . . 182

.
8.2 Microstructure of C 3 S . . . . . . . . . . . . . . . . . . . . . . . . . 185

.
8.2.1 SE M . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

.
8.2.2 Sorption isotherms . . . . . . . . . . . . . . . . . . . . . . . 187

.
8.2.3 Volume fractions . . . . . . . . . . . . . . . . . . . . . . . . 191

.
8.3 Eigenstress development during C 3 S hydration through a multiscale
thought-m odel . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 201

.
8.3.1 LEVEL I (The C-S-H gel) . . . . . . . . . . . . . . . . . . . 202

.
8.3.2 LEVEL II (C-S-H gel with capillary porosity) . . . . . . . . 205

.
8.3.3 LEVEL III. Reinforcement by Rigid (but Slippery) Inclusions 206
8.3.4 Eigenstress development . . . . . . . . . . . . . . . . . . . . 210

.
8.4 Speculation about the origin of Eigenstresses . . . . . . . . . . . . . 215

.
8.4.1 Powers and Brownyard Model . . . . . . . . . . . . . . . . . 216
8.4.2 Densification of C-S-H . . . . . . . . . . . . . . . . . . . . . .
. 219
8.4.3 Meso-scale simulations . . . . . . . . . . . . . . . . . . . . . 223
.

8.5 Sum m ary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227


.

9 Expansion of O-C 2 S 229


9.1 Kinetics of C 2 S hydration . . . . . . . . . . . . . . . . . . . . . 229
.

U.2 E
I!VoLUtU"- I VoULUcklln ofvlmefatin AUIng-L the h1ydratlin Uf C-20

9.3 Eigenstress development during the hydration of C 2 S . . . . . . 233


.

9.4 Sum m ary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 233


.

10 Development of Eigenstresses During the Hydration of Class-G Ce-


ment 235

10
10.1 Evolution of volume fractions during the hydration of Class-G cement 236

10.2 Bulk volume changes and multiscale microporomechanics model for the

hydration of Class-G cement . . . . . . . . . . . . . . . . . . . . . . . 247

10.2.1 Bulk volume changes in the Rubber Diaphragm Configuration 251

10.2.2 Bulk volume changes in the Liquid Barrier Configuration . . . 254

10.3 Sum m ary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 263

V Conclusions and Perspectives 264

11 Conclusions and Perspectives 265

11.1 Summary of main findings . . . . . . . . . . . . . . . . . . . . . . . . 265


11.2 Research contribution . . . . . . . . . . . . . . . . . . . . . . . . . . . 268

11.2.1 New experimental apparatus . . . . . . . . . . . . . . . . . . . 268

11.2.2 A multiscale microporomechanics model . . . . . . . . . . . . 268

11.2.3 Driving forces of bulk volume changes in CBM . . . . . . . . . 269


11.3 Industrial benefits and impact . . . . . . . . . . . . . . . . . . . . . . 269

11.4 Limitations and Perspectives . . . . . . . . . . . . . . . . . . . . . . . 270

A Evolution of volume fractions during hydration reactions 291

A.1 Hydration kinetics and mechanisms . . . . . . . . . . . . . . . . . . . 291

A.2 Evolution of volume fractions . . . . . . . . . . . . . . . . . . . . . . 297

A.2.1 Minimum water content and initial porosity for complete reaction304

A.2.2 Porosity of the individual solid phases . . . . . . . . . . . . . 305

B Elementary concepts and tools for micro-poro-mechanics modelling309

B.1 Representative volume and scale separability . . . . . . . . . . . . . . 309

B.2 Stress and strain field averages . . . . . . . . . . . . . . . . . . . . . . 310

B.3 Statistical indentation . . . . . . . . . . . . . . . . . . . . . . . . . . 311

11
Part I

Overview

12
Chapter 1

Introduction

1.1 Industrial context

The main objectives of the primary cementing of an oil well is to support the casing

string and provide zonal isolation by filling and sealing the annular gap between the

casing string and the borehole [136]. Lack of understanding of the state of stress in

the cement sheath and how the sheath may fail can lead to loss of zonal isolation.

Among all the possible reasons that can lead to failure of cement sheath, bulk vol-

ume changes that cement-based materials (CBM) experience during hydration in a

constrained environment is the focus of this study.

When cement slurry develops its mechanical properties, it also undergoes volume

changes during the early age of hydration. If constrained, these volume changes will

lead to the development of stresses in the hydrating cement. To predict failure due

to loading during the service life of the well, one needs to understand initial stresses

endured by these volume changes.

Prediction of stress and strain development in the cement sheath at early ages

is critical for performance evaluation. Yet, such a prediction calls for a profound

understanding of the chemical and physical mechanisms at early ages in the cement

sheath.

13
This study will involve the development of an experimental program designed for

the measurement of volume changes, and eigenstresses of hydrating cement pastes and

slurries under controlled pressure and temperature conditions. Appropriate mechanics

and kinetics models will also be developed for the interpretation of the experimental

data. The models will eventually be used for the development of an engineering tool

necessary for the prediction of the cement sheath behavior.

1.2 Research question

Can the driving forces found in cement literature explain the bulk volume changes

observed in saturated CBM? If not, then what are the driving forces of bulk volume

changes in CBM during hydration under saturated conditions? More specifically, can

the growth and densification of C-S-H explain the observed bulk volume changes in

saturated CBM?

Bulk volume changes of Cement-Based Materials (CBM) under autogenous (sealed)

conditions and drying (through evaporation) conditions have been investigated ex-

tensively in the cement literature. Driving forces of such bulk volume changes have

been attributed to capillary tension, disjoining pressure, crystallization pressure, and

surface energy of the solids as detailed in later sections. Yet, these driving forces

cannot explain the bulk volume changes observed under saturated conditions during

the hydration of the CBM. In this study, we investigate the bulk volume changes

of saturated CBM during hydration in an attempt to understand and quantify the

driving forces of bulk volume changes.

The driving forces generally accepted in the literature for autogenous and dry-

ing shrinkage cannot explain the bulk volume changes of saturated CBM. For this

purpose, alternative driving forces thought to control the bulk volume changes are

considered in this study. The first driving force is related to changes in colloidal forces

14
leading to contraction of the Low Density C-S-H (LD-CSH) whether it was due to

the progression of the chemical reaction or due to densification of the LD-CSH that

originally existed in an out-of-equilibrium state. This kind of force has never been

explored before but the recently published work of Ulm et al [195] shows that col-

loidal particles precipitating inside the porous C-S-H gel, can lead to volume changes.

The second driving force is due to the formation of High Density C-S-H (HD-CSH)

which is believed to form through a topochemical reaction leading to expansion of

the material. Expansion can also be a consequence of the volume difference between

the dissolving solids and the hydration products where the volume of the solids can

double at full hydration.

The formation of the LD-CSH and the subsequent densification depends primarily

on the reaction mechanism, rate of the reaction, presence of nucleation sites (seeds),

particle size distribution (PSD) of cement, and curing temperature. Expansion is

attributed to the volume difference between the reactive solids (anhydrous cement)

and the hydration products and leads to the observed expansion.

1.3 Methodology

This section details the approach to study the origin of bulk volume changes including

a general description of the experimental program employed to pinpoint the driving

forces of bulk volume changes. This section will also describe how to use the micro-

poromechanics model to relate the macroscopic measurements to their microscopic

origin.

Advances in experimentation and multiscale microporomechanics models made it

possible to establish the relationship between the macroscopic behavior of a heteroge-

neous material and its finer-scale origins (see, for example, [20] [153] [61]). With such

models, mechanisms and intrinsic properties that dominate the behavior of the ma-

terial can be incorporated at their corresponding scale, and their contribution to the

macroscopic behavior can be obtained. Investigating the origin of volume changes by

means of multiscale microporomechanics modeling of macroscopic scale experiments

15
will be the focus of this study. With focus on CBM, a number of reactive porous
materials will also be investigated, including the hydration of magnesium oxide and
pure cement phases.
Relevant measurements in this study will be obtained at macroscopic scale and
a multiscale microporomechanics model will be developed in an attempt to relate
the macroscopic observations to their finer scale origin. Multiscale micromechanics
in conjunction with understanding of the microstructure properties of the porous
media are necessary to quantify the mechanical properties and bulk volume changes
at the macroscopic scale. Bulk volume changes of CBM are possibly controlled by
several mechanisms that operate in various parts of the microstructure and at mul-
tiple scales. The origin of these volume changes will be introduced in a multiscale
micromechanics model as eigenstresses at the appropriate scale based on the under-
standing of the driving forces. A model for volume changes is developed by combining
the contribution of the different mechanisms in the multiscale micromechanics model.

The primary experimental setup employed in this study is a new apparatus de-
signed to measure bulk volume changes during the hydration of cement-based materi-
als under controlled curing pressure and temperature. The apparatus is also capable
of monitoring the hydration kinetics by measuring the volume of the reaction (chem-
ical shrinkage) of saturated samples. Other standardized tests will also be employed
to test the proposed mechanisms. These tests include:

1. Thermogravimetric Analysis (TGA) to monitor the progression of the hydration


reaction and the chemical composition of the material.

2. Scanning Electron Microscope (SEM) to investigate the microstructure of the


material.

3. Electron Probe MicroAnalysis (EPMA) to investigate the chemical composition


at the microscale.

4. Sorption isotherms to investigate the microstructure.

16
5. Isothermal calorimetry to monitor the reaction kinetics.

6. Indentations to investigate the microstructure.

The methodology followed to answer the research question can be summarized as


follows:

1. Bulk volume changes are measured at macroscopic scale using a new exper-
imental setup designed to measure the bulk volume changes under saturated
conditions while simultaneously measuring the volume of the reaction.

2. Systematic testing of a set of materials ranging in complexity from magnesium


oxide, C 3 S and C 2 S, and the Class-G cement are investigated. The tested
parameters also included the particle size distribution, curing temperature, and
the w/c ratio.

3. The microstructure of the material and the reaction kinetics are investigated
using a set of standardized tests.

4. Based on the understanding of the microstructure and the evolution of material


properties, a multiscale microporomechanics model is developed to interpret the
experimental data.

1.4 Outline of thesis

The goal of this thesis is to study the possible contribution of the densification of C-
S-H and the evolution of solid volume fractions to the observed bulk volume changes.
To achieve this goal, we start in chapter 2 by a review of the literature relevant to
the research question including driving forces and models of bulk volume changes,
mitigation strategies of bulk volume changes, and the available methods for the mea-
surement of the bulk volume changes. The main variables manipulated in this research
are presented in chapter 3 in form of a set of materials ranging in complexity from
the simplest Magnesium Oxide, then introducing the more complex C 3 S and C 25

17
followed by the most complex of all, Class-G cement. Chapter 4 introduces the main
experimental setup used to study bulk volume changes for the different materials
and test parameters, including curing pressure and temperature, effective stress, w/c
ratio, and the stress and displacement boundary conditions. Chapter 5 details the
framework for the analysis of the measured bulk volume changes and the different test
configuration obtained in the main experimental setup. Chapter 6 presents the ob-
tained results from the different tests on the studied materials. Chapters 7-10 discuss
the results obtained on the studied materials along with the multiscale microporome-
chanics model developed for each material for the analysis and interpretation of the
experimental measurements. In chapter 11, we draw the conclusions by presenting a
summary of the main findings as well as a highlight of the research contribution and
the possible industrial benefits, and gives suggestions for future research.

18
Chapter 2

Literature Review

In this chapter, we review the measurements and interpretations of bulk volume


changes in general while focusing on the relevant conditions. This review is necessary
to highlight the contribution of this study and set the ground for the experimental
design and data interpretation.

2.1 Measurements of bulk volume changes

Bulk volume changes have been the subject of extensive research due to their critical
role in durability and performance of CBM, especially in high performance concrete.
The main body of the literature focuses on autogenous shrinkage, drying shrinkage
and, to a much lesser extent, the bulk volume changes under saturated conditions.
Of special interest in the study of bulk volume changes are saturated and autogenous

(sealed) curing conditions. Autogenous volume changes are the bulk volume changes
(expansion or shrinkage) of cementitious materials during the hydration of cement
paste. Autogenous volume changes are measured under isothermal conditions for
sealed (no mass exchange) and free of stress samples. Autogenous shrinkage of CBM
have been measured both by volumetric and linear methods. Although designed to
measure the same quantity, volumetric and linear measurements are fundamentally
different due to experimental considerations related to boundary conditions.
In volumetric measurements, cement slurry is placed inside a tight rubber mem-

19
brane and immersed in water; the change in slurry volume is then measured by
dilatometric or gravimetric measurements [7,115,117,132,172]. One of the main chal-
lenges in the volumetric measurement of autogenous shrinkage is the sedimentation
and the associated bleeding water. Investigation of the bleed water on the measured
autogenous shrinkage has been the subject of investigation of many researchers (see,
for example, [24,107,130]). While bleeding in volumetric measurements resulted in an
overestimation of autogenous shrinkage at early ages [107], it could have been the rea-
son expansion was observed in linear measurements [130]. The linear measurement is
carried out by placing the cement slurry in a prismatic mold with displacement sensors
on one or both ends to measure the deformations [6,7,9,24,59,84,95,108,117,126,127].
In both volumetric and linear measurements the focus of the measurement is on the
bulk volume changes without pore pressure measurement (and effective stress) or
monitoring the hydration kinetics.

Bulk volume changes have been measured for samples with different mix designs,
compositions, and curing conditions. In general, shrinkage is observed and is typically
attributed to the self-desiccation resulting from the negative volume of the reaction
(chemical shrinkage). In some cases, an expansion was measured at early ages during
the hydration of CBM under autogenous conditions [7].

Another interesting observation is the impact of particle size distribution on the


measured autogenous shrinkage. Bentz et al [17,18] studied the influence of particle
size distribution on the bulk volume changes under autogenous conditions. By com-
paring the bulk volume changes of cements with different particle sizes at the same
water-to-cement ratio, they observed significant variation in the bulk volume changes
and the associated eigenstresses. These variations were then attributed to the pore
size distribution and the subsequent decay in relative humidity which reduces the
associated capillary stresses in the hydrating CBM.

Among the main parameters that control the properties of the hydrating and
hardened CBM is the initial water-to-cement mass ratio (w/c), which can also be
viewed as the initial porosity. Another way to view the initial w/c is by relating it
directly to the average spacing between cement particles [12]. The impact of w/c on

20
w

1400 -1400-
4" 1200 C 1200-
1000 -- W/C=0.25 M + W/C=0.25
CD 800 -o- W/C=0.30 40 800 -o- W/C=0.30

-
0
-W/C=0.35
600- W/C=0.35 .E 600-
.

400shrn + W/C=0.40. -e- W/C=0.40


400 shrnka -%-/C=0.45 400 -U-W/C=0.45
.C4)

-
0) 200 -- W/C=0.50 8 200- -o- W/C=0.50
0 0 1 -+- W/C=0.60 r_
0 i +- W/C=0.60
01 swelling W
CD Of
-200 2 -200- I I
-400 4 -400-
0 50 100 150 200 250 300 350 400 0 10 20 30 40 50
Age (days) Age (days)
a) 0 - 400 days b) Zoom: 0 - 50 days

Figure 2-1: Linear autogenous deformations vs. age., initialized at Vicat setting time
up to 1 year on c'p20 x 160-n1 sealed samples of cement pastes with variable w/c. at,
T=20 C. Reproduced from [9].

the bulk volume changes has been investigated extensively ill the literature of cement

and, ill general, the lower the w/c the higher is the autogenous shrinkage [9, 18].

A relevant observation related to the effect of w/c on the bulk voluimne changes

can be found in the work of Baroghel-Bounv et al [9]. where cenenlt pastes with

high w/c experienced expansion up to 20 days followed by shrinkage (see figure 2-

1). Shrinkage in low w/c materials is usually attributed to the smaller porosity and

smaller separation distance between cement particles as a result of capillary tension.

In the light of the previously discussed densification of C-S-H. one can argue that

capillary tension imight be a consequence of the ongoing chemical reaction, but it is

the densification of C-S-H that is controlling the overall behavior of the material.

Densification of C-S-H and bulk volume changes seem to follow similar trends, yet.

only the densification of C-S-H can explain the expansion anld shrinkage of saturated
naterials as will be shown in the main results of this study.

2.2 Driving forces and models of bulk volume changes

This section reviews the drivilig forces of bulk vohune changes in mature ald hydrat-
ing CBM. This section will also help determiie the lmeasllralble quantities required to

build the microporomechanics model.

The driving forces of bulk volume changes will be listed ill the following subsections

21
with the corresponding models and experimental measurements.

It is important to build some understanding of the driving forces of bulk volume

changes in order to be able to isolate and elucidate the contribution of each force.

In this section, we review the driving forces found in the literature with focus on the

controlling parameters of each force.

The fact that hydration of CBM is accompanied by bulk volume changes indi-

cates the development of eigenstresses inside the material. These eigenstresses can

originate at multiple scales depending on the mechanism that is causing them. A

number of volume-change-inducing mechanisms can be found in the cement litera-

ture, among which, capillary tension predominates as the most accepted mechanism

(see for example [94]). In addition to capillary tension, eigenstresses in hydrating or


drying cement paste are generally attributed to stresses like changes in surface ten-

sion of colloidal particles [209], or changes in disjoining pressure [10]. Eigenstresses

can also result from temperature changes, crystallization pressure [206], and possibly

the densification and hardening of hydration products [195]. Whether it is one or a

combination of these mechanisms that dominate the bulk volume changes in cement,

it has been established experimentally that changes in relative humidity is necessary

for significant bulk volume changes to occur in drying specimens. Similar to drying

shrinkage the driving force of autogenous shrinkage is usually attributed to capillary

tension resulting from drop in relative humidity. While in drying shrinkage drop in

relative humidity is due to loss of water through evaporation, under autogenous test

conditions it is attributed to consumption of water in the chemical reaction [94]. It

should be noted here that bulk volume change due to change in relative humidity

does not necessarily imply that capillary tension is the driving force due to the fact

that drying is associated with various physical and chemical changes in CBM.

A bulk volume change of a porous material is a manifestation of the interplay be-

tween a number of possible driving forces. Significant bulk volume changes of CBM

can be induced experimentally by desaturation, solvent exchange, and temperature

variation. There exists ample experimental evidence that variation of degree of satu-

ration (degree of saturation, Sw, is the volume fraction of the pore space occupied by

22
the pore solution) is one of the most influential means to induce bulk volume changes.
Irrespective of the driving forces, which will be discussed later on, the change in de-
gree of saturation is associated with certain changes in the pore space and the pore
solution. These changes include the formation of gas bubbles and changes in the
chemical and physical properties of the pore solution.

Desaturation is the loss of water from saturated porous solids, whether it was due
to the chemical reactions (Self-desiccation [158]) or through evaporation (drying). It is
also necessary here to understand the difference between self-desiccation and drying.
Drying involves the evaporation of all volatile phases from the pore solution while
keeping the amount of non-volatile solute constant. Drying is also associated with
the development of stress/strain gradients and the dimensions of the test specimen
are critical for the analysis of such phenomenon. That is to say, homogeneous test
conditions for bulk volume changes are virtually impossible to achieve in a drying test,
nonetheless, reasonably-thin specimens can be employed to create near homogeneous
conditions. Self-desiccation in sealed specimens on the other hand, involves the loss
of only water to the chemical reaction and it is possible that this kind of test is the
only test to provide homogeneous conditions.

Pore solution is always saturated with all ionic species that cement can provide

(see for example [167]). Calcium ions are released continuously in the pore solution
from the dissolution of gypsum, free lime, and all the major cement phases. The
concentration of calcium ions is maintained slightly above the portlandite saturation
point [182]. Alkali ions (Na+,K+) are mainly generated from the dissolution of alkali
sulfates in the pore solution [182] and the major part of the alkalis are introduced
into the pore solution at very early ages during the mixing process [180,188]. The
very high solubility limit of alkali products ensures that the pore solution is always
under-saturated with alkali ions [150].

Precipitation of hydration products during the hydration reaction is driven by


the dissolution of cement. The chemical composition of the pore solution during the
reaction depends on the composition of cement, and it can change during the reaction.
Another way to trigger precipitation of hydration products is the loss of solvent by

23
self-desiccation or drying which can increase the supersaturation with respect to a
number of hydration products. The precipitation of hydration products due to loss
of solvent can change the chemical composition of the pore solution especially in the
presence of alkalis.

The change in chemical composition of the pore solution, in general, will have an
impact on the activity of water (Raoult's law and relative humidity) [64, 145], the
surface tension [140,175], the dielectric permittivity [203], and the viscosity of the
electrolyte solution [63,93, 103]. The type of ions in the pore solution will also have
an impact on the colloidal forces between charged surfaces by changing the counter
ions and consequently changing the electrostatic coupling [154]. pH of the solution is
also expected to change if Ca(OH) 2 is to precipitate and consequently changing the
surface charge density of the colloids.

Chemical composition of the pore solution can also change in saturated material
due to the decrease in the porosity, where depending on the w/c ratio, porosity can
decrease significantly. The decrease in porosity can also lead to the increase in the
concentration of alkali ions in the pore solution and change its chemical composition.

As part of this study, an evaporative analysis is conducted to illustrate the impact


of alkali ions on the properties of the pore solution due to evaporation. To investigate
the impact of evaporation, half a liter of pore solution is extracted from Class-G ce-
ment mixed continuously for 24 hours in a sealed container with w/c = 1.0. The pore
solution is extracted in a MILLIPORE Stericup with 0.22 pm filter under vacuum.
A sample of the pore solution is then collected, and the remainder of the solution is
placed in an evaporation dish inside a desiccator under vacuum with silica gel beads.
Vacuum is used to prevent interaction with carbon dioxide from the atmosphere and
the silica gel is used for evaporation through capillary condensation. Samples of the
pore solution are then collected at different evaporation stages. Figure 2-2 shows
the color change during evaporation and the development of pH resulting from the
presence of alkali ions and the common ion effect.

24
13.4

-
13.3

13.1

13

-M G"12.9-

-
12.8
0 20 40 60 80 100
Water loss (%)
(a)
(b)
Figure 2-2: Evaporative behavior of pore solution of Class-G cement. (a) a picture of
the collected samples of the pore solution showing the color change due to evaporation.
(b) development of pH during evaporation due to the common ion effect. Water loss
is the percentage of water evaporated from the original pore solution.

2.2.1 Capillary pressure

Drop in relative humidity due to moisture exchange with the environment or self-

desiccation will lead to the formation of a gas-liquid interface (gas bubbles). which if

present in narrow pores, will result in pressure drop due to the formation of menisci

[65]. Drop in pore pressure resulting from the formation of liquid-vapor interface can

stress the solid skeleton and cause the material to shrink [94]. This section is a review

of the capillary pressure as a possible driving force of bulk volume changes in hydrating

and mature cement-based materials. We start with a general discussion about, the

possible causes that can lead to the formation of liquid/gas interface followed by a
discussion of the Kelvin-Laplace equation and an estiiate of the bulk volume changes

induced by capillary pressure.

Pure liquids can be driven into a imetastable state either by superheating above

the boiling point T (boiling) or by stretching below its saturated -vapor pressure

p,t (cavitation). The liquid will eventually return to equilibrium by nucleation of


vapor hu lles [39]. If a vapor Nibble nucleus is formed ii a pure liquid. it tends

to grow or collapse depending on whether the radius of the mewlv formed mucleus is

25

I
larger or smaller than the critical radius r, [76]. The nucleation of the vapor bubble is
associated with energy gain proportional to the bubble volume. However, a minimum
amount of work is required to form the bubble in a metastable liquid

4
E(r) = 7rr3(p - psat) + 47r 2'7 (2.1)
3

where -y is the liquid-vapor surface tension, p is the pressure in the liquid, and r is
the radius of the bubble. This competition results in an energy barrier

167733
Eb= l~v
3 (p - psat) 2
(2.2)

resulting in a critical radius


_ 27y
2 -y (2 .3

)
(Psat - P)

and a nucleation rate (how long one should wait for the nucleation to occur) [89,152]

F = Fo exp Eb= Fo exp (2.4)


kBT 3kBT

where kB is Boltzmann's constant and T is the absolute temperature. The prefactor

FO represents the number of nucleation sites and the frequency for nucleation [89].

Cavitation pressures of tens of MPa's were reported for pure water [88,215], but pore

solution in CBM is not a pure liquid.

In addition to boiling and cavitation, if a gas or volatile vapor is dissolved in

the liquid, then changing the conditions to create a supersaturated solution can lead

to the formation of gas bubbles [116]. Supersaturation can be achieved in three

ways: First, decreasing the solubility of the gas in the liquid by decompressing the

solution which has been saturated at a higher pressure or increasing (typically) the

temperature of the saturated solution [31]. Second, supersaturated solutions can be

prepared by reducing the amount of liquid (solvent) in the saturated solution while

keeping the amount of dissolved gas constant (see for example the formation air

bubbles in ice [38,118,173] or crystals [207]). Third, gas can be generated chemically

26
or electrochemically in a solution [31, 147].

The saturation concentration ceq of a gas in a liquid is determined by the partial

pressure Ph of the gas above the solution and is described by Henry's law:

Ceq = rKhPh (2.5)

where Kh is Henry's constant. The presence of dissolved gas in water strongly affects

the initiation of nucleation in the solution. In general, the higher the gas concentra-

tion, the lower the energy barrier for nucleation, the smaller is the critical radius, and

the higher the nucleation rate [204]. The gas bubble nucleation rate in the presence

of dissolved gas is estimated in the same way as for the pure liquid, i.e. Eq. (2.4)

with the difference in the critical radius, where the critical radius becomes [191,204]:

2 'y
rc =- (2.6)

where c is the concentration of the gas in the liquid, p is the pressure in the liquid,

and

,= exp(P- sat ) (2.7)

It is evident that the critical radius is smaller than in pure liquids and nucleation will

therefore be easier. Dissolved gases can also act as surfactants at the solution-gas

interface and consequently, reducing the energy barrier [116]. Ionic strength [34,54],

impurities [3, 54, 216], microbubbles [26, 105,119], and solid surfaces [116, 131, 214]
can all reduce the energy barrier for nucleation. Cavitation has also been reported

in porous media under tension [81,141] and cavitation due to drying was observed

experimentally in micro capillaries [29,30].

Air is typically present in CBM as dissolved air in the mixing water. Air bubbles

can be entrapped in the slurry during the mixing process [156]. Gas can also be

introduced as part of the reaction products like the release of hydrogen gas from

aluminum or zinc reactions in an alkaline environment [40, 134,147, 151] and such

technology is used in oilwell cementing to maintain the hydrostatic pressure and

27
counteract the negative volume of the reaction of the cement sheath slurry [86,178].

For a conservative estimation of the critical radius of air bubbles in pure water at

25'C with psat = 3000 Pa and y = 0.072 N/m under 0.5 atmospheres, the critical

radius using Eq. (2.3) is approximately 3 microns. Pore sizes of larger than 3 microns

are abundant in cement-based materials. In conclusion, formation of vapor bubbles

in cement is feasible whether due to drying or self-desiccation.

Formation of gas bubbles in cement is possible due to drying or self-desiccation in

sealed samples. Following is a discussion of the capillary pressure and the associated

stresses. Capillary pressure is the jump in pressure across the interface between two

immiscible fluids in a capillary:

Pc = Pnw - Pw (2.8)

where pc is the capillary pressure, Pn, is the pressure in the non-wetting (less wetting)

phase, pw is the pressure in the wetting phase (more wetting). In capillary tubes, the

pressure jump across the interface depends on the interfacial tension between the

fluids -y, contact angle 9, and capillary radius r. For 0 = 0, it is reasonable to

assume that the meniscus is hemispherical with radius of curvature equal to that of

the capillary.

Pc = - cos() (2.9)
r

In porous media, the pressure jump across the interface depends on the interfacial

tension between the fluids -y, contact angle 9, pore size distribution, and degree of

saturation Sw. The pore size distribution represents the volume fraction of each of

the pore sizes present in the porous medium and the degree of saturation Sw is the

volume fraction of porosity occupied by the liquid phase. Depending on the pore size

distribution, the loss of saturation will drive the liquid-gas interface (meniscus) into

pores with progressively smaller diameters, which results in drop in relative humidity.

28
Relative humidity can be related to the curvature of the meniscus by Kelvin equation

In ( =_ 2Vm (2.10)
Po rRT

where p is the partial vapor pressure of wetting fluid, po is the partial vapor pressure

at equilibrium, -y is the surface tension between the two immiscible fluids, Vm is the

molar volume of the wetting fluid, T is the temperature, r is the radius of curvature,

and R is the universal gas constant. Leading to the Laplace-kelvin equation

RT
PC= -In - (2.11)

In partially saturated porous media, volume changes due to drop in capillary

pressure can be attributed to changes in relative humidity, degree of saturation (pore

size distribution) and the compressibility of the material. Detailed discussion and

derivation of the asymptotic shrinkage due to capillary pressure can be found in [192]

as summarized below:

For an elastic porous solid, the free energy of the solid and the interfaces is a

function of the strain, porosity, and degree of saturation,

T = 'I(e, , SW) (2.12)

where Sc, is the degree of saturation of fluid a:

Sa =S= (2.13)

where # is the total porosity and # is the fraction of porosity occupied by fluid a.

The state equations are obtained in the form:

S.+ SnwPnw =
- , #PC (2.14) o- - P
FmOdwE= ,h and Eq(.)ed

For undeformable solid, for which 0 = Oo, the state equation and Eq. (2.8) reduce

29
to:

0P dI(S) (2.15)
dSw
For an undeformable solid skeleton, the energy T reduces to the interface energy qoU;
and use in Eq. (2.15) yields:

dU
Pc = Pnw - Pw= dSU (2.16)

with U the integral over the capillary pressure-saturation curve. Then, Kelvin's law
(2.10) provides the link between relative humidity and capillary pressure.

The mean stress in a non-saturated porous solid, reads:

oUm = KE - b7r(t) (2.17)

The asymptotic shrinkage due to drying in a stress-free material (Urn = 0), thus is:

b
E(t) = -ir(t) (2.18)
K

where b is Biot coefficient, K is the drained bulk modulus, and 7 is defined as:

, = S.pw + Sn.pnw - U = S.pw + SnwPnw -] P(S)dS (2.19)

Strain due to capillary pressure requires changes in the relative humidity, which
can be achieved by drying for example. To prevent possible contribution of such
pressure in this study, samples are maintained saturated throughout the tests by
providing water to the sample under controlled pressure.

2.2.2 Disjoining/Joining pressure

One of the driving forces of bulk volume changes is the disjoining pressure. Disjoining
pressure is the hydration pressure built up between surfaces at nanometer separation
due to the interaction between the water molecules and the hydrophilic surfaces of

30
the solid [120]. During the hydration reaction of cement a nanoporous C-S-H gel

is formed in addition to Portlandite and Calcium-Alumino-Ferrite phases. C-S-H

gel has especially a large internal surface, which in addition to the cations adsorbed

on these surfaces will give rise to attractive and repulsive colloidal forces. At high

relative humidity, water adsorbed on the solid surface is at the origin of disjoining

pressure [10,209]. Disjoining pressure is defined as the sum of attractive and repulsive

colloidal forces including Van der Waals forces, the double layer repulsion and other

colloidal forces [72] and it is activated due to loss of moisture from the gelpore space

in C-S-H (changes in relative humidity).

2.2.3 Surface energy

The hydration products of cement have a very high internal surface with the main con-

tribution being from the C-S-H. Down to very low relative humidity, large quantities

of electrolytic liquid are adsorbed on the internal surface of the hydration products.

Adsorption of water on a solid surface at very low relative humidity changes the

solid surface energy which can create considerable hydrostatic stresses at the col-

loidal scale [209]. In general, adsorption of water on the solid surface will result in

relaxation of the surface tension while desorption of water will increase the surface

tension and compress the solid [66,94]. The change in free energy of pure adsorbent

can be written as [69,162]:

AF rRT P2 dp (2.20)

where n is the number of moles of adsorbate on a fixed mass of adsorbent, S is

the solid surface area, R is the gas constant, T is the temperature, and p is the

vapor pressure. The length strain i due to change in surface energy can be written

as [5,66, 157, 208]


--l kAF; k = Ps (2.21)
E

And
-L (ki<n) In P2 (2.22)
1 Ss P1
31
where ps is the density of the solid and E is the elastic modulus of the solid. One way

to change eigenstrains is by changing the relative humidity to very low values, which

is not relevant to our study at this stage. Another way to change the eigenstrains of

the surface energy is through solvent exchange [68] [210].

2.2.4 Crystallization pressure

Stress exerted by growing crystals in confined space and the dissolution (and melting)

of crystals under pressure have been the focus of research since 1849 [189,190]. The

driving force required to lower the freezing point or increase the solubility of crystals

under stress can be related to the degree of supersaturation by Correns' equation

[50,51,75]:

ac = R In - (2.23)

where co is the saturation concentration for the unloaded crystal, c is the actual

supersaturated concentration, R is the gas constant, T is the absolute temperature.

This definition of crystallization pressure assumes that the activity coefficient of the

different constituents of the system to be unity.

The fundamental driving force of crystallization (precipitation) pressure in aque-

ous supersaturated solutions is the difference between the chemical potentials of the

ions in the supersaturated solution yi and the corresponding value at equilibrium

Peq [110,170]:
Af = feq - 1-t,= RTIn ( ao (2.24)

where a and ao are the activity coefficients of the solution and the solution at equilib-

rium, respectively. If the activity coefficients are ignored, as well as the deformability

of the system, then the stress that is required to suppress the growth of the crys-

tal can be estimated using Correns equation Eq. (2.23). The derivation of Correns

equation can be found in the original article [511 and the commented translation [75].

The degree of supersaturation required for the crystal to grow will increase and shift

the equilibrium towards the reactants and eventually stop the reaction by increasing

32
the stress on the reaction products [50,206]. Stress will increase the solubility of the
crystals and it does not differentiate between the reactants and the products when
both materials are present in the same space. It is worthwhile to mention here that
the stress on the solid has a different effect on the chemical reaction when compared
to the pressure in the pore solution. Stating the Le Chatelier's principle

When a system at equilibrium is subjected to change in concentration,


temperature, volume, or pressure, then the system readjusts itself to (par-
tially) counteract the effect of the applied change and a new equilibrium
is established [49].

In a system where the volume of the reaction products is larger than the volume
of the solid reactants, stress on the products will shift the equilibrium towards the
reactants and decelerate the reaction, while stress on the reactants will shift the
equilibrium towards the products and accelerate the reaction. The effect of the pore
pressure is the opposite of the "effective" stress on the solids in that, increasing the
pressure will accelerate the reaction if the volume of the reaction is negative.
A number of experimental designs can be found in the literature to measure the
crystallization pressure. The earliest types of experiments as conducted by Correns
et al. [51], by applying stress on a crystal in a supersaturated solution and monitor
the change in length of the crystal. Other set of experiments involving the cyclic
crystallization of salts in porous solids was also used as a proof of crystallization
pressure (see for example [74]).

2.3 Mitigation strategies of bulk volume changes

Autogenous shrinkage of CBM is often hypothesized to result from the capillary pres-
sure, which develops due the negative volume of the reaction (chemical shrinkage)
in sealed samples and resulting in the self-desiccation. The magnitude of the au-
togenous shrinkage as a consequence of the self-desiccation depends on the surface
tension of the pore solution, pore size distribution and geometry, and the elastic and

33
visco-elastic properties of the porous solid [14]. An understanding of the successful

mitigation strategies can give a critical insight into the driving forces of bulk volume

changes, given that a proper model and interpretation of the impact of these strate-

gies is possible. Here, we take a look at some of the effective mitigation strategies

and the possible interpretation of such strategies.

The mitigation strategies discussed here include the application of super absorbent

polymers (SAP) [101, 125], porous lightweight aggregates (LWA) [11, 111, 205], and
shrinkage-reducing admixtures (surfactants) [13, 137, 161]. Both the soaked light-

weight aggregates and super absorbent polymers provide internal reservoirs of water

that prevent large drop in the relative humidity. The LWA should possess larger pore

size distribution than that of the C-S-H in the cement paste such that the resulting

capillary pressure upon drying (or self-desiccation) is limited by the pore size distri-

bution in the LWA. The water content of the SAP, on the other hand, is dependent

upon the relative humidity and the pH of the pore solution and delivers water to

the hydration reaction in a similar fashion. The main function of shrinkage-reducing

admixtures is to reduce the surface tension of the pore solution and consequently,

reduce the resulting capillary pressure.

In general, the interpretation of the shrinkage mitigation strategies is based on the

concept of the self-desiccation (or drying) and the resulting drop in capillary pressure.

SAP and LWA prevent large drop in relative humidity by providing an additional

source of water inside the paste, while shrinkage-reducing admixtures reduce the

surface tension of the pore solution.

It is also important to note here that shrinkage reducing admixtures are known

to significantly reduce the concentration of alkalis in the pore solution [161]. And

the autogenous shrinkage of cement paste mixed with 0.5M potassium hydroxide was

reported to be double that of pastes mixed with deionized water [112,210]. The ob-

servations regarding the effect of alkalis on the volume changes, should be considered

when interpreting the mechanisms of mitigation strategies, yet, such analysis is ab-

sent from the literature. It is possible that the mechanisms of shrinkage-mitigation

strategies are related to changes in the chemical composition of the pore solution

34
rather than the changes in relative humidity.

2.4 Colloidal forces in C-S-H gel

In cohesive-frictional materials like moist sand, strength of the material can be at-
tributed to the friction force between the grains and the cohesion due to the formation
of capillary bridges between sand grains [25]. Strength of cohesive clays on the other
hand, is dominated by cohesion forces due to the significantly smaller grain sizes.
However, clays are much less sensitive than sand to changes in relative humidity,
which indicates the presence of other cohesive forces. Forces prevalent in sand and
clays cannot yet explain the strong cohesion in CBM, which is only slightly sensitive
to changes in relative humidity [149].

Advances in molecular modeling of cement based materials provide a powerful and


versatile means to exploring the structure and properties of C-S-H [96,121,122,148].
The binding phase in cement (C-S-H) which also forms the matrix, in which all other
phases are embedded, can be modeled as a dense dispersion of colloidal particles
with charged surface interacting through a liquid (see for example [98,106,149,157]).
Forces between such colloids can be categorized into electrostatic, van der Waals, and
solvation forces [97,113,148].

While van der Waals and solvation forces cannot explain the strong cohesion
in CBM, it has been suggested that ionic correlation forces (ICF) resulting from
thermally activated dynamical fluctuations are at the origin of cohesion in CBM at
early ages [1491. In general, the direction and magnitude of ICF depends on the
strength of the electrostatic coupling [60]:

= |- qlRhd
(2.25)
4EOCrkT

where o-, is the surface charge density, q is the ionic charge, Rhyd is the ionic radius,
E, is the relative permittivity and T is the temperature. For the most part it is
not possible to independently manipulate and keep any of these parameters constant

35
during the hydration of cement except for the temperature. ICF prevail at the scale

of C-S-H colloids and can be modeled as eigenstresses in a multiscale micromechanics

model at that scale as detailed in chapter 5.

Evolution of mechanical properties of CBM at later ages can be attributed to

several possible mechanisms. The first mechanism is due to the precipitation of larger

amount of the binding phase (C-S-H) due to the progression of the hydration reaction,

which leads to the reinforcement of the porous structure. Evolution of mechanical

properties due to such mechanism can be related to the mechanical properties of

C-S-H phase and the porosity of the system and is proportional to the degree of

hydration.

Development of mechanical properties of cement can also be due to changes in

the structure and composition of the C-S-H grains themselves as is the case with

aging and polymerization of the silicate chains causing the C-S-H to become stiffer,

stronger, and denser [187]. Another possible mechanism is related to the evolution

of the interaction forces between the colloidal C-S-H grains. The evolution of the

interaction forces between the surface of C-S-H particles can be a consequence of

changes in the chemical constitution of the pore solution.

Independent of the causes that can lead to changes in the interaction potential

between the C-S-H colloids, it was possible to parametrize the interaction potentials

to generate the stiffness of C-S-H known from nano-indentation studies [121, 122].

At early ages of hydration interaction potential was measured experimentally with

AFM experiments by Plassard et al. [154]. Precipitation and densification of C-S-H

was studies with meso-scale simulations by Masoero et al. [121,122] and Ioannidou

et al. [96]. In these simulations, the precipitation of the C-S-H gel is simulated by

successively and randomly inserting particles into a cubic periodic simulation cell. At

early stages of the simulations and due to the absence of mechanical percolation, all

the inserted particles are at equilibrium and the overall stress in the system is zero.

Once mechanical percolation if reached, however, insertion of a new particle will

induce stresses in the system because the existing particles are not free to rearrange

anymore. Stresses calculated from the experimental results agree qualitatively with

36
those from meso-scale coarse-grained simulations of C-S-H precipitation originating

from the electrostatic coupling between charged C-S-H particles mediated by the

electrolyte pore solution. These simulations are discussed further in Section 8.4.3.

2.5 Summary

This chapter is a review of the existing measurements, driving forces, and the mit-

igation strategies of bulk volume changes. The existing measurements are typically

concerned with the measurement of the bulk volume changes under controlled curing

conditions (sealed, saturated, and drying) at ambient temperature and pressure. The

existing measurements are also focused on the measurement of volume changes with-

out monitoring the hydration kinetics or the effective stress on the solid framework.

Driving forces found in the cement literature are focused on the change in relative

humidity due to drying or self-desiccation and as a consequence, all the driving forces

require a change in relative humidity to be activated. Such driving forces cannot

explain the shrinkage observed in saturated samples or the expansion of pure calcium

silicate phases. This chapter also discussed the shrinkage mitigation strategies and

as is the case with the driving forces, all the mechanisms are attributed to changes

in relative humidity and mainly to capillary pressure.

37
Part II

Materials and Methods

38
Chapter 3

Materials

The main goal of this study is to investigate the driving forces of bulk volume changes

in hydrating cement-based materials. The materials investigated in this study include

Class G cement, CS, C 2 S, belite cement, and light burn magnesium oxide. This

chapter will include the details about the materials used in this study as well as the

reasoning behind the choice of these materials.

3.1 Light-burn magnesium oxide

The first reaction to be investigated is the hydration reaction of light burn magnesium

oxide. The light burn MgO (also called caustic magnesia) is produced by calcination

at temperatures ranging from 700 C-1000 C. The material used for this investiga-

tion is a light burn magnesium oxide MagChem@30 produced by Martin Mrietta

Magnesia Specialties. MagChem@30 is a high purity technical grade of magnesium

oxide processed from magnesium-rich brine. Composition and physical properties of

magnesium oxide MagChem@30 are shown in Table 3.1.

The hydration of magnesium oxide will serve as the proof of concept for the

measurement of bulk volume changes of materials subjected to different boundary

conditions. The hydration of MgO will be especially necessary to compare bulk

volume change measurements from the Oedometric boundary conditions in the liquid

barrier test and hydrostatic state of stress in the rubber diaphragm test. Table 3.2

39
lists all the experiments conducted on MgO with different boundary conditions, curing
pressure, w/s mass ratios, and effective stresses.

Table 3.1: Composition and physical properties of magnesium oxide MagChem@30


from the product data sheet

Material Typical Specifications


Magnesium Oxide (MgO), % ignited basis 98.2 97.0 min.
Calcium Oxide (CaO), 0.8 1.0 max.

%
Silicon Oxide (SiO 2 ), 0.35 0.5 max.

%
Iron Oxide (Fe 2 0 3 ), 0.15 0.3 max.
Aluminum Oxide (Al 2 03), % 0.1 0.2 max.

%
Chloride (Cl), 0.35 0.5 max.
%

Sulfate (SO 3 ), 0.05 0.3 max.


%

Loss on Ignition, 1.7 2.5 max.


%

Bulk Density, loose, (g/cm 3 0.35


)

Median Particle Size, microns 3-8


Surface Area, m 2 /g 20-30
Activity Index, seconds 18 16-21
Screen Size, % passing 325 mesh, wet 99 95 min.

3.2 C 2 S and C 3 S

The second set of materials investigated in this study are the pure cement phases

C3 S tri-calcium silicate and C 2 S di-calcium silicate. Pure C 3 S and C 2 S materials are

obtained with three different particle size distributions shown in Figure 3-1.

C 3 S and C 2 S pastes are mixed with different water-to-solid mass ratios (w/s) and
different curing temperatures as listed in Table 3.3.

3.3 Class G cement

The main goal of this study is to investigate the bulk volume changes of oil well cement

during hydration, and for this purpose, oil well Class G cement is investigated. The

composition of the class G cement used in this study is shown in Table 3.4.

To investigate the effect of particle size distribution on the bulk volume changes,

Class G cement powder was separated into different particle sizes. Particle separa-

40
Table 3.2: MgO samples tested in the Pressure vessel
I
Speime 0 Pressure Test Effective stresstB
Specilinen w/', BCT
(g) (Mpa) Config* (MPa)
MgO-PVC-1 1 38.4 0.1 PVC t CBC
MgO-RDC-1 0.9 43.65 0.1 RDC ~ 0 HS
MgO-RDC-2 1 37.86 0.1 RDC 0.003 HS
MgO-RDC-3 1 38.2 1 RDC 0.003 HS
MgO-RDC-4 1 38.67 1 RDC 0.003 HS
MgO-RDC-5 1 36.84 1 RDC 1 HS
MgO-LBC-1 1 64.72 1 LBC t CBC
MgO-LBC-2 1 68.2 1 LBC t CBC
MgO-LBC-3 1 62.82 1 LBC t CBC
MgO-LBC-4 1 68.61 1 LBC t OBC
MgO-LBC-5 1 67.18 1 LBC t OBC
MgO-LBC-6 1 61.81 1 LBC t OBC
MgO-RDC-6 1 37.28 1 RDC 0.003 HS
MgO-RDC-7 0.71 35.27 1 RDC 0.064 HS
MgO-RDC-8 0.53 36.25 1 RDC 0.064 HS
MgO-RDC-9 0.5 38.24 1 RDC 0.014 HS
t Effective stress in RDC is applied through the check valve and the effective stress in
the LBC depends on the boundary conditionsas discussed in Sections (5.2)(5.3)(5.4).
CBC is the constrained boundary conditions. HS is the hydrostatic boundary conditions
and OBC are the oedometric boundary conditions. See chapter 5 for details.
* RDC is the rubber diaphragm configuration, LBC is the liquid barrier configuration and
PVC is the perneability and volume of the reaction configuration., see Sections (5.2)(5.3)
for details

t 7*I
10

8 --- C3S-G(62/100)
_____ __________ _____ ~- ~I
->C3S-M (32/62)
0
6 C3S-F (32)-
:E
LI 4

0
0.1 1 10 100 1000
Diameter (pm)

Figure 3-1: Particle size distribution (PSD) of C:IS and CS. PSD is shown for the
three C.5s sizes while CS has identical PSD.

41
Table 3.3: Properties of the tested C3 S and C 2 S samples. w/s is the
water-to-solid mass ratio at mixing, m' is the mass of solid used in the
test, Volume is the volume of the sample at the termination of the test,
T is the curing temperature, Density is the density of the sample at the
termination of the test, is the degree of hydration at the termination
of the test, and w/s final is the estimated final water-to-solid mass ratio
after placing the sample inside the pressure vessel.

W/S mO Volume T Density* W/C


initial (g) (mL) 0C g/mL final
C 3 S-F-RDC-01 0.49 63.97 49.65 25 1.97 0.74 0.46
C 3 S-M-RDC-01 0.40 72.55 49.74 25 2.03 0.34 0.37
C 3 S-M-RDC-02 0.42 70.06 49.32 25 2.02 0.60 0.39
C 3 S-G-RDC-01 0.34 80.28 48.75 25 2.17 0.48 0.29
C 3 S-F-RDC-02 0.49 65.10 48.50 25 1.95 0.81 0.45
C 3 S-M-RDC-03 0.42 68.65 48.90 25 2.02 0.70 0.39
C Q-'Dn-n1 n 4n A ) 71 1 4. 40 . An ) no .n on n An

C 2 S-M-RDC-01 0.38 77.25 49.80 40 2.10 0.34 0.35


C 2 S-G-RDC-01 0.34 80.78 48.20 40 2.18 0.29 0.30
C 3 S-M-RDC-04 0.42 68.59 48.45 45 2.01 0.72 0.39
C 3 S-M-RDC-05 0.42 69.71 48.74 35 2.03 0.71 0.39
* density measured at the termination of the test corresponding to the degree of
hydration
.

42
tion was achieved by two methods. The first method was through sedimentation in
isopropanol in which a 150g sample was sedimented in a IL sedimentation tower for
five cycles of 2hr sedimentation before the finer part (suspended) was removed in
each cycle. Sedimentation is not the most efficient method for particle separation but
with the aforementioned sedimentation procedures, the finer 40% of the sample was
separated.
The second separation technique was Air Jet Sieving. For this purpose, the sample
is sieved on the Air Jet AS200 by Retsch. The sample was sieved by on sieve No.635
(20 pm opening). 40% of the sample was retained on the No.635 sieve. An XRD
analysis is conducted on the different samples of Class-G cement to check for variation
in composition due to sieving. Figure 3-2 shows the XRD spectra for the different
CGC sample and indicated identical composition of all materials.
Samples were manually mixed for 2 minutes and degassed under vacuum for an-
other 2 minutes before placing in the rubber diaphragm apparatus (discussed later
on). Check valves with different cracking pressure were used to control the effective
stress as shown in Table 3.5.

Table 3.4: Chemical composition of Class G cement.

Oxide/mineral mass
%

Silica (Si0 2 ) 21
Alumina (A1 2 03 ) 3.5
Iron Oxide (Fe 2 03 ) 3.9
Calcium Oxide, Total (TCaO) 62.3
Magnesium Oxide (MgO) 4.4
Sulphur Trioxide (SO 3 ) 2.8
Loss on Ignition 1
Equivalent Alkali (as Na20) 0.54
C3 S 57
C2 S 17.3
C3A 2.5
C 4 AF 12
C 4 AF + 2X C 3 A 17
CSHO.5 3.8

43
__ - - - - -- - - - _.C - - - - - - __ - - - - - -_-_ :=

10000
- Class G
8000 -_ - -i ass Gfine (Sieying)
Class G course (sieving)
6000 -Class G fine (Sedimentation)
V)

4000

2000
-,
0
10 20 30 40 50 60 70 80 90
20

Figure 3-2: XRD patterns for Class G cement samples with different particle sizes as
obtained from sieving and sedimentation. All samples show identical XRD patterns
indicating identical composition and no preferential particle separation.

44

I
Table 3.5: Tested Class G cement samples in the rubber diaphragm test. Water-to-cement ratio was chosen to obtain workable
pastes and prevent significant sedimentation in the course samples. Effective stress (a + p) in the RDC tests is applied through
the check valve to maintain the direct contact between the rubber diaphragm and cement sample during the test.

Specimen w/C PSD p Volume Thickness Density W


(kPa) (MPa) (mL) (mm) (g/mL)
CGC-PVC-01 0.3 Full OBC 5 105.57 46.2 0.40
CGC-PVC-02 0.42 Full OBC 5 106.94 46.8 0.40
CGC-PVC-03 0.35 Full OBC 5 90.03 39.4 0.40
CGC-LBC-01 0.42 Full OBC 10 100.77 44.1 1.98 0.62 0.40
CGC-LBC-02 0.42 Full OBC 1 100.77 44.1 1.98 0.66 0.40
CGC-LBC-03 0.42 Full OBC 5 28.34 12.4 1.982 0.683 0.40
CGC-LBC-04 0.42 Full OBC 0.2 42.50 18.6 1.983 0.71 0.40
c-~1
CGC-LBC-05 0.42 Full OBC 0.2 40.68 17.8 1.9845 0.734 0.40
CGC-LBC-06 0.42 Full OBC 0.2 108.09 47.3 1.981 0.663 0.40
CGC-PVC-04 0.42 Full OBC 0.3 96.01 42.0 0.59 0.40
CGC-LBC-07 0.42 Full OBC 0.3 89.00 38.9 0.67 0.40
CGC-LBC-08 0.42 Full OBC 0.3 82.53 36.1 0.67 0.40
CGC-RDC-01 0.42 Full 14 1 48.24 - 2.07 0.64 0.35
CGC-RDC-02 0.36 C2 14 1 45.54 - 2.24 0.50 0.25
CGC-RDC-03 0.38 F1 14 1 47.09 - 2.09 0.72 0.33
CGC-RDC-04 0.38 Full 60 1 50.04 - 2.11 0.61 0.32
CGC-RDC-05 0.38 Full 60 1 50.04 - 2.12 0.62 0.31
CGC-RDC-06 0.315 Cl 60 1 47.31 - 2.20 0.38 0.26
CGC-RDC-07 0.49 F2 60 1 49.50 - 1.92 0.78 0.45
CGC-RDC-08 0.38 Full 60 1 49.44 - 2.11 0.62 0.32
3.4 Summary

This chapter introduced the materials and the test parameters investigated in this
study. Three materials are chosen for the study ranging in complexity from the sim-
plest MgO to the more complex C 3S and C 2 S and finally, the Class-G cement. The
hydration of MgO produces only one hydration product (Mg(OH) 2 ) and the resulting
porous material is strictly expansive with well-known properties for both the reac-
tants and the products. C 3 S and C 2 S react with water to form C-S-H gel in addition
different quantities of portlandite. The difference in reaction rate and the quantity
of CH produced from the hydration reaction helps determine the contribution of CH
to the observed bulk volume changes. The most complex material investigated is the
Class-G oil well cement comprising the known 5 phases in portland cement. The in-
vestigated parameters included w/c ratio, PSD, and curing pressure and temperature
as well as the stress and displacement boundary conditions.

46
Chapter 4

Experimental Setup

This chapter will focus on the design and the capabilities of the primary experimental

setup used in this study. The design of a new experimental setup is necessitated by the

main goal of this study, which is the measurement of bulk volume changes of saturated

cement-based materials (CBM) during hydration. To design the experimental setup,

it is necessary to specify the desired capabilities of the setup. For this purpose, we

consider the poromechanical quantities required to model a reactive porous solid. The

main quantities required for such model include the pore pressure, curing temperature,

bulk volume changes while monitoring the reaction kinetics under controlled stress,

displacement, and hydraulic boundary conditions. Unlike the existing measurements

of length changes, the current apparatus is designed to measure bulk volume changes

while monitoring the hydration kinetics through the measurement of the volume of

the reaction of saturated samples.

Measurements of bulk volume changes of sealed samples in the cement literature

(autogenous shrinkage) does not include the measurement of the pore pressure inside
the sample during hydration. The current design can be used to measure bulk volume

changes of sealed samples while monitoring the evolution of the pore pressure under

controlled effective stress. The current design can also be used for the direct and

indirect measurement of the evolution of the permeability during the hydration of

CBM.
The pressure vessel is an apparatus capable of simultaneous measurement of bulk

47
volume changes and chemical shrinkage under controlled curing conditions. Simul-
taneous measurement is necessary for the estimation of eigenstresses and the con-
struction of a microporomechanical model to interpret the experimental data. Ex-
perimental apparatus capable of such simultaneous measurements are scarce in the
cement literature with focus on either chemical shrinkage measurement or bulk volume
changes separately. The available apparatus are not sufficient for a proper microp-
oromechanical interpretation.
The apparatus is capable of measuring the bulk volume changes while controlling
the curing temperature, curing pressure, sample thickness, effective stress, chemi-
cal composition of curing fluids, as well as variations of stress, strain and hydraulic
boundary conditions. Experimentally, there are three possible boundary conditions
for the stress and strain.

1. Hydrostatic: Radial and axial stresses are identical and controlled, resulting
in a uniform volumetric strain. This state of stress is made possible with the
rubber diaphragm configuration as detailed in the following sections.

2. Oedometric: In this configuration, the radial strain is zero in case of swelling and
the volumetric stain is manifested entirely as axial strain. In case of shrinkage,
specimen can debond from the walls of the vessel at which point the radial
strain will no longer be zero. These boundary conditions are employed in the
liquid barrier configurations as detailed in the following sections.

3. Constrained: This configuration is similar to the oedometric boundary condi-


tions with constrained displacement on the wall of the vessel. These boundary
conditions are employed in the permeability and volume of the reaction mea-
surements as well as some liquid barrier test as detailed in later sections.

To control the hydraulic boundary conditions experimentally, the specimen can be


completely sealed or flow is allowed through the bottom surface only or through the
top and bottom surfaces.
Figure 4-1 shows the layout of the main apparatus used in this study. The appa-
ratus consists of two end caps attached to the cylindrical body of the pressure vessel

48
through shear blots. Two sources of pressurized water (S-1 and S-2) are connected
to the vessel through the top and bottom caps. The first source (S-1) is connected to
eccentric port (port-1) in the top cap. The water is injected into the vessel through
the eccentric port (port-1) while the central port (port-2) allows air to leave the vessel
while filling it with water. The second source (S-2) is connected to the vessel through
the port in the bottom cap (port-3).
Volume of the reaction (chemical shrinkage) is measured by injecting water inside
a control volume under constant pressure. In the first configuration (Figure 4-6(a))
water is injected through port-1 and has access to the top surface of the specimen
only. The bottom surface of the specimen is resting on a filter and connected to a
pressure transducer to monitor the evolution of the pressure at the bottom of the
specimen. Not all the water injected into the system (Vx) is contributing to the flux
through the surface of the sample as part of it will compensate for the bulk volume
change. The bulk volume changes thus needs to be measured directly by covering
the specimen with a liquid barrier while valve-1 is open and valve-2 is closed. The
volume injected through port-1 is Va,,k and the volume injected through port-3 is

Vab. This configuration is shown in Figure 2(b). Another possible configuration to be


used for the simultaneous measurement is the rubber diaphragm (Figure 2(c)). The
rubber diaphragm design has the advantage of creating hydrostatic state of stress and
eliminating the possibility of leakage.

4.1 General Layout

This section describes the general layout of the experimental setup with focus on the
design goals and specifications of the different components. Following is a description
of the general layout and the main components of the flow network.

4.1.1 Pressure cell

The pressure cell is a cylindrical vessel with top and bottom caps that allow full
access to the inside of the vessel (see Figure 4-1). The vessel used in this study is a

49
modified double cap test cell by OFI Testing Equipment, Inc., model number 170-45.
The cell is made out of grade 316 stainless steel with an inner diameter of 53.98 mm,
outer diameter of 82.55 mm, and inside vertical clearance of 65.8 mm with a pressure
rating of 13.8 MPa (2000 psi). O-rings are used as the sealing mechanism between
the end caps and the body of the vessel and the end caps are attached to the body
of the vessel through shear bolts (set screws).

The top cap has an eccentric port (port-1) used to inject water inside the vessel
under controlled pressure during the test and it is also used to fill the vessel with water
at the beginning of the test. The top cap also has a concentric port used to vent air
out of the vessel at the beginning of the test and as a port for the temperature probe
during the test. The bottom cab has only a concentric port, which can be used to
inject water inside the vessel or it can be connected directly to a pressure transducer
to measure the evolution of pressure at the bottom surface of the sample.

4.1.2 Syringe pumps

The syringe pumps (S-1 and S-2 in fig.4-1) are used in this test to supply water under
controlled pressure and provide accurate volume measurements. The syringe pumps
used in this study are low-flow precision solvent delivery pumps by Teledyne ISCO,
model number 100DM Pump Module. The 100DM pump module has a displacement
resolution of 4.8 nl and a pressure range from 0.07 to 69.5 MPa (10 to 10000 psi).
For the main part of this study the syringe pump S-1 is mainly used to fill the vessel
with water at the beginning of the test and to restore the pressure in the bladder
accumulator.

The major draw back of the syringe pumps is the flow pulsation and the inaccurate
control of pressure if the two pumps are used to control the effective stress on the
sample. The main advantages of using the syringe pumps is the capability of delivering
programmed pressure profile and the fact that it measures volumes directly instead
of the flow rate. To avoid the problem of pulsation, the syringe pumps are replaced
by a bladder accumulator.

50
M, am. M -- p - a

-
Air Bladder
Femperature Accumulat or
probe
S-1

Submerged
Temperature
probe

FM-2

T alve-2
Port-2 Port-i FM-1
Top cap
Eccentric mass Aluminur
vibrator collar Valve-1

.4
O-ring
Hydraulic fluid
S-2
Filter
Vessel body
Specimen
O-ring
Valve-3

- Set screw
-heck
Bottom cap
Port-3

FM-3

PT-1
Valve-4

Figure 4-1: Schematic of the apparatus showing a cross section in the pressure vessel
and the flow network. The pressure vessel consists of two end caps attached to the
cylindrical body of the pressure vessel through shear bolts (set screws). S-1 and S-
2 are sources of pressurized water (syringe puilps) or a bladder accnniulator that
can deliver water to the network at controlled pressure. Water is injected into the
vessel through the eccentric port (port-1) in the top cap while the central port (port-
2) allows air to leave the vessel and is also used to insert the temperature probe.
Pressurized water is also connected to the vessel through the port in the bottom cap
(port-3). Valve-i, Valve-2, Valve-3, and Valve-4 are shtit-off valves. The filter is a
saturated filter paper placed on top of a perforated stainless steel disk that allows
the flow of water but not cement grains, while allowing access of water to the entire
bottoli surface of the specimen. The filter is also designed in such a way to provide
a rigid support for the specimen. FM-1, F-M-2, and F1\1-3 are two-way flow meters.
The purpose of the check valve is to create a differeintial pressnre between the top and
bottom ports when the vessel is connected to only one source of pressurized water.

51

___________ _________ I
4.1.3 Bladder accumulator

Delivering pressurized water without pulsation is the only reason a bladder accumu-
lator is used in this study. A bladder accumulator is a pressure (energy) storage
device in which an incompressible fluid (water) is held under pressure by compressed
gas (nitrogen). The compressed nitrogen is enclosed inside an elastic bladder, which
can be charged through a gas charge valve as shown in fig.4-2. The accumulator can
deliver pressurized water to the hydraulic line The bladder accumulator used in this
design is a Parker model (BA002B3T01A1) with 1 liter capacity and 20.7 MPa (3000
psi) pressure rating.
A common way to estimate the pressure variation in a bladder accumulator is to
use the ideal gas law and in isothermal conditions pV = const.. It is obvious that
the larger the volume occupied by the gas the smaller is the pressure drop due to
loss of water out of the accumulator. As water is injected into the pressure cell and
out of the bladder accumulator, the pressure in the gas and water will drop at a
magnitude proportional to the compressibility of the particular accumulator state. In
other words, if most of the volume of the accumulator tank is occupied by gas, the
drop in pressure will be insignificant due to the very high compressibility of the gas.
The bladder accumulator cannot be used to measure the volume directly, but the
smooth (pulsation-free) flow allows for the use of flow meters and accurate monitoring
of the hydration kinetics of cement-based materials.

4.1.4 Flow meters

Volume measurements in the current design can be carried out by three methods. The
first method is a direct measurement of volume injected through the top and bottom
ports by a syringe pump equipped with direct volume measurement capability (see
Section 4.1.2). Volume can also be measured by monitoring the pressure drop in a
bladder accumulator by relating the pressure, volume ,and the compressibility of the
accumulator (this method is not used in this study). In the third method, flow meters
are mounted at the desired branches of the flow network and the flow rate is measured

52
- - - - - --= - 7 - - _1W

Gas charge

Compressed
gas

Bladder

Accumulator
Pressurized tank
water

To flow
network

Valve

Figure 4-2: Schematic of the bladder accumulator.

at these locations (see FM-1, FM-2, and FM.-3 in fig.4-1). Flow meters can provide

an accurate measurement of the flow rates at different branches of the network, while

syringe pumps can only measure the volume injected in the entire system.

Flow meters employed in this design are the ultra high pressure very low flow

rate sensors based on thermal microsensor by Sensirion model number SLG64-0075.

These flow meters are bi-directional with maximum flow rate of 5 pl/mini in precision

imode and up to 20 pl/min in the extended mode with a resolution of 0.5 nl/mim.

Pressure drop across the flow meter is approximately lkPa/lpl/mnin: see Figure 4-3.

4.1.5 Check valve

Cement bleeds! Depending on the geometry and the boundary conditions of the

samnlple this bleeding can imipede the mieasurement of the bulk volume changes (see.

for example. [7. 115. 117, 132. 172]). In the absence of a pressure differential in the

rubber diaphragn test, water pocket can form between the rubber diapliragni and the

top surface of the sample due to sedinentation. Such problem can result in false bulk

53
1.4 1.2
y = 1.3298 + 0.0197 y = 1.104 x + 0.0225
1.2
1
1 - 0.8 - --- --

-
0.8
0.6
0.6
0.4
0.4
0.2
0
0
-

0.5
I
_______________

1
0.2

0
0 0.5 1
Flow rate (ml/min) Flow rate (ml/min)

(a) Bottom (FM-3) (b) Top (FM-2)

Figure 4-3: Head losses through the bottom (a) and top (b) flow meters. The mea-
surement range is much larger than the typical rates measured in cement tests, but
the head loss shows a linear response throughout. The measured ordinate intercept
is irrelevant.

volume changes due to the absorption of water by the hydrating sample. To overcome

the problem of sedimentation, a pressure difference is necessary to maintain a direct

contact between the sample and the rubber diaphragm, which applies an effective

stress on the sample. In the absence of such pressure difference, the stiffness of

the rubber diaphragm itself can play a role in the measured bulk volume changes.

Creating two different pressures in a flow network with a single source of pressurized

water can be achieved by means of a check valve as shown in Figure 4-4.

4.1.6 Filter

The purpose of the filter is to provide the sample with water by allowing the flow of

water while retaining cement grains. The filter is also designed in such a way to allow

water access to the entire bottom surface of the sample and provide a rigid support

in case of effective stress application. The filter is also built of inert material to limit

the reaction with the material especially at high pH found in cement-based materials.

To meet the design requirements, the filter is built of perforated stainless steel

disk with 0.5mm perforations distributed over the exposed sample surface (figure 4-

5). A filter paper is also placed between the perforated disk and the cement slurry to

54
P-
Syringe
Pa Pump
PIaM

Check
Valve

Figure 4-4: The concept of a check valve to control effective stress on the sample
during hydration. A check valve is placed in a flow network, as shown in the figure,
and creates a pressure difference between the top branch with pressure pt and the
bottom branch with pressure PA once the flow is established in the network. The
difference in pressure depends on the flow rate through the network, but for the
hydration of cement-based materials where the flow rate is very small at all times,
the pressure difference is equal exactly to the cracking pressure pr of the check valve
Pt - Pb = pr The pressure difference is necessary in the rubber diaphragm test
to maintain a direct contact between the sample and the rubber diaphragm, which
applies an effective on the sample equal to the cracking pressure. In this configuration,
pressure in the top branch is equal exactly to the pressure inside the source pt = Pa
while the pressure in the bottom branch is PA = Pa - Pcr. Creating two different
pressures in a flow network with a single source of pressurized water can also be
achieved by means of a pressure regulator.
--MMPMMRIL --

prevent the flow of cement grains across the filter.

. . 236 02.496

. - .
- . . Flow grooves

Top Bottom

0.106
I U 11 11 11 1 11 1 11

2.496

Cross section

Figure 4-5: Schematic of the perforated disk.

4.1.7 Water bath

The pressure cell along with part of the hydraulic lines are immersed in a water

bath where the temperature can be controlled to within 0.01'C. The water bath is a

custom made one cubic foot stainless steel tank connected to a Julabo circulator to

control the test temperature.

4.2 Configurations

The experimental setup can be configured to obtain a number of measurements,

including the volume of the reaction (chemical shrinkage), permeability, and bulk

volume changes. This section describes the different possible configurations that can

be obtained with the new experimental setup. The coifigurations are the Permeability

and Volume of the reaction Configuration (PVC). the Liquid Barrier Configuration

(LBC), and the Rubber Diaphragm Configuration (RDC).

56
Port-2 Port-I
Top cap
I'

0-ring
Curing fluid Hydraulic fluid Rubber
Vessel body Hydraulic fluid
Liquid Barrier Diaphragm

Filter Specimien
Spechien Specimen
O-ring

Set screw (r RDC


Bottom car
P )rt-3

(a) PVC (b) LBC (c) RDC


Figure 4-6: Three test configurations. In configuration (a). Port-i is connected to
a source of pressurized water and the specimen is exposed to water through the top
surface. Bottom surface is resting on the filter and Port-3 is connected to a pressure
transducer to measure the evolution of pressure at the bottom surface of the speci-
men. In this configuration. port-3 can also be connected to a second syringe pump
instead of the pressure transducer to allow for direct permeability measurement. This
coifiguration is referred to as the Permeability and Volume of the reaction Configura-
tion (PVC). In configuration (b), Port-1 is connected to a source of pressurized water
and the top surface of the specimen is covered with the liquid barrier. Bottom sur-
face is resting on the filter and Port-3 is connected to the same source of pressurized
water as Port-1 or a separate source of pressurized water if a differential pressure is
desired. This configuration is referred to as the Liquid Barrier Configuration (LBC).
Configuration (c) is the same as configuration (b) except for the liquid barrier being
replaced with a rubber diaphragm and the oedometric state of stress with hydrostatic.
This configuration is referred to as the Rubber Diaphragm Configuration (RDC). In
all configurations, Port-2 is first used as a vent to let the air out of the vessel while
being filled with water and then used to insert the temperature probe inside the vessel
before pressurizing the vessel.

57
4.2.1 Permeability and Volume of the reaction Configuration
(PVC)

In this configuration (see Figure 4-6(a)), cement slurry is cast directly inside the vessel

and is then covered with the desired curing fluid before pressurizing the vessel. This

configuration is employed to measure simultaneously the volume of the reaction and

transport properties of expansive-reactive porous solid. Expansion (or shrinkage that

cannot cause debonding between the sample and the vessel ) is a condition required

to guarantee the complete bonding between the specimen and the wall of the vessel.

This requirement in not necessary if the sole purpose of the test is to measure the

volume of the reaction.

4.2.2 Rubber Diaphragm Configuration (RDC)

Rubber membranes have been used extensively to measure the bulk volume changes

in cement-based materials under sealed conditions (autogenous shrinkage) as detailed

in Section 2.1. The premise of the rubber diaphragm is illustrated in Figure 5-12(a),

where the rubber membrane acts as a barrier between the fluid phases inside and

outside the porous sample while transmitting the effective stress to the solid skeleton.

In this study, the rubber diaphragm (membrane) is used to prevent the flow through

the sample surface while providing the capability of bulk volume measurement.

4.2.3 Liquid Barrier Configuration (LBC)

The premise of the liquid barrier test is based on the understanding of the ingression of

a liquid through the outer surface of a porous solid. An ideal liquid barrier is a liquid

that can function as an impermeable and flexible membrane (like a rubber membrane).

The function of the rubber membrane (in tri-axial soil testing for example) is to

impose a total stress (O-i = -ptj) boundary condition (Figure 5-12(a)) and more

importantly, prevent flux at that boundary (q - ni = 0). The pressure differential

between the hydraulic fluid and the pore fluid is resisted by the deformation of the

rubber membrane and transferred to the solid skeleton as the effective stress o . The

58
3 attachment 3 detachment
screws screws

6 vent holes 10-32 Tap

r 01.84"

10-32 Tap

P Half
B

Bottom Half

Monolithic O-ring
Dash Number 228

Figure 4-7: The Rubber Diaphragm Mold. The rubber diaphragm mold is a two-
halves mold made out of aluiinum with the shown dimensions. The Rubber Di-
aphragm is a 0.5 nmn-thick membrane built monolithically with a dash number 228
O-ring.

effective stress can be written as [53]:

CT . = O-i + pl)i6j (4.1)

where pb is the pore pressure. 6ij is the kronecker delta.

The idea of the liquid barrier test is based on the assumption that we can find a

liquid that can function as a flexible-imperieable barrier. Hence, a funetional 1iquid

barricr is a liquid that prevents flow through the top surface of the specimen.

qz = 0 (4.2)

This entails that the pressure gradient at the top surface vanishes:

K dp
q-- - 0 (4.3)
pi 6

Complete analysis of the performance of such configuration is shown in detail in

Section 5.4

59

I
4.3 Summary

This chapter introduced the main experimental setup used in this study to investi-
gate volume changes of cement-based materials under controlled curing conditions.
The apparatus is capable of measuring the bulk volume changes while controlling the
curing temperature, curing pressure, sample thickness, effective stress, chemical com-
position of curing fluids, as well as variations of stress, strain and hydraulic boundary
conditions. Three test configurations are possible with this design including the Per-
meability and Volume of the reaction Configuration (PVC), which can be used to
measure the volume of the reaction (chemical shrinkage) and the evolution of perme-
ability during the hydration of CBM. The second configuration is the Liquid Barrier
Configuration (LBC), which can be used to measure both volume of the reaction and
bulk volume changes under oedometric boundary conditions. The third configuration
is the Rubber Diaphragm Configuration (RDC), which can also be used to measure
the volume of the reaction and bulk volume changes, but unlike LBC, the state of
stress in the RDC is hydrostatic.

60
Chapter 5

Analytical Framework and


Boundary Value Problems

The experimental setup introduced in the previous chapter can be configured to obtain

a number of measurements, including bulk volume changes, volume of the reaction,

and permeability. The displacement, stress, and hydraulic boundary conditions of

the main three configurations (see Figure 4-6) can be varied to investigate the impact

of the boundary conditions on the measured quantities.

In this chapter, we develop the poroelastic state equation before and after the solid

percolation threshold. To this end, we consider the idealized case of the hydrating

CBM as a two-phase system, more specifically, as a pore-solid composite composed

of a solid phase and the pore space. In the first step, we upscale the behavior of

the composite; and in the second step, we recall the field equations, which the solid

and fluid phases need to satisfy. Then we apply the combination of the state equa-

tions and field equations to the particular boundary conditions of the experimental

configurations.

5.1 Microporomechanics framework

To model the two-phase composite (porous solid) we employ the framework of con-

tinuum microporomechanics [61] used to describe the response of the system to two

61
mechanical loads to which the solid is subjected. For a porous solid, these loads are
the regular displacement boundary condition at the boundary and pressure bound-
ary condition at the pore-solid interface. For hydrating cement-based materials, in
which the solid mass increases, the homogenization is carried out at a constant degree
of hydration and incremental changes in the boundary conditions. Following earlier
work [193-195], the heterogeneous stress distribution in a porous solid composed of a
solid phase and saturated pore space with eigenstress can be written in incremental
form:

(Vz E V) do m(z) = k(z)de(z) + duP (z) (5.1)

dsij = 2Gdeij (5.2)


1 1
d@| = bdc + du* +I dp (5.3)
No, Np

dpjk = bd + I(du* + dp) (5.4)


N

where z is the position vector, dam(z) is the incremental heterogeneous mean stress,
k(z) is the bulk modulus, de(z) is the incremental volumetric strain, and p is the
pore pressure. de.(z) is the incremental mean eigenstress that develops in the solid
phase due to thermal and chemical phenomena associated with the solid. That is,
stress in the solid phase that forms the matrix in which pores are embedded can be
represented by Eq. (5.1). 0 is the porosity, G is the shear modulus, b = 1 - 2 is the
Biot coefficient and K is the drained bulk modulus, N = ks where N is the solid
Biot modulus.

All poroelastic constants are related to the degree of hydration through the link
between volume fractions and texture parameters. a-. is the mean stress (hydrostatic
stress) component and sij is the deviatoric stress component.

1
0-m SOi-; Sij = 6i
9moij (5.5)

62
Similarly, c is the volumetric strain and eij is the deviatoric strain component

C= Eij; ey = Ej4 - 3i (5.6)


3

For a porous solid under drained conditions and in the presence of eigenstresses,
we shall apply the following distribution:

Jzk' for z E V, d d* for z E V

0 for zGV, -dp for zEV

where kS is the bulk modulus of the solid, V, is the volume of the solid, V is the
volume of pores, p is the pore pressure, and do-* is the incremental mean eigenstress
that develops in the solid phase due to the progression of the hydration reaction.
Homogenization of Eq. (5.1) with Eq. (5.7) entails the macroscopic incremental
stress state equation:

dEm = (dam(z)) =(1- #)(dam(z))Vs + #(dam(z))VP (5.8)

dEm = (dom(z)) = k' (1 - b) dE, + (1 - b) do* - bdp (5.9)


K d>2P

where angle brackets imply volume averages. The stress on the sample can also be
written in terms of incremental mean effective stress dM

dE' = d(Em + p) = kS(1 - b)dE, + (1 - b)d(J* + p) (5.10)

where Em is the macroscopic mean stress, p is the pore pressure, d(Em + p) is the
incremental effective stress, dE, = (dF(z)) = r(de(z))vs + (dE(z))VP is the incremen-
tal macroscopic volumetric strain, n is the solid volume fraction and # is the porosity
(= 1 - ), and b = #5(A)V = 1 - T(A)Vs = 1 - K is the Biot coefficient. (A)VP
and (A)Vs are the volume averages of the volumetric parts of the fourth order strain
localization tensor, dE(z) = A(z)dE,.

Similarly, microporomechanics provides the tools to determine the change in

63
porosity due to mechanical loading and eigenstress of the solid. The incremental
change in the solid volume fraction at constant degree of hydration can be written as

dq = ldc(z))Vs = r,(dc(z))vT(do + q (dcv(z))V dE=0 (5.11)

where

y7(dEv(z))vsd,= = (1 - b)dE, (5.12)

T/(dcv(z))VS dEv=0 = (/ dorm(z) - douP(z))vs ) (5.13)

Then note that (d-P(z)) Vs = do-* and that dEm = 77 (do-m(z)) " -dp, while dEmIdEv=0
dEP; so that

r/ (dom(z)) dEv=0 = dEP + #dp . (5.14)

q (dom(z) - do-P(z)) V dEv=0 = dEP + dp - 77 du* (5.15)

Thus:
d = (1 - b)dE- bks ) (du* + dp) (5.16)

Finally, if we note that dEvjI = dr/Ig + doj , we thus obtain:

d = bdE + k ) (du* + dp) (5.17)

The incremental change in porosity can then be related to the incremental change in
mass content of the system at constant degree of hydration

dM*| -MW
- d4g+ dpw r 1R
(5.18)
PW
P fl

It is useful to define here the isothermal compressibility

S= -- - -1- (5.19)
V \p T P OP T

64
And the volumetric thermal expansion coefficient

av = T (5.20)
P pOP

The density change of the pore solution depends only on the pressure and temperature

is:

dpw pT P )dT = pw (13wdp - a"'dT) (5.21)

The rate of change in the density of the pore solution is:

aP Op\ pWQ ap (5.22)


at (&TpW
\\avI x
~ + \T (t V atJ
)

Then, using Eq. (5.21) in (5.18) yields:

dMw
= b-q5 (da* + dp) + # (!3wdp - adT)
bdEv + (b)
b (5.23)
= bd+Eb+ do-*
-
qac4dT+ + )dp
-

(
ks
1/N 1/M

where N is the solid Biot modulus, M is overall Biot modulus, and kw is the bulk
modulus of the pore solution. Finally, the state Eq. (5.9) can be rewritten in the

form:

d-*
dBm = (da-m(z)) = K. (dE, + (1 - B) ks + ac"'$dT - Bdd) (5.24)

where K, and B stand for the undrained bulk modulus and the Skempton coefficient,
respectively:
- b)(ks/kw - 1)# + b
= K + b2 M = k'(1 (5.25)
(ks/kw - ) + b

B BbM (
(1 -
bb )(ks/kwb - 1)# + b
(5.26)
Kv,

65
5.1.1 Poroviscoelasticity

It is the goal of this study to measure the bulk volume changes during the hydration
of CBM due to the evolution of eigenstresses under constant and controlled effec-
tive stress. Such controlled constant effective stress can be achieved in the rubber
diaphragm configuration (RDC) by means of a check valve. In the RDC, a small
effective stress (7-60 kPa) is required to maintain contact between the sample and
the rubber diaphragm. In the liquid barrier configuration (LBC), on the other hand,
such constant effective stress is not possible due to the radial confinement of the ves-
sel. The radial confinement of expansive samples creates a compressive radial stress
that can induce time dependent deformations. To account for the time dependent
deformations, we review here the relevant viscoporoelasticity relations.

When both elastic strain and viscous strain (creep) coexist in a continuum, the
total strain can be considered to be composed of two parts:

= de + dev (5.27)

where dEij is the total strain increment, dEq is the elastic strain increment, and
d,'. is the viscous strain increment The total strain inerement de 2 can bP fiirther

decomposed into deviatoric and volumetric strain components dEig = dei +-lde5ij;
and the viscous strain can also be decomposed into deviatoric and volumetric strain
components dE . = de- + ldcv6i. The viscous volumetric strain dev undergone by
the porous solid is attributed to the viscous volumetric strain of the solid phase de"
and the viscous change in porosity do'

dc' = (1 - )dev + ddv (5.28)

For incompressible solid phase dc = 0 and the volumetric strain of the porous solid
is due solely to viscous change in porosity d/v = dcv. In the general case where the
solid is compressible, the viscous change in porosity can be related to the viscous
volumetric strain #v = bcv, where b is the tangent Biot's coefficient.

66
When viscous strain is present in an elastic material, the elastic strain is the
difference between the total strain and the viscous strain. The incremental state of
stress including the viscous deformations can be written as

dur+ dp = K(dc - dc) + (1- b)(do-* + dp) (5.29)

dsij = 2G(deij - de') (5.30)

where am is the mean stress (hydrostatic stress) component and sij is the deviatoric
stress component. The rate of viscous deformations can then be related to the effective
stress and deviatoric stress through the volumetric viscous coefficient mqand the
deviatoric viscous coefficient r7d

Or + bp = KeV + yvi"v; sij = 2Ge' + 2


Tld% (5.31)

5.1.2 Useful relations and tools

Effective stress The principle of effective stress was introduced in 1923 by Karl
Von Terzaghi in which he related the deformation of the porous solid to the excess
in the total stress over the pore pressure. The excess in the total stress is called the
effective stress a- and here it will be referred to as Terzaghi's effective stress and it
is defined as [52]:
U' 1, (5.32)
m = kk

where UiJ is the total stress applied on the porous solid and p is the pore pressure. For
the derivation of the principle of effective stress, Terzaghi assumed an incompressible
solid matrix and as a consequence, an increment in the volumetric strain dE is mani-
fested entirely as incremental change of the porosity of the system, dE = do [53]. For
a porous solid with incompressible solid matrix, it is valid to relate the volumetric
strain increment to an increment of effective stress through the undrained bulk mod-
ulus K, a' = KE. However, most porous solid are composed of compressible solids
and hence, the compressibility of the solid matrix should be taken into account. To

67
account for the compressibility of the solid phase, Maurice Biot [21] introduced the

Biot coefficient b, which separates the part of volumetric strain due to change in

porosity be and the part of strain that is caused by the compressibility of the solid

phase (1 - b)E. Biot's effective stress is then defined as [52](p. 76):

1/ = + 1
=crj + bp3ij; afm - (5.33)

,
-1i -kk (.3

Biot's effective stress shows that volumetric strain is possible under zero "Terzaghi's
effective stress" by simply changing the pore pressure. The term effective stress in

this study will refer strictly to Terzaghi's effective stress because only such effective

stress is imposed and controlled experimentally.

Displacements, strains, and stresses The displacement field is a function of r

and z only and it is defined by the two components

u(r, z) = urer + uzez + useO (5.34)

where ur is the radial displacement and uz is the axial displacement. The circumfer-
ential displacement, uO, is zero due to the rotational symmetry.

The infinitesimal strain and stress tensors in cylindrical coordinate system can be
written as

Eij=

L
Err Erz ErO
Ezr Ezz

Ear EOz
EzO
1rr
I00asL
n
ij =

Due to axisymmetry, OrO, Ozo, ErO and Ezo vanish leaving only fon r components
-zr

GOr
Urz 0rO
Uzz

0Oz
0

UOO
-zO

I (5.35)

Err Erz 0 1[ rr (-rz 0


Eij = Erz Ezz 0 ;
0
-ii = Orz o-zz 0 (5.36)
0 0 oo 0 0 (Too

68
Rearranging the non-vanishing terms

Urr

Ezz 07 o-zz
(5.37)

2
Erz Urz

where each of the non-vanishing components is a function of r and z only.

Kinematic equations The radial strain in the r direction

Our
Err = er ' -= (5.38)
ar 4r

The circumferential strain in the 0 direction and considering axisymmetry

lBu l Bu9 ur _Ur


Eoo = eo -- =u
-- + - = - (5 .39)
r OO r aO r r

In the axial z direction, we get

Ou Ou(
,zz = ez - = (5 .40)
az 49z

Finally, the non-vanishing shear strain component in axisymmetric problems

Erz =- (er- Ou+ez - -- = -(r+ OUz(5 .41)


2 Rz r 2 z Ur

Rearranging in matrix form gives:

Err '9 0
0 a Ur
5z
(5.42)
1
eo 0 uz
a a
2rz (9z r

69
Incremental momentum balance (equilibrium equations) The general equa-
tions of equilibrium without body forces in cylindrical coordinates in incremental form
are

{
(dErr) + (dr) + (dErz) + (dErr - doo)= 0
div (dEij) = A (dEzr) +1 (dY zr) + i (dEzo) + a (dEzz) 0
(dEOr) + 1 (dY 00) + A (dEoz) + (dEZr) =0

For axisymmetric problem, these relations reduce to

div (dEij) =
{ (dErr) + A (dErz) +(dErr
" (dEzr) + '(d~zr) + a (dEzz)

For oedometric conditions, these relations can be further reduced to


- doo) = 0
= 0
(5.43)

(dEr) =ar0 (d oo) = 0


div (dE ) = ao(5.44)
7a (dEzz) = 0

Constitutive equations (for elastic solid without eigenstress) The isotropic

constitutive law can be written using index notation as:

6
=i 1+i j~- 7 -Ukk ij
(5.45)
ort iE arxfr E

or, written in matrix form

Err 1 -1 -V 0 07rr

1 -V 1 -V 0 ozz
(5.46)
01 0

0 0 0 2(1+ v7) Urz

And the stress is

E 3K- 2G
= E{ + -E k6i = 2G Ej (5.47)
6G
+

1-12v
+

70
or, written in matrix form

Urr 1-v V V 0 Err

-zz 3K V 1-I iV 0 Ezz


or = = (5.48)
0 (1+v)+I I-V0 900

Urz 0 0 0 (1- 2v) 2Erz

where E 9KG
3K+G and V 3K-2G
2(3K G)

Constitutive equations (for elastic solid with eigenstress) For an elastic


solid experiencing volume changes due to eigenstrains, the constitutive equations can
be extended to include such contribution as follows:

1+ v V
6
Eij =
E a,-
E-kk ij + E*6ij (5.49)

07 j=
1 + VE ij + 1 -Ekki -- (5.50)
I - 2v -O*i
Such contributions are included explicitly for a porous solid in Eq. (5.1) and will
be detailed later in the analysis of each of the experimental configurations and the
corresponding boundary conditions.

5.2 The Permeability and Volume of the reaction


Configuration (PVC)

The PVC test can be run in two modes of operation (hydraulic boundary conditions).
In the first mode, water under controlled pressure is injected through the top port
(port-1 in Figure 4-6) while the bottom port (port-3) is connected directly to a pres-
sure transducer. In the second mode, water is allowed access through top and bottom
ports at identical or separately-controlled pressures.
To estimate the pressure variation and the evolution of permeability during the
reaction, we employ the first mode where water is allowed access through the top

71
port and pressure is measured at the bottom surface. In this mode, the pressure at
the bottom surface of the specimen is initially identical to the controlled pressure at
the top surface and the pore pressure is constant throughout the entire specimen.
This situation will persist until the specimen develops significant strength (setting).
Once the slurry sets, pressure at the bottom surface will start dropping at a rate
proportional to the rate of the reaction, thickness, and permeability of the specimen.
The stress and strain boundary conditions, assuming full bonding between the sample
and the wall of the vessel, are shown in Figure 5-7(a).

5.2.1 Stress and displacement boundary conditions

The state of stress in the specimen depends on the boundary conditions and the
evolution of mechanical properties and bulk volume changes during the reaction.
Ajhen the slurry develops its mechanical properties, it also undergoes volume changes

during the early age of reaction. If constrained, these volume changes will lead to
the development of stresses in the reacting material. In the absence of bulk volume
changes, stress can develop inside the material by external loads only.

The PVC test starts by placing the slurry inside the pressure vessel and cover
it with the desired curing fluid (see Figure 5-1). Under saturated conditions, the
slurry is most likely to experience a net expansion and as the paste gains strength,
compressive stress will start to build up on the walls of the specimen (Err < 0). The
extent of the expansion and consequently the stress on the wall will depend on the
mix design and the chemical composition of the slurry as will be discussed later on.
Figure 5-7(a) shows the boundary conditions of the test where the displacement is
fully constrained at the vertical wall and the stress at the top and bottom surfaces

72
I p
Port-2 pt Port-i pt
J

-i

4
F

Curing fluid C uring fluid

Specimen C pecimen

Port-3
Pb

Pb

(a) (b)
Figure 5-1: Hydraulic boundary conditions of the PVC test. In the boundary condi-
tions shown in (a), pressure oi the top surface p, ald bottom surface pb are controlled
separately. These conditions allow the direct nieasureinent of permeability of set sain-
ples by applying a differential pressure across the samnple. Boundary Couditionls shown
in (b) are employed to measure the permeability by controlling the pressure on the
top surface and monitoring the evolution of the pressure pi) at the bottom surface. In
both boundary conditions. the volume of reaction is also lmeasured.

73
are prescribed. These boundary conditions read:

V(r, 9), z= 0 :uz = 0 (5.51)

V(z, 0), r= R :uz = ur =uo = 0 (5.52)

V(r, 0), z = -h/2 : EZ = -Pb (5.53)

V(r, 0), z = h/2 :E = -pt (5.54)

where PA is the water pressure delivered through the bottom port, pt is the pressure
delivered through the top port, and R is the radius of the sample. From axisymmetry,
one can show that Ezo = Ero = 0 and ozo = oro =0 and q, = q0 = 0.

Stresses and strains developed in the sample subjected to the boundary conditions
shown in Figure 5-7(a) are complex and it will not be analyzed here due to the non-
homogeneous nature of the stress and strain fields. These boundary conditions will
only be used for permeability measurements to guarantee complete bonding between
the sample and the walls of the vessel.

5.2.2 Volume of the reaction Vn (chemical shrinkage)

Volume of the reaction can be obtained in the PVC and all other saturated configu-
rations by simply measuring the volume or the flow rate of water injected inside the
vessel under constant pressure. As stated previously, compressibility of the cement
paste constituents in conjunction with its permeability will influence the chemical
shrinkage measurement and the extent of this influence will depend on the specimen
size. One way to minimize the effect of the sample size is to simply use a sample with
small thickness. This can also be achieved by introducing the curing water through
the top and bottom surfaces at the same time (Figure 5-1(a)). Volume of the reaction
can be measured in most of the test configurations with reasonable accuracy as long
as air bubbles are not introduced into the sample. Volume of the reaction is inde-
pendent of the stress and displacement boundary conditions as long as the sample is
maintained in a saturated state.

74
The hydraulic boundary conditions on aQ

V(r, 6), z = -h/2 q= qbot (5.55)

V(r, 6), z = h/2 q,= qtop (5.56)

V(r, z, 6): q, =o = 0 (5.57)

where qbOt and qtop are measured quantities. The breakup of the flow rate between
the top and bottom surface will be discussed later on, but the sum of the flow rate
through the top and bottom ports represent the volume of the reaction.

dVrxn Mc
qbot + qtop = dt Ac (5.58)

where Vxn is the volume of reaction measured in pL/gcement, M, is the initial mass of
cement powder in the paste, A 8 is the surface area of the sample in direct contact with
water, which also corresponds to the cross-sectional area of the cylindrical container.

In a situation where both bottom and top ports are connected to the same source
of pressurized water and in the absence of thermal gradients and when the density
of the slurry is identical to the curing fluid (water) and when the permeability of the
sample is spatially uniform and the head losses through the network are identical for
both ports, then and only then, the flow rates through top and bottom ports will be
identical. But that is obviously not the case. To elucidate the distribution of flow
rate between the top and bottom surfaces of the sample, we run two identical tests
with variable head losses through the flow network by swapping the connectivity of
the flow meters with different head losses as shown in Figure 4-3.

Both tests were carried out under identical conditions with water-to-cement ratio
of 0.42 and cured under 25'C and 0.2 MPa curing pressure in the PVC configuration.
The main outcome of the test is the volume of reaction represented by the black
lines in Figure 5-3. As expected, it was identical for both samples. An interesting
observation was the division of the flow rate between the top port representing the
water uptake of the sample through the top surface and the bottom port representing

75
the water uptake in the sample through the bottom surface, where flow rates are in
pL/g/hr. Flow rates through top and bottom ports were measured with two different
flow meters with slightly different head losses as measured experimentally and shown
in Figure 4-3.

In Test(1) the flow meter with the higher head loss is connected to the bottom
port, and the flow meter with the lower head loss is connected to the top port. Then
the flow meters were swapped for Test(2). During approximately the first 15 hours
of the test flow rates were controlled by the head losses in the external flow network
(flow meters) and after 30 hours of hydration the flow rates were controlled by the
permeability of the sample. It is evident from these tests that sedimentation creates
a permeability gradient along the sample.

Flow rate through the main inlet as shown in Figure 5-2 is Qtot and flow rates
through the top and bottom branches are Qt0 p and Qbot, respectively. The origin of
the water sink causing the flow through the network is the negative volume of the
reaction Vxn and from mass conservation:

dVrx
Qtot Qtop + Qbot = dtr Mc (5.59)
dt

where M, is the original mass of cement powder in the mix. Assuming that head
losses are only significant through the flow meters, pressure in the tubing upstream of
the flow meters is the pressure delivered by the syringe pump pa shown as blue tubes
in Figure 5-2. The pressure in the tubing downstream of the flow meters depends on
the head loss through each flow meter, which in turn depends on the flow rate and the
geometry of the capillary tubes of the flow meters. In other words, pressure on the
top surface of the sample is pt = pa - Apt shown as red tubes in Figure 5-2 and the
pressure at the bottom surface is PA = Pa - APb shown as green tubes in Figure 5-2,
where Apt and APb are head losses through the flow meters in the top and bottom
branches, respectively. When the permeability of the sample is sufficiently high, the
assembly is expected to perform as a parallel channel flow with identical pressure
drop through both flow meters. Using the head losses measured experimentally in

76
Figure 4-3 we can calculate the flow rates during the first 15 hours. In this case,

Qt, _ kbob (5.60)


( 2 bo k.top

where k1 p and k)ot represent the headlosses through the top and bottom flow meters.

Such test can also be used to measure the evolution of permeability and the gradient
there of, and it can be further improved by adding a resistive element such as a

capillary tube with known head loss to the flow network. To this end, this test is only

used to measure the volume of the reaction and to give a qualitative insight into the
impact of sedimentation on the evolution of permeability.

Syringe
Pump

Tubing
FM-I

/7

Qtop

Spec imen
Vessel
Qtot
z (t) Rigid Filter
Z(bot)
FM-2 (a) Z(a)

Qbot
Figure 5-2: A simplified schematic of a single-pump flow network. where the syringe
pum1p is a source of pressurized water. The filter is a saturated stack of filters that
allow the flow of water but not of cement grains, while allowing access of water to the
entire bottom surface of the specimen. FM-1 and FM-2 are flow meters that measure
the flow rate through the top branch Qjj, and the bottom branch Qb.t, respectivehy.
For a typical test, cement slurry is placed inside the vessel and the desired conditioiis
are applied. Flow meters FM\-1 and FM-2 have different head losses as shown in
Figure 4-3. The result of swapping the flow meters is shown in Figure 5-3.

77
9EN91L-

1.5 ---

-
-Qtot Test(1)
1.2 -Qtot Test(2)
Qbot (1315) Test(1)
Qtop (1302) Test(1)
Qbot (1302) Test(2)
- ----
3- 0.9--- - Qtop (1315) Test(2)
On

0.6- --

0.3
-

0- II I

0
0 10 20 30 40 50 60
Time (hr)

Figure 5-3: Flow rate through top and bottom ports and the effect of head losses
through the flow network. Two tests were carried out under identical conditions with
0
water-to-ceient ratio of 0.42 and cured under 25 C and 0.2 MPa curing pressure
in the PV configuration. The miain outcome of the test is the volume of reaction
represented v the black lines iin the Figure and as expected. it was identical for both
samples. Al interesting observation was the division of the flow rate between the
top port representing the water uptake of the sample through the top surface and
the bottoim port representing the water uptake in the sample through the bottom
surface. where flow rates are in pL/g/hr. Flow rates through top and bottol ports
were measured with two different flow meters with slightly different head losses as
measured experimentally and shown in Figure 4-3. In Test(1) the flow meter with
the higher head loss is connected to the bottom port and the flow meter with the
lower head loss is connected to the top port. Then the flow meters were swapped for
Test(2). During approximately the first 15 hours of the test flow rates were controlled
hours of
by the head losses in the external flow network (flow meters) and after 30
hydration the flow rates were controlled by the properties of the sample. It is evident
from these tests that sedimentation creates a perimieability gradient along the saimiple.
The number (1302) and (1315) arc used to refer to the different flow meters.

78
5.2.3 PVC test with bottom surface connected to the pres-
sure transducer only

Another application of the PVC test is the measurement of permeability during the

hydration of CBM. One way to measure permeability is by providing water through

the top surface of the sample under constant or cyclic pressure while the bottom

surface is connected directly to a pressure transducer. Pressure evolution at the

bottom surface is then measured and related to the permeability of the sample. The

following is an analysis to estimate the evolution of pressure and permeability during

the chemical reaction for the configuration shown in Figure 5-1(b) by application of

a constant pressure. For this analysis, we consider an incompressible flow through

an incompressible reactive porous solid (in the absence of bulk volume changes).

Boundary conditions and the state of stress in the specimen related to this analysis

are presented in Figure 5-7(a). It is worthwhile here to point out that the assumption

of incompressibility is based on the fact that the rate of the water sink (volume of the

reaction) and changes in permeability due to the progression of the chemical reaction

are orders of magnitude larger than any contribution from strains and change in

porosity due to mechanical loading (external stress).

To obtain the equations for the conservation of mass and momentum, we consider

the differential volume shown in Figure 5-6. Mass conservation in the axial direction

(there is no flux in the radial or hoop directions due to axisymmetry) requires that

the difference between the rate of mass flow in and the rate of mass flow out of the

control volume must be equal to the rate of accumulation inside the control volume.

The flux field will be described as

qj = q - ni (5.61)

where q is the flux vector of water and i = r, 0, z. The infinitesimal volume of the

REV is
dVREV = rdrd~dz (5.62)

79
For the purpose of this analysis, we will assume that all the constituents of the control

volume are incompressible and the accumulation is only due to the ongoing chemical

reaction. The rate at which the fluid mass (m) is lost or gained by the control volume

during the transient process can be written as:

p(Qin - Q att)
at at
rdrdOdz (5.63)

where Qi2 and Qout are flow rates of volume per unit time into and out of the REV.

p is the density of the fluid and q, qo, and q, are fluxes in the r, 0, and z directions

respectively. Eq. (5.63) defines ( as volume fraction of fluid stored or lost from

the REV due to chemical reactions or mechanical loading (dimensionless). In this

analysis, ( is assumed to be only due to chemical reactions. This is a valid assumption

if the pressure variation is small (more specifically, 2 and - are very small), or the

constituents are incompressible. Hence

frxn (5.64)

where frxr is the fractional volume of the reaction and is measured directly in the

experiment. The conservation of mass reads in explicit notation for cylindrical coor-

dinates:

I a(prqr)+ - (pqo) + a-(pqz) (pfrxn) (5.65)


r Or ro00 z at

Here we assume that the flow is one-dimensional in the z direction and qr = qo = 0

due to axisymmetry. For one-dimensional flow as shown in Figure 5-7(a) and constant

density, we obtain:
Oqz Ofrxn
- at (5.66)
Assuming that the pressure gradient is the only driving force for the flux and in the

absence of concentration and thermal gradients and assuming isotropic permeability,


the flux (velocity) is obtained from Darcy's law

K
q(t, x) = - Vp(t, x) (5.67)
P

80
where K(t) is the permeability, which evolves over time as a function of the degree

of hydration and the accompanied change in porosity. p is the viscosity of the liquid

and Vp is the pressure gradient. From Eq. (5.66) and (5.67), we obtain:

_V2p(t,x)= frx (5.68)


P at

where V 2 p(t, x) is the Laplacian of the pressure. In one-dimensional flow it is simply


02
the second derivative of the pressure P (pressure distribution along the specimen is

non-linear):
a 2 p(t,z) P Ofrxn (5.69)
az2 K (t) at
Integration yields:
ap(tZ) / A z + f(t) (5.70)
az K (t) 8t

p(t, z) = K (t) at Z2 + f(t)z + g(t) (5.71)

where f(t) and g(t) are functions to be determined from the boundary conditions.

In the test described in Figure 5-7(a), the boundary conditions are flow rates and

pressures at top and bottom of the specimen and the unknown is the permeability

K(t). It is also assumed here that pressure at the top and the bottom of the specimen
as well as the flux through the top and bottom surfaces vary over time. Viscosity of

the fluid is assumed to be constant but the permeability is evolving over time (or the

progression of the chemical reaction) The flux boundary conditions in Eq. (5.67) for

flux through the bottom surface reads:

h) K 4p(t, z=b
qz tz= ) 2) _qb (5.72)
2 y z

For flux through the top surface, we have:

tKapt
qt (5.73) qz t z 2 -
2) p az

81
Based on these boundarv conditions. the pressure gradient is

Up(t. z)
(5. 74)
U: M ot) K(t) ( 2

By substituting the pressure boundary conditions in Eq.(5.71). the pressure at the

bottom surface is
i (t) =
11 2 /1 + (5.75)
p t,
2 ,;(t) t 8 ,(t) \ 2 ) 2
)

In return, the pressure at the top surface is:

/1 oftr, h? - t ) + q, I
(I 1) (t. 2)
- At (5.76)

+
Ut 8 t;(t) ( 2 )2

The pressure (listribution thus is:


Pressure distribution

2I
-(q,
qb p , ( / ( p) - pt
(5.77)
+

x(t) Ot 2 8 x(t) 2 2

The perlleability can thus be estimated from:


Permeability in PVC test 1
K
_________________ ________

n~t) = -p
qt + (b
(A - A)
(5.78) j
)

If frxn is known from a separate measurement, then the flux through the bottom
surface can be calculated fron flux boundary conditions as follows:

(b = (I + 1? (5.79)
Ut

For the special case where flow is restricted at the bottom surface (valves 3 and 4

in Figure 4-1 are closed), pressure at the bottomi surface is (pressure at the sealed

surface):
2 2 1(
]) (t. :
2,) ]4Llb(_ 2 i + ]) (5.80)
=0
But, since the flux through the bottom surface is not completely restricted, we need

to consider the compressibility of the dead volume between the pressure transducer

and the bottom surface of the specimen.

The dead volume flux (qb)

A pressure-transfer medium (water) is required to measure the evolution of pres-

sure at the bottom of the specimen (Figure 5-1(b)). The pressure-transfer medium

has a finite volume (dead volume Vd), which in addition to the wetted parts, will
3
determine the compressibility of this dead volume / :

1 OVd
/ 3d (5.81)
Vd pb

Since the exact volume and properties of the dead volume are not known and cannot

be measured directly, the compressibility will be measured experimentally. The com-

pressibility of the dead volume is then defined as the increment of volume injected

inside the dead volume Vi as a result of change in pressure Pb:

fOd = 0 " (5.82)


Pb

To measure the compressibility of the dead volume, a cyclic pressure was applied

through the top and bottom ports of the pressure vessel with a low permeability

specimen inside the vessel. The specimen was hydrated for 10 days prior to testing and

the rate of pressure variation was high enough to prevent any possible contribution

from the compressibility of the specimen. The result is shown in Figure 5-8, that is:

#d - in - 505 x 10--4 m3 /Pa (5.83)

Pore pressure inside the reacting porous solid will drop as a consequence of the on-

going hydration reaction and the evolution of permeability as shown previously. The

fact that the dead volume is compressible and pore pressure inside the specimen is

dropping will cause a flux of the pressure-transfer medium (water) inside the speci-

men. Flux through the bottom surface (qb) can then be related to the compressibility

83
of the dead volume:

q- -2.4 x 10-" nm/s (5.84)


Aspdt dt A4,) (It

where A = 2.287 x 10- 3 m is the section of the specimen.

A sample differential pressure development and permeability in Class-G ceiment

with w/c = 0.35 cured at 25'C and thickness of 33.3 nun are shown ii Figure 5-4
and the resulting pressure profile inside the sample in the axial direction for 25 hour

intervals is shown in Figure 5-5. The test is carried out in the configuration shown in

Figure 5-1 with curing pressure of 400 kPa delivered through the top port.

300 1E-14

250
1E-15
__I [_

1E-16
200
1E-17
150
1E-18
100
1E-19

I
50 1E-20

0 1E-21
0 50 100 150 200 0 50 100 150 200
Time (hr) Time (hr)

(a) (b)

Figure 5-4: (a) Development of pressure differential across Class-G cement samiple
with w/c = 0.35 cured at 25'C. (b) evolution of permeability of the same sample as
alcllatedl with Eq. (5.78).

In a PVC test. a knowNi miiass of the slurry is placed inside the vessel and covered

with the desired curing liquid before pressurizing the vessel. The experimental pro-

cedures should be designed in such a way to iiminimize discrepancy and variations in


the mneasureiment. Some possible sources of error are:

1. The first possible source of error in this imeasureiment is the leakage through

the network tubing or the sealing umechanisum of the vessel. Leakage cal be
prevented by regular testing of the vessel and the nietwork and a quick check
0.50

0.30 --

0.10

-0.10

-0.30 -

-
-0.50
-

0.3 0.4 0.5 0.6 0.7 0.8 0.9 1


p/pt(kPa)

Figure 5-5: Development of pressure distribution inside Class-G cement sample in


the PVC test with constant pressure on the top surface using Eq. (5.77) with the
pressure differential and permeability shown in Figure 5-4.

of the network before running the test. To check for leakage. the vessel is

pressurized b.riefly without a specimen in a condition similar to that of the


PVC test while monitoring flow through the network.

2. The second possible source of error in the PVC test is the flow resulting from

temperature fluctuations. Temperature fluctuation of 1 'C is enough to create

significant volume changes and pressure variations. This problem is solved by

placing the entire pressure vessel inside a water bath.

3. Initial curing tenperature and initiation of the test is another source of error

in this test. It takes around 25 minutes to mix and place the slurry as well as

assembling the vessel to initiate the test. Depending on the test temperature,

the temperature equilibration can take up to 40 minutes. It takes about an

hour to start collecting meaningful data. If the test temperature was close to

the lab temperature. then this will not, create significant deviation but if the

test temperature was significantly different from the lab temperature then this

deviation should be taken into account. One way to eliminate this error is

b1y following strict test procedures to get reproducible results at similar test,

temperatures.

85
Mmmk

4. Heat of reaction and thermal runaway is another source of error in this test. The

goal of the PVC test is to measure the volume of the reaction and pernability

and hence, it, is imperative to keep the temperature of the entire system as

constant as practicably possible during the test. The system is innersed in

a water bath to control the teniperature, but temperature rise of about 0.8 C

was estimated inside the specimen during the test.

5. Sedimentation can play an important role in the measurement of permeabil-


ity. The well-known bleeding in slurries can create a deilsity and permeability

gradient along the specimen.

z UoO

Uzz(z=D) qz (z=h)2 UU
Uz

-0
r 4 urj~ UrTr 0> r U 9
--- - - -UrzR

Specimen -~-r+ Uz

h
\N

-) qz (z=-2)

(a) (b)
Uzz
z
Uz
AUzU

h--h --------- ----

Specimen
0 27

(C)
Figure 5-6: Stress and displacellent conventions.

86
z z
qz=O

qe=O
qr=O
u8 =0 r
Uz=O
ur=O

Cfzz=-pb qz =qb Cfzz=-pb qz* 0


uz=O
(a) (b)
z z
Cfzz=-p qz=O Cfzz=-p, qz=O

qr=O
qr=O u8 =0
ur=O Uz=O
u8 =0 r ur=O r
an=O

Cfzz=-pb
uz=O
(c) (d)
Figure 5-7: Boundary conditions and state of stress in all configurations. Fig. (a)
Shows the stress and displacement boundary conditions of the PVC test. Fig. (b)
shows the stress and displacement boundary conditions for the RDC test. Fig. (c)
shows the boundary conditions in the LBC test with greased wall. Fig. (d) shows
the boundary conditions in the LBC test with clean walls.

87
120

100

80

60

S40

20

0
0 500 1000 1500 2000 2500
Pressrue (kPa)

Figure 5-8: Volume and pressure variation for the dead volunne at the bottom of the
vessel. It can be seen that the respolise is nonlinear and can be due to the presence of
3 /Pa.
0-rings in the compressed part of the vessel. The average slope is 5.5 x 10-n

5.3 The Rubber Diaphragm Configuration (RDC)

The rubber diaphragm configuration can be used to simultaneously neasure the vol-
uie of the reaction 11', and the bulk volunie changes ERD under hydraostatic state

of stress and controlled effective stress. The superscript RD will be used hereafter

to refer to the bulk volume changes measured under Hydrostatic state of Stress and

will be later compared to volunme changes measured under oedoietric boundary coil-

ditions (OBC). The RDC test can also be used to measure the bulk volume changes

of sealed sanples and to imonitor the evolution of the pore pressure for such samples.

Figure 5-11 shows the two possible tests in the RDC for saturated and sealed samples.

In a tyl)ical saturated RDC test, the slurry is placed inside the rubber diaphragm

while the exposed surface of the sample is resting on a filter paper supported by the

perforated disk to allow flow of water through the entire bottom surface of the sample

(see Section 4.2.2). A controlled effective stress is then applied to the sample by means

of a check valve placed in the network to create an effective stress equal to the chosen
cracking pressure of the check valve (see Figure 4-1 and Section 4.1.5). Dimensions

and -eoinetry of Ihe rubber (iaphragn are chosen such that it can maintain the

intended shape while nimiilmizing any contributioii of permneability to the pore pressure

88
variation while providing a reasonable signal-to-noise ratio. The material and the
thickness of the rubber diaphragm are chosen to allow for free volume changes with
minimal confining stress except for the stress applied through the check valve.

Cracking pressure of the check valve is chosen to represent the desired conditions
with a range between 3 kPa and 60 kPa. The choice of the 60 kPa check valve is based
on the pressure drop in sealed samples (autogenous conditions). Pressure evolution
at the bottom surface of sealed samples is shown in Figure 5-9 for Class-G cement
samples in the rubber diaphragm test with w/c = 0.42 at 25'C and 300 kPa curing
pressure. Upon setting (about 4 hours), pore pressure starts dropping due to the
negative volume of the reaction (chemical shrinkage) until it reaches a low enough
pressure when it starts increasing again. The jump increase in pressure is most likely
due to the formation of gas bubbles inside the sample and the fittings of the pressure
vessel.

5.3.1 Boundary conditions

Coordinate system and stress/strain convention are shown in Figure 5-6 and the
boundary conditions are shown in Figure 5-7(b). The mechanical boundary conditions
are zero displacement at the bottom surface z = -h/2, where h is the thickness of
the sample and z is the axial coordinate. The boundary conditions read:

V(r,9 ), z = -h/2 : uz = 0 (5.85)


V(r, 0), z = -h/2 : ERD = IRD = ERD = 0 (5.86)
V(rO6), z = -h/2 : ERD -Pb (5.87)

V(r, ), z = h/2 : ERD = -Pt (5.88)


V(z, ), r = R: ERD = -pt (5.89)

where Pb is the water pressure delivered through the bottom port, pt is the pressure
delivered through the top port and exerted directly on the rubber diaphragm, and R

89
300
-

250
-

200

150
(iz

F-
0.
100 _-

50

0 - -H
- I -- 7

-50
-

-100
0 10 20 30 40 50
Time (hr)

Figure 5-9: Evolution of pore pressure as measured at the bottoin surface of sealed
Class-G cement saniples in the RDC test for three identical samples with u,/c = 0.42
at 25'C in the sealed configuration shown in Figure 5-11. Zero pressure corresponds
to atmnospheric pressure.

90

I
is the radius of the sample. The hydraulic boundary conditions on OQ

K ap d (VrxnMc dEvD Vsp


V(r,9), z=-h/2:qz=--z d- dt A (5.90)
p 9z dt Asp d ASP
V(r, 0), z = h/2: q, = 0 (5.91)

V(r, z, 0) :q, = O = 0 (5.92)

where Vxn is the volume of reaction measured in pL/gcement, M, is the initial mass

of cement powder in the paste, Asp is the exposed bottom surface area of the sample,

/- is the permeability of the sample, p is the viscosity of the pore solution, and Vsp is

the volume of the sample.

Assuming constant effective stress and negligible shear stresses at the bottom

surface of the sample, the radial displacement at the walls of the stress-free specimen

is

uRD
r = ED rD
3V RDz (5.93)
Z 3 v

And the Strains are

Ezz = Err = O= -ERD (5.94)


3v

Due to bulk volume changes, the initial radius of the specimen R becomes R'

R'= 1 + EV R (5.95)
3

And the initial height h becomes h'

ERD
h'= (+ E ) h (5.96)
3

5.3.2 Field equations

Within the sample, the two conservation laws to be satisfied are linear momentum

and mass conservation.

91
Incremental Momentum Balance

For axisymmetric problem, divergence of the incremental momentum reads

div (dEj4) =
{ r(dErr) +1 (dErr - dE oo ) = 0
( (dEzz) = 0

The total stress imposed on the sample in the RDC is the constant pressure applied
(5.97)

through the top port while the effective stress is applied through the check valve and

maintained constant thorough the experiment

ERD =RD _ ERD -Pt (5.98)

The incremental stress (see Section 5.1), reads:

dERD _ dERD _ dRD = KdERD + ( - b)du* - bdp = 0 (5.99)

The incremental effective stress in the RDC also vanishes; that is:

d(ERD + p) = KdERD + (1 - b)d(T* +p) = 0 (5.100)

This leads to the incremental volumetric strain in the RDC test with constant total

stress:

dERD = (bdp - (1 - b)du*) (5.101)


V K
1 d
B) ku Bdm
*

- (5.102)
K ((1 ks - B
)

The incrpmpnftal vonlulimptric ftrain in the PDC tst wifth onswtant ffIctfivp strps reads:

dERD =
V (- b)d(* + p) (5.103)
K

In summary, under constant pore pressure with both the total stress and the

effective stress maintained constant, the following holds:

92
END

Incremental volumetric strain in the RDC test with pore pressure, total stress,

and effective stress all maintained constant throughout the test

dER, RD = (1 - b) do-* = 1 dT* (5.104)

Viscous deformations (creep)

In the rubber diaPhiragni configuration, the effective stress is applied by means of a

check valve to maintain the contact between the rubber diaphragm and the sample.

The magnitude of the effective stress is controlled and maintained constant through-

out the test. Due to the nature of the cement slurry, the effective stress cannot be

applied until the setting of the slurry is reached. Once the slurry is set, the effective

stress will be applied gradually due to the progression of the hydration reaction and

the negative volume of reaction. Typical development of the effective stress applied

through the check valve is shown in Figure 5-10, where the onset of pressure drop

measured at the bottom port is affected by setting and consumption of bleed water

in the hydration reaction.

Pc (CGC-Full-RD-W38-T25)
-Pc (CGC-Full-RD-W38-T25)
-- Pc (CGC-FuII-RD-W38-T25)
1030
1020
1010
1000
) 990
980
970
960 - - ---
950
940 -
-

0 20 40 60 80 100
Time (hr)

Figure 5-10: The evolution of pressure at the botton surface of the sample during
hydration in the rubber diaphragm test, where the pr{'ssure drop is due to the check
valve. Sample CGC-RDC-01 is tested with a 14 kPa check valve while all the other
saimples are tested with 60 kPa check valve.

93
The application of the effective stress at very early age leads to irreversible vol-

umetric deformations in the sample due to the high compressibility. Any volume

changes measured after the sample develops significant strength to resist such defor-

mations, is attributed to the development of the eigenstresses in the sample. The

constant effective stress applied throughout the test is controlled by the cracking

pressure of the check valve pc, that is:

E' = Em+ p = -pr; dEm+dp= 0 (5.105)

The total stress is also maintained constant and equal the applied pressure (curing

pressure):

ERD = ZRD R DRD _DRD=O (5.106)


rr 00 zz pidr=d 0 z(516

Following the analysis presented in Section 5.1.1, and assuming that the small and

constant effective stress (7-6OkPa) cannot contribute to viscous deformations, we have

dc' = 0, and thus

dE RD - b) (du* + dp) (5.107)


K

5.4 The Liquid Barrier Configuration (LBC)

The use of a liquid barrier to measure bulk volume changes was applied independently

by Bosma et al. [27] [28] and Abuhaikal et al. [1] to measure the bulk volume changes of

hydrating cement-based materials. Bosma et al. did not specify the material that was

used as a liquid barrier, nor did they specify the required properties except for stating

that it was a hydrophobic liquid. The liquid of choice in the case of Abuhaikal et al.

was the Poly [methyl(3,3,3-trifluoropropyl)siloxane] Dow Corning FS-1265 chemically

inert liquid.

94
Port-2 pt Port-I Pt

-4

~1'
Hydraulic fluid

Specimen
2 Hydraulic fluid

Specimen
F
Rubber
Diaphragm

P
-

Port-3
b

Pb

(a) (b)
Figure 5-11: Hydraulic boundarv conditions of the RDC test. In the boundary condi-
tions shown in (a), pressure through the top port pl, and pressure on bottom surface
pl can be identical, controlled separately, or a differential pressure is applied through
a check valve. These conditions allow the direct measurement of bulk volume changes
under controlled effective stress. Boundary conditions shown in (b) are enployed to
measure the bulk volume changes of sealed sample while measuring the evolution of
pressure at the bottom surface Pb. Volume of the reaction is only measured in the
boundary conditions (a).

95
5.4.1 Boundary conditions

Coordinate system and stress/strain convention are shown in Figure 5-6 and the
boundary conditions are shown in Figure 5-7(c) and (d). Only the oedometric bound-
ary conditions of 5-7(c) will be analyzed here. The mechanical boundary conditions
are zero displacement at the bottom surface z = -h/2, where h is the thickness of
the sample and z is the axial coordinate. Neglecting the self-weight of the sample
and the barrier liquid, we have as boundary conditions:

V(r, 0), z = -h/2: = 0 (5.108)

V(r, 0), z = -h/2: ELB = -Pb (5.109)


V(r, 0), z = h/2: E = -pt (5.110)
V(z, 6), r = R:ELB = -pt (5.111)

where Pb is the water pressure delivered through the bottom port, pt is the pressure
delivered through the top port and exerted directly on the liquid barrier, and R is
the radius of the sample. The hydraulic boundary conditions on read:

K OP d (VrxnMc\ dELB Vss


ep 19 -h /9 -n = n, -
\Vf(r 09
~ = ,/1z
pb dt \ Asp,! dt Asp
V(r, 0), z = h/2: q = 0 (5.113)

V(r, z, 0) : = q0 = 0 (5.114)

where Vrxn is the volume of reaction measured in pL/gcement, Mc is the initial mass
of cement powder in the paste, Asp is the exposed bottom surface area of the sample,
K is the permeability of the sample, 1- is the viscosity of the pore solution, and Vsp is

the volume of the sample.

The liquid barrier functions in two possible ways. The first mechanism can be
achieved if the pore fluid (fluid initially present inside the porous solid) is much more
wetting than the liquid barrier and the interfacial tension between the liquids is high
enough to allow it to behave like the rubber membrane (Figure 5-12(b)). An excellent

96
example of such liquid is toxic mercury. The second mechanism can be activated if

the viscosity of the barrier liquid is very high such that the flow inside the specimen

is prohibited by the viscous forces. Following is a discussion of the mechanisms that

allow the use of such experimental setups to measure the bulk volume changes and

its limitations.

5.4.2 Wetting, interfacial tension and viscous forces

In the case when the pore fluid is more wetting than the liquid barrier (Fig. 5-

12(b)), ingression of the liquid barrier is limited by the pore size distribution of the

porous solid. The liquid will readily flow through if the pressure differential between

the two immiscible fluids away from the meniscus is sufficient to push the liquid

interface deeper into the porous solid. The rate of ingression depends on the pressure

differential, viscosity of the fluids and the pore size distribution. For the ingression

to proceed, the pressure differential should be greater than the capillary pressure:

2'-
Pt - Pw > Pc Pnw - Pw = -- cos(08 ) (5.115)
r

where pt is the pressure at distance z behind the meniscus in the liquid barrier, PA is

the pressure in the pore fluid in front of the meniscus, pc is the pressure jump across

the interface, Pnw is the pressure on the less wetting side of the meniscus, p, is the

pressure on the more wetting side of the meniscus, -y is the interfacial tension between

the two fluids, r is the pore radius, and 0, is the static wetting angle at the triple

point.

The liquid barrier will ingress, however, at a lower pressure or even against the

pressure differential if the barrier liquid is more wetting than the pore fluid (Figure

5-13(b)). In this case and since the pressure in the liquid barrier is always maintained

equal or higher than the pore pressure (Pt ;> Pb), the liquid barrier will continue to

ingress inside the porous solid until it replaces the pore fluid completely. Such fluid

should not be used as a liquid barrier.

To illustrate the principle of the liquid barrier, we consider a case where two

97
immiscible and perfectly wetting fluids flow through a capillary (Figure 5-14). In

a capillary tube, the equilibrium contact angle 0, results from equilibrium between

several forces as a direct consequence of the molecular interactions among the three

materials at the contact line [58] (Figure 5-14(a)). The equilibrium contact angle 0,
0
is different from the dynamic contact angle d when the interface between liquids is

set in motion (Figure 5-14(b)).

Once the interface is set in motion, the rate of displacement of the meniscus will

depend on the dynamic contact angle. A pressure differential Ap must be imposed

between the pore fluid in front of the meniscus pw and the oil at a distance z behind

the meniscus pt in order to produce a flow rate Q based on Hagen-Poiseuille equation

[58] [62]:
2
8QpLBx -y
Ap= - -cos(Od) (5.116)
rr r

where PLB is the viscosity of the liquid barrier. In the problem at hand, the viscosity

of liquid barrier ALB is 4 orders of magnitude larger than the viscosity of the pore

fluid [pf. Eq. (5.115) can then be modified for porous solids:

AP = P.
- Pt PLB 1LBz - pc
K
(5.117)

where q = -'Vp is the Darcy flux and K is the permeability of the porous solid.

Now consider a porous solid specimen with height h supported at the bottom

surface uz(z = 0) = 0 and a free top surface (Figure 5-15). The specimen is initially

saturated with the pore fluid only (Figure 5-15(a)). Flow is allowed only through the

top and bottom surfaces where the top surface is in direct contact with the liquid

barrier at pressure pt. The bottom surface is in contact with a fluid that has the

same properties as the pore fluid at pressure Pb. The vertical surface of the specimen

is free to slide vertically (greased).

Then consider a liquid barrier that has penetrated to a depth z inside a porous

solid with porosity 0 in a sample with cross sectional area A (Figure 5-15(b)). The

volume of the liquid barrier that crossed the surface of the porous solid (assuming

98
that the liquid barrier completely replaced the pore fluid) is

VLB = Ax (5.118)

And by definition, the flow rate QLB (measured in m 3 /s) is

QLB dVLB QLB (5.119)


dt A

Hence

QLB = d (5.120)
dt

Now using Eq. (5.117) and (5.120) with the pressure gradient given by Vp = Ap/z,

and Ap = p.(t) - pt in Figure 5-15(c) where p.(t) is the pressure in the pore fluid

at the wetting end of the meniscus. It is also possible to relate pw to the pressure at

the bottom of the sample PA using Eq. (5.80). pt is the pressure in the liquid barrier

at the top surface of the specimen. Then we have

dx K AP+
(5.121)
dt p x

Assuming Pc, Pt, K, # are constants, we have

xdx = (p. - pt + pe)dt (5.122)

Since z (t = 0) = 0, then

x(t) ( ] (Pw - Pt + PC)dt) (5.123)

The ingressed volume is then

VLB(t) = OAx(t) =5AQ J (pw - pt + pC)dt (5.124)

In case the pressure at the top and bottom surfaces of the sample are identical in

99
a liquid barrier test, the inequality (Pc > Pt - Pw) should be satisfied for the liquid
barrier to function.

5.4.3 State of stress and strain in the liquid barrier test

It is virtually impossible to obtain perfectly homogeneous fields in a CBM sample


during hydration except in autogenous conditions due to the presence of a water
sink. Homogeneous fields can be assumed if the permeability of the sample is high
enough or the dimensions of the sample are small enough to prevent the development
of a pressure gradient. For a reactive porous sample with thickness h where the
bottom surface is exposed to water at pressure PA and the flux is restricted at the top
surface, then the following relation must be satisfied to obtain homogeneous pressure
distribution (Pb a pt based on Eq. (5.80)):

h2 [l<fxn < 1
2
(5.125)
Pb K Ot

The viscosity of the pore fluid yt, the permeability of the porous solid K and the rate
of the water sink af-xn
at are material properties and cannot be manipulated for a given
material and curing conditions. The minimum thickness of the sample is limited by
the hardware to obtain a reasonable signal-to-noise ratio. That leaves out the bottom
pressure as the main variable that can be adjusted to obtain near homogeneous stress
and strain fields.
The liquid barrier test can be carried out under two different displacement and
stress boundary conditions as shown in Figure 5-7(c) and (d). Figure 5-7(d) shows
the constrained boundary conditions where the sample is fully bonded to the wall
of the vessel. These boundary conditions are not analyzed in this study due to the
unnecessarily complicated non-homogeneous stress and strain fields developed during
hydration.
The goal of this analysis is to find a relationship between the bulk volume changes
during the hydration of a reactive porous solid as measured from a free-of-stress sam-
ple and the volume changes in a radially confined sample. For the free-of-stress spec-

100
imen, the bulk volume changes were measured directly inside the rubber diaphragm

test under controlled (small) effective stress. The result will then be compared to

the bulk volume changes as measured in the liquid barrier test. To establish the

relationship between the bulk volume changes as measured from both boundary con-

ditions, we start with the analysis of the oedometric boundary conditions. The main

assumption here is the absence of shear stresses on the walls of the vessel, which was

achieved by greasing the walls of the vessel.

Following is the analysis of the state of stress in a cylindrical reactive porous solid

sample confined inside a rigid cylindrical container with smooth walls under drained

conditions. The stress Eij on the top and bottom surfaces of the specimen is the

effective stress E'.. The pore pressure inside the specimen is p and the flux through

the top and bottom surfaces are qt and qb, respectively. The pressure in the liquid on

the top surface pt and the pressure at the bottom surface PA are identical to the pore

pressure.

Figure 5-7(c) shows the boundary conditions of the liquid barrier test with greased

walls. These boundary conditions will be referred to as the Oedometric BC (OBC)

and the superscript LB will refer to the quantities measured under these conditions.

In case where the sample experiences a net expansive volumetric eigenstrains under

hydrostatic state of stress EvB and considering the axisymmetry of the problem at

hand, then for OBC:

6
Err = E00 = r9 = Ezo = Erz = 0; Ozz = -p (5.126)

Err = Eo =Ero Ezo Erz = 0; Ezz = -p (5.127)

The volumetric strain measured under OBC is manifested entirely as axial strain

101
ELB ELB; thus the stress and strain boundary conditions entail

dErr = dEcO = (K - 2G) ]E:; + (1 - b)du* - bdp


d('., = dl' = (K - 2G) dE 2 + (1 - b)(du* + dp
(5.128)

)
dZ = (K + 4G) dEz: +(1 - b)du* - bdp
dE'z = dE + dp = (K + 4 G) dE 2 + (1 - b)(du* + dp)

where the ilodullus (K + 4G) is referred to as the oedometric modulus [52](p. 76).

Unlike the rubber diaphragm test. the total and effective stresses in the hoop and

radial directions in the LBC are non-zero Iue to the radial confinement. Alternatively

stresses can be represented in terms of the undrained bulk modulus and the skenipton

coefficient as

{ dErr

dE
-

2 -=
dEoo
(KI +
(K - 2G)

G) dE 2 + K, ((1
dE 2 + Kl ((1

Hence. the strain increment is obtained for constant total stress E-


-- B)
- B)

- B;
1 0- B
(5.129)

constant and

dE = 0
dE- 3 = dEll - 1 4 ( b)dj* bdp] (5.130)
K + -G
And the incremental volumetric strain in the LBC test with constant effective stress

V, constant and( dE' = 0. is:

dELB = = - (1EB b)*(/ + dp) (5.131)


' K+ 4G

Thus, under constant pore pressure with both the total and effective stresses main-

talied constant the incremental volunetric strain in the LBC is:


The incremental volumetric strain with constant pore pressure and effective
StreSS

K d(T*
dEV 3 dE' LB -
P - 4b) da* = () (5.132)
K + fK+ iG k a

It is of special interest here, to compare the volumietric strains as measured for the

102
hydrostatic state of stress (RDC) and the oedometric boundary conditions (LBC).

We recall Eq. (5.104):

dERID J (1- b)d (5.133)


K

Then the ratio between the volumnletri(c strain as neaslred from the LBC relative

to the Volumietric strain in the RDC is


The ratio between the volinnetric strain as neasured from the LBC relative to

the Volumetric strain in the RDC

dEB K 1 + Iv
(5.134)
dE D K + 4G 3(1 -v

where v is the Poisson's ratio of the porous solid.

The deviatoric strain in the liquid barrier test is:

deg = d 1- 1-dcoij = dcez X. e- 1


- -deij (5.135)
3 3

where the volumetric strain d6 is manifested completely as axial strain dg,. Thus,

1 2
derr = d o - = d dc., = -dc (5.136)
3 3

The incremental (leviatoric stresses are:

1
dij = dcr-I do-,s1 = 2Gei (5.137)
3

Thus.
2 4
dsr= ds-- = Gdc: dszz =-IGde (5.138)
3 3

Viscous deformations (creep)

Bulk volumne changes in the liquid barrier configuration gives rise to radial and hoop

stresses in the sample (ie to the radial confinement of the pressure vessel. Such

stresses can lead to elastic and viscoelastic deforinations in the confined expand-

ing/shrinking sanple. To take such viscoelastic deformations into account, we con-

103
sider here the viscoelastic deformations of the sample in addition to the elastic stresses
and strains developed in the preceding sections. It is possible in the LBC to maintain
constant the total axial stress or the axial effective stress or both.
As shown in Section 5.1.1, the viscous strains can be decomposed into deviatoric
viscous strains e - leading to the distortion of the sample and volumetric viscous
strains 6' leading to volume changes of the sample dE' = de,- + !dcvojj. Following
the elastic analysis in the previous section and including the viscous deformations
lead to the deviatoric stresses:

do-, = K(dc - dc') + (1 - b)du* - bdp (5.139)


dum+dp K(dE-dc') + (1- b)(da* +dp) (5.140)

dsi= 2G(deij - de') (5.141)

ds ; = 2G (deii - de
)

dsrr =2G (-dc - de)


r
3 (5.142)
dsoo = 2G (-1dE - devo)
dszz = 2G (idE - devz)

douij =dam 3dsij


+

dar, = K(dc - dWV) + (1 - b)do-* - bdp - 2G (Idc + de ,)


3 rr (5.143)
do-oo = K(de - de) + (1 - b)do-* - bdp - 2G (Ide + de o(
do-zz = K(de - dWV) + (1 - b)do-* - bdp + 2G ( dc - devz)

In the liquid barrier test where the effective stress is maintained constant during the
to +b effe ivn stress increomen+ in th1 -ir" tion vhAnis

do-zz+dp= K(dE-d-W)+ (1 -b)(do* +dp)+2G (d-de z) =0 (5.144)

The measurable strain increment in the z-direction is due to the eigenstress buildup
and due to volumetric and deviatoric creep under constant effective stress in the

104
z-lirection:
I
1
d - ((1 - b)(dr* + (11)) - Kdc" - 2Gdc"z) (5.145)
(K+ 'G)
v+1)
_(V+
d/-*+dp __+1_ ,
(de'
1 -2v

,
=- (0*+II+(V + I drl (5.146)

+
3(1- ) k3 3(1- v)

)
If the total stress is maintained constant during the test, the stress increment in

the z-direction vanishes:

do(- = K(de - dc") + (1 - b)du* - bdp + 2G (dE


- de = 0 (5.147)

The measurable strain increment in the z-direction in the LBC with constant total

stress reads:

1
b)du* + bdp - Kd" - 2Gdlc (5.148)

)
(K + 'G)

When both total and effective stress are maintained constant, we thus have:

Incremental volumetric strain in the LBC test with constant total and effective

stresses in the axial direction

1
(K + 4{G) (1 - b)du* - Kdc"' 2Gde, } (5.149)
:K+3G
Eigenstrain volumderic creep deviatoric creep

I/ + /(*+ d+ (5.150)
) 3(1- )
+

3(11 k d 1- )v -"

All A2 A 3
I

Eq. (5.150) shows that the measured volunetric strain can be decomposed into

the contribution from eigenstrain. volumetric creep, and deviator creep in the axial

direction. In Eq. (5.150), the coefficients A1 1 and M2 range from 1/3 to 1 for Poissoni's

ratio of 0 to 0.5. respectively and the coefficient Al1 ranges from 0 to 1 for Poissoi s
ratio of 0.5 to 0. respectively.

In case of expansion. the eigenstress is negative d&* < 0 and the contribution of the

eigenstrain to dEl ranges from 1/3 to 1.0 of the volumetric strain in the RDC, dEl'

105
as also shown by Eq. (5.134). In expanding samples and due to the radial confinement
dc' < 0. The radial strain due to the confinement and the resulting creep will further

reduce (relax) the radial stress and the contribution of the elastic deformations to the
measured axial strain and reduce the contribution of the eigenstrain.
Considering that de', = de'O and that tr(dsij) = 0, the deviator stress can be
rewritten as:
dsij = d-ij - dam
dSrr = d-rr-d-m = dm+ dp)
2 (5.151)
dsoo = duoo - d-m = (d-m + dp)
dszz = d-zz - dm = -(d-m + dp)

The effective radial stress in the liquid barrier test is:

darr + dp = da-m + dSrr + dp = -(dar- + dp) (5.152)


2

The deviatoric viscous strains become

de'. deii - dsi-


de' = -(dam + dp) - 1dE
rr 3 (5.153)
deoo = -(dorm + dp) - lde

[ dev = 3(d-m + dp) + qdE

The volumetric strain measured in the liquid barrier test can then be written in terms
of the eigenstress and volumetric creep:

dELB V _ _ * (5.154)
K

5.5 Summary

A microporomechanics framework is developed to quantify the state of stress in each


of the experimental configurations while taking into account the effect of the eigen-
strains. The resulting state equations are then employed to estimate the state of

106
!a) Hydraulic fluid (b) Liquid barrier

t
Rubber
membrane
pw

Solid 0 Solid

Figure 5-12: The concept of the liquid barrier in comparison with the rubber meni-
brane. (a) Is a typical situation in a tri-axial test where the pressure differential
between the hydraulic fluid and the pore fluid (p, - pb) is resisted by the deforination
of the rubber membrane and transferred to the solid skeleton as the effective stress.
(b) Will function exactly as in (a) if the pore fluid is more wetting than the liquid
barrier and the interfacial tension between the liquid barrier and the pore fluid can
resist the pressure differential. In (b) if the differential pressure across the interface is
larger than the capillary pressure, then the flow will commence and the liquid barrier
will ingress inside the specimen.

(a) Liquid barrier (b) Liquid barrier

Pt

Solid | Solid Z

Figure 5-13: Behavior of the liquid barrier is dependent upon the wetting properties
of the liquids. (a) If the liquid barrier is less wetting than the pore solution, capillary
pressure will act against the ingression. (b) If the liquid barrier is more wetting than
the pore solution, capillary pressure will enhance the ingression.

I
107
(a) (b)
1 - - - h I h--

oil porefluid r r

P P, P /" -d ) z

-
p p)
P,

Pb=Pi4- --------------
~-- z
z Pb --------------------------

Figure 5-14: Capillary flow. (a) In a capillary tube the equilibrium contact angle
0, results from equilibrium between several forces as a direct consequence of the
molecular interactions among the three materials at the contact line. (b) The dynanic
contact angle 0 1 when the interface between liquids is set in motion.

Liquid barrier pt Liquid barrier Pt p" p"..

,
z 7:1
2)
rfA

Pore fluid P
Pore fluid P
vz
(a) (b) (C)
Figure 5-15: Axial effective stress and the ingression of the liquid barrier iln porous
solid with height 1. (a) is the initial condition where the specimen is saturated
with the pore fluid only. (b) The specinen is exposed to a viscous liquid (the liquid
barrier) from the top at pressure P, and a less viscous liquid from the bottomn at
)ressuir'e p)I). (c) shows the pore pressure aid the effective stress along the specimen.
The displacement is restrained at the bottom surface (u. (z = 0) = 0) 1)but fluid camn
move freely in or out of the porus solid through he top and bottomll surfaces om1ly.
*

108
stress for the oedometric and hydrostatic boundary conditions. This chapter also
presented the calculation of the evolution of permeability of hydrating CBM in the
PVC configuration as shown in Eq. (5.78). Another important finding in this chapter
is the relationship between the bulk volume changes measured in the RDC and LBC
configurations as described in Eq. (5.134).

109
Part III

Results

110
Chapter 6

Results

This chapter presents results of hydration kinetics measurements from isothermal

calorimetry, volume of the reaction, and TGA for all investigated materials. For

each of the investigated materials, kinetics are then used to calculate the evolution of

volume fractions of the various phases during the hydration reaction as a necessary

input quantity for the multiscale microporomechanics model. Once the kinetics and

volume fractions are established for each material, the results of the bulk volume

change measurements are then presented. Nomenclature presented in Figure 6-1 is

used for all the results presented in this chapter and the following chapters.

6.1 Magnesium Oxide

Magnesium oxide (MgO) is investigated in this study due its simplicity with well-

known chemical reaction and properties of the both the reactants and the hydration

products. Due to its simplicity, MgO is first used to validate the experimental mea-

surements in the pressure vessel and to help understand the mechanisms of expansion.

Unlike the hydration products in cement, the hydration product of MgO (Mg(OH) 2
)

is an electro-neutral crystalline phase, which helps in eliminating any contribution

of colloidal forces (discussed in chapter 8). The resemblance between the hydration

kinetics between MgO and cement will also help understand the kinetics-controlling

mechanisms in cement-based materials.

111
uple thickness (mm1)
Curing pressure (MPa
Curing lemperature (fC)

CGC-F -RD-W42-T25-PJ-H25
Test configuration

Material

Figure 6-1: The nomenclature used throughout this study. The material name can be
CGC for Class-G cement, MgO for magnesium oxide, C 3 S. or CGS. The Particle size
can be F for fine, M for Medium. G for coarse C 3 S or coarse C 2 S, F1 for fine Class-G
obtained by jet sieving, F2 for fine obtained by sedimentation, CI for coarse CGC
obtained by jet sieving. C2 for coarse CGC obtained by sedimentation, and Full for
CGC with full particle size distribution. Test configuration can be RD for the rubber
diaphragm configuration, LB for the liquid barrier configuration, and PV for the
permeability and volume of the reaction configuration. w/c is for the mixing water-
to-solid mass ratio. T for curing temperature in 'C and P for the curing pressure in
lPa and H is for the thickness of the specimen in nun.

The material investigated in this study is light-burn MgO (MagChem@30) with


2
a surface area of 25 m /g. Samples are prepared with water-to-solid mass ratio

(w/s) of 1.0 and lower w/s ratios are prepared by squeezing of water out of the

sample under controlled effective stress. Bulk volume changes of MgO were also

studied under different boundary conditions; including hydrostatic state of stress in

the RDC configuration, Oedonmetric boundary conditions in the LBC configuration.

and constrained in the LBC configuration as well.

The hydration kinetics of Mg() was monitored for two different batches of MagChem@R 30.

R esults are shown in Figure 6-2. Another method to monitor the hydration kinetics is

through the volume of reaction. Figure 6-3 shows the volume of reaction as measured

in the pressure cell at 25'C

Bulk volume changes during the hydration of IgO were measured for a mnumber

of samples under different boundary conditions as listed in Table 3.2. In this section.

results of the bulk volume cliange measurements are first presented for the RDC and

LBC to compare the hydrostatic state of stress to oedometric state of stress. Figure

6-4 shows the complete bulk volume change curves of the samples tested in the RDC

with different w/s ratios. Due to the high surface area of the magnesium oxide of

112
12
IrL 900
800
10
- 700
8 - 600
/ 0-, dHr n/dt (n W/g) (MagChem30-B2) - 500
6 -- dfHrxn/,dt ( W/g) (MagChem30-BI)
-e - 400
Hrxni (J/g) (MagC ien30-B1)
4 -_ - - -Hrii (J/g) (I agC1 fei l3O-B2) - 300
- 200
2
-e -- I - 100
0 -A
-0
0 20 40 60 80 100 120 140 160
Time (hr)

Figure 6-2: Heat of hydration as measured with isothermal conduction calorimetry on


MagChem@30 with water-to-solid ratio of 1.0 at 25 C. The total heat of hydration
was 860J/g with the peak of the reaction rate occurring at 10 hr.

6 120

5 100
Vrxn/dt (MgO-PVC 01)
-

4- 4Vrxn/4dtMgO-LBCUO3) 80 -

dVrxn/dt (MgO-LBC-06)
3- - - ------------ 4Vrxn/dt-(Mg-RDC-07) 60
---- Vrxn (MgO-PVC-01)
2- ----- rxn (MNgO-LBC-03) 40
_rxn (MgQ-LBC-06)
1- 20
'4 ----- Vrxn (MgO-RDC-07)
I-

0- 0
0 20 40 60 80 100 120 140
Time (hr)

Figure 6-3: Volume of reaction as measured in the pressure cell on MagChei 30


with water-to-solid ratio of 1.0 at 25 'C. The total measured volume of reaction was
about 110PL/g with the peak of the reaction rate occurring at 10 hr.

I
113
about 25 m 2 /g, it was only possible to mix the slurry at w/s = 1.0. The w/s ratio
was then reduced by subjecting the samples to a differential pressure (effective stress)
of 10 MPa inside the rubber diaphragm test and squeezing the water out for different
periods of time. The initial fast expansion observed in the lower water-to-solid ratio
(w/s) is time-dependent deformation due to removal of the effective stress. The initial
fast shrinkage of the high w/s ratios is due to sedimentation (bleeding) and setting of
the solid framework. Both the high and low w/s samples displayed similar expansion
behavior.
Figure 6-5 shows the bulk volume changes during the hydration of MgO in both
RDC and LBC initialized at a degree of hydration = 0.5 (see appendix A for
definitions) for comparison. The hydrostatic state of stress is obtained with the
rubber diaphragm configuration with controlled effective stress while the oedometric
and constrained boundary conditions are obtained with the liquid barrier test. All the
samples presented in this figure have an initial water-to-solid mass ratio of 1.0 except
for one sample at 0.71 as seen in the legend. Details of the boundary conditions are
presented in Sections 5.2, 5.3, and 5.4. The symbol o- represents the effective stress
in kPa as applied by the check valve in the RDC test. Figure 6-6 shows the effect of
w/s ratio on the measured bulk volume changes in the RDC test.
To study the microstructure of MgO and the evolution of mechanical properties,
a microindentation investigation was conducted on a set of samples with different
w/s ratios at full hydration. Table 6.1 shows the details of the samples and the a
summary of the indentation results on samples pressed inside the rubber diaphragm
and in a piston apparatus.

114
---- - Ev(I) (M gO-RD-W100-a3) Ev (() (MgO-RD-W100-a3)
- Ev ()(IV gO-RD-W71-u60) Ev (() (MgO-RD-W53-a60)
Ev (N)(M gO-RD-W50-u14)
40000

30000

20000
-

10000

0 i i i i

-10000
-

N'

-20000
0 0.2 0.4 0.6 0.8 1

Figure 6-4: The complete bulk volume change curves of the samples tested in the RDC
with different w,s ratios il function of hydration degree . Due to the high surface
area of the magnesium oxide of about 25 m2/g, it was only possible to mix the slurry at
vw/s = 1.0. Tie u/s ratio was then reduced by subjecting the samples to a differential
pressure (effective stress) of 10 MPa inside the rubber diaphragm test and squeezing
the water out for different periods of time. The initial fast expansion observed in the
lower tv/s is tinie dependent deformation due to removal of the effective stress. The
initial fast shrinkage of the high w/s ratios is due to sedimentation (bleeding) and
setting of the solid framework. Both the high and low w/s samples displayed similar
expansion behavior.

115

I
-- Wm -- - mm - MFEINEOPMw - -- - - - - -Npmpmmr.EIL- - - - - --- - -=q

- Ev (() (MgO-RD-W71-o60) -- Ev (() (MgO-RD-W100-o3)


-Ev (() (MgO-RD-W100-o3) Ev (() (MgO-LB-W100)
Ev (() (MgO-LB-W100) Ev (() (MgO-LB-W100)
8000

7000
-

6000

5000

3 4000

3000-

2000

1000

0
0.5 0.6 0.7 0.8 0.9 1

Figure 6-5: Bulk volume changes during the hydration of 1\gO in both RDC and
LBC where details of the samples are listed in Table 3.2. The hydrostatic state of
stress is obtained with the rubber diaphragmn configuration with controlled effective
stress wile the oedoietric and constrained boundary conditions are obtained with
the liquid barrier test. All the samples presented in this figure have an initial water-
to-solidi mass ratio of 1.0 except for one sample prepared at w/s = 0.71. Details of
the boundary conditions are presented in Sections 5.2, 5.3. and 5.4. The symbol u-
represents the effective stress in kPa as applied by the check valve in the R.DC test.

116
Ev () (Mg-RD-W100-a3) - Ev () (MgO-RD-W71-o60)
Ev () (MgO-RD-W53-u60) - Ev (() (MgO-RD-W50-o14)

9000 I
8000
-

7000
-01-
-

6000
-

5000
-

-l
4000
-
-

3000
-

2000
-

1000
-

0
0.5 0.6 0.7 0.8 0.9 1

Figure 6-6: Effect of water-to-solid mass ratio on the measured bulk volume changes
during the hydration of MgO in the RDC test. For this test, samples of MgO are
mixed with water at water-to-solid ratio of 1. then part of the water is squeezed out
of the sample while inside the vessel by applying a controlled effective stress through
the rubber diaphragm.

117
Table 6.1: Hydrated Mg(OH) 2 samples for the indentation testing. Summary of the indentation results on samples pressed
inside the rubber diaphragm and in a piston apparatus. Samples are initially mixed at w/s = 1.0 and an amount of water is
squeezed out of the samples by applying a controlled effective stress in the rubber diaphragm test and in the piston apparatus.
After pressing, samples are allowed to hydrate under a small effective stress in a carbon-free environment until full hydration is
obtained. See chapter (7) for the complete discussion.

Sample MgO-RD-07 MgO-RD-08 MgO-RD-09 MgO-Pis-01 MgO-Pis-02 MgO-Pis-03


Initial water-to-solid ratio 1.1 1.1 1 1.2 1.2 1.2
Final water-to-solid ratio 0.71 0.53 0.5 0.53 0.77 0.6
Compaction Effective Stress (MPa) 1 10 10 15 0.1 3.5
Age at compaction (min) 10 10 10 5 5 5
Duration of compaction 5min 5min 5min lhr+const. vol. 5min+const. vol. 5min+const. vol.
Measured porosity (at 105C) 0.381+0.004 0.26410.002 0.248+0.010 0.222 0.005

-
Maximum Indentation load (mN) 2000 5000 5000 5000 2000 2000
Contact depth hc (pm) 19.9 0.6 17.9 0.3 18.0 0.2 15.7 0.3 16.5 0.7 12.4 0.4
Maximum depth hm (ytm) 20.8 0.5 19.2 0.2 19.2 0.2 16.9 0.3 17.5 0.7 13.2+0.4
Indentation hardness (GPa) 0.21 0.01 0.63 0.02 0.63 0.01 0.83 0.03 0.31 0.02 0.53 0.03
00
Indentation modulus (GPa) 14.7 0.4 29.6 0.5 29.3 0.4 35.4 0.6 18.6 0.9 27.4 0.9
Calculated porosity (1-r7) 0.396 0.003 0.292 0.004 0.293 0.003 0.248 0.004 0.369 0.007 0.307 0.006
6.2 Monoclinic C 3 S

The second material of choice in this study was the monoclinic C 3 S with three dif-
ferent particle size distributions. C 3 S reacts through a well know chemical reaction
and produces both calcium hydroxide and calcium silicate hydrate (C-S-H) and it
forms the main phase in cement-based materials and has major contribution to both
the enthalpy and volume of the reaction. These properties make C 3 S an excellent
candidate to study the bulk volume changes in CBM.

Bulk volume changes of C3 S pastes were studies for different particle size distri-
butions (PSD) and curing temperatures as well as w/s ratios. Different PSD are
studied to investigate the possible contribution of reaction mechanism on the bulk
volume change resulting from the diffusion-limited reactions. Different curing temper-
atures are investigated to understand the effect of hydration rate on the bulk volume
changes. Different w/s ratios are also investigated to elucidate the effect of porosity
and microstructure on the bulk volume changes.

Hydration kinetics of C 3S is monitored with enthalpy of reaction with isother-

mal calorimetry and volume of reaction in the pressure vessel. The enthalpy of the
reaction is measured with a Multi-Cell Differential Scanning Calorimeter (TA IN-
STRUMENTS MC DSC). Enthalpy is measured for C 3 S with three different particle
size distributions (see Section 3.2) mixed at water-to-solid ratio of 0.42 and cured at
25 C. Volume of the reaction is measured in the pressure vessel for C 3 5 with three
different particle size distributions (fine F, medium M and coarse G) with w/c ratios
of 0.49 for the fine C 3 S, 0.42 for the medium and 0.34 for the coarse.

Rates of enthalpy and volume of reaction in function of time are reported in Figure
6-7. Rates of enthalpy and volume of reaction in function of the degree of hydration
are reported in Figure 6-8. Degree of hydration is calculated assuming an enthalpy
of reaction of 520 J/g for the complete hydration of C 3 S. The degree of hydration as
measured from the volume of reaction is also based on the total enthalpy of 520 J/g,
which corresponds to 60.0 2.0 J/pL as shown in Figure 8-2(a). Figure 6-9 shows the
cumulative enthalpy and volume of the reaction for C 3S with the different particle

119
sizes.

Bulk volume changes during the hydration of C:3S with different particle sizes is

measured in the pressure vessel with the rubber diaphragim configuration (RDC) (see

Section 5.3 for the analysis of the RDC). Effective stress on the hydrating samples

is maintained constant by means of the check valve at 60 kPa throughout the test.

The results of bulk vohume change measurements in function of time and degree of

hydration for the different particle sizes is shown in Figure 6-22b.

Similar to the investigation of the different particle sizes., the effect of curing

temperature on the hydration of C 3 S with medium particle size is investigated with

the enthalpy and volume of the reaction. The rate of entlhalpy and volume of the

reaction for the C3S with medium particle sizes is shown in Figure 6-11 and in function

of the degree of hydration in Figure 6-12. The cumulative enthalpy and volume of

reaction are shown in Figure 6-13. Bulk volume changes during the hydration of C3 S

hydrated at 25. 35. and 45'C are shown in Figure 6-14.

dHrxn/dt (mW/g) (C3S-M-W42-T25) dVrxn/dt (C3S-M-RD-W42-T25-P1)

-dHrxn/dt (mnW/g) (C3S-G-W42-T25) dVrxn/dt (C3S-G-RD-W34-T25-P1)

-dHrxn/dt (mnW/g) (C3S-F-W42-T25) -- dVrxn/dt (C3S-F-RD-W49-T25-Pl)

7 3

6 2.5

-.- 5--- -_ 2__- -


-

0Z0
3 _._
-

1'10 100 1 1 0
Te(r)Time (hr)

(a) (b)

Figure 6-7: Rate of enthalpy and volume of the reaction as a function of tinie for
0
C 3 5 with three different particle sizes cured at 25 C. Enthaipy is iieaslired with
isothermal calorimetry aild volume of the reaction is measured iii thme pressure vessel.
See Figure 6-1 for the nomnclature.

120
dHrxn/dt (mW/g) (C3S-M-W42-T25) dVrxn/dt (C3S-M-RD-W42-T25-P1)
- dHrxn/dt (mW/g) (C3S-G-W42-T25) dVrxn/dt (C3S-F-RD-W49-T25-P1)
dHrxn/dt (rnW/g) (C3S-F-W42-T25) - dVrxn/dt (C3S-G-RD-W34-T25-P1)
8
7
2.5
bfl
6
CU 2
E 5
4 1.5
3 -I- - - _____ Z 1
2
1
0.5 - - -- 4-0--
- -.---
0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8 1

(a) (b)

Figure 6-8: Enthalpy (a) and volume of the reaction (b) as a function of progression
of the hydration reaction. Enthalpy of hydration as measured with isothermal con-
duction calorimetry on C:3S with three different particle size distributions. Volume of
reaction is neasured with the pressure vessel and the degree of hydration is calculated
assumilg an enthalpy of reaction of 520 J/g for the complete hydration of C:3S.

Hrxn (J/g) (C3S-M-W42-T25) Vrxn (C3S-M-RD-W42-T25-P1)


- Hrxn (J/g) (C3S-G-W42-T25) --- Vrxn (C3S-F-RD-W49-T25-P1)
Hrxn (J/g) (C3S-F-W42-T25) -Vrxn (C3S-G-RD-W34-T25-P1)
400 50

350 45
40
300 -t
35
250 30
200 25
-

150
100
20
15 -.
10
50 5
0 0
0 50 100 150 200 0 100 200 300 400
Time (hr) Time (hr)

(a) (b)

Figure 6-9: Ciiiiiiilative enthalpy and volume of the reaction as neasured with isother-
mal conduction calorimetry on C 3 S with three(differeIt partidle sizes culred at 25 0 C.

121
Ev (C3S-F-RD-W49-T25-P1) --- Ev (C3S-F-RD-W49-T25-Pl)
Ev (C3S-F-RD-W49-T25-PI) - Ev (C3S-F-RD-W49-T25-Pl) -
Ev (C3S-M-RD-W42-T25-P1) Ev (C3S-M-RD-W42-T25-P1)
Ev (C3S-M-RD-W42-T25-P1) (C3S-M-RD-W42-T25-P1) -Ev
Ev (C3S-G-RD-W34-T25-P1)
800
700
V Ev (C3S-G-RD-W34-T25-P1)
800
700
600 - -v 600
500 500
-/
400 400
.2- -0

-
300 300
200 200 -- - ~
- -- -

-
100 100

0 0
0 200 400 600 0.2 0.4 0.6 0.8 1

Time (hr)

(a) (b)

Figure 6-10: Bulk volume changes during the hydration of C3S at 25'C. Fig. (a)
shows the volumetric strain as a function of time. Fig. (b) Shows the volumnetric
strain as a function of the degree of hydration.

I
dHrxn/dt (mW/g) (C3S-M-W42-T25) - dVrxn/dt (C3S-M-RD-W42-T25-P1)
dHrxn/dt (mW/g) (C3S-M-W42-T35) - dVrxn/dt (C3S-M-RD-W42-T35-P1)
dHrxn/dt (mW/g) (C3S-M-W42-T45) dVrxn/dt (C3S-M-RD-W42-T45-P1)
12 4.5
4
10 3.5 ---
-- -- 3 7
-

8
E 2.5
6 -7- - - - -- 2
-

4
1.5 --
----- - - 1
-

2- 0.5
0
0- 10 100
1 10 100 Time(hr)
Time (hr)

(a) (b)

Figure 6-11: 1ate of enthalpy and volume of the reaction as a function of time for
0
CmS with medium particle size cured at 25 C. 3 5C. 45'C. Enthalpy is mlleasulred with
isothermal calorimetry anid volume of the reaction is leasure1d in the pressure vessel.
See Figure 6-1 for the nomenclature.

122

____ _ I
- dHrxn/dt (mW/g) (C3S-M-W42-T25) - dVrxn/dt (C3S-M-RD-W42-T25-P1)
dHrxn/dt (mW/g) (C3S-M-W42-T35) -dVrxn/dt (C3S-M-RD-W42-T35-P1)
dHrxn/dt (mW/g) (C3S-M-W42-T45) dVrxn/dt (C3S-M-RD-W42-T45-P1)
12 4.5

-
4
10
3.5
8 3

7Ii~
E 2.5
6 _____________ _____________ _____________
2
4 1.5
1
2
-

0.5
0 0 00-
0 0.2 0.4 0.6 0.8 0 0.2 0 .4 0.6 0.8

(a) (b)

Figure 6-12: Rate of enthalpy of reaction (a) and rate of volume of reaction (b) as
function of progression of the hydration reaction. Enthalpy of hydration as measured
with isothermal conduction calorimetry on C 3 S-M cured at three different temper-
atures. Volume of reaction is measured with the pressure vessel and the degree of
hydration is calculated assuming an enthalpy of reaction of 520 J/g for the complete
hydration of C 3 S. The degree of hydration as measured from the volume of reaction
is also based on the total enthalpy of 520 J/g. which corresponds to J/pL as discussed
later on in chapter 8

Hrxn (J/g) (C3S-M-W42-T25) -- Vrxn (C3S-M-RD-W42-T25-P1)


-- Hrxn (J/g) (C3S-M-W42-T35) -Vrxn (C3S-M-RD-W42-T35-P1)
Hrxn (J/g) (C3S-M-W42-T45) Vrxn (C3S-M-RD-W42-T45-P1)
350 45
40
300
35
250
30
x 200 25
20
150
15
100 10

50 5
0 1-
0 100 200 300 400
0 50 14 00 150 200
Time (hr)
Time (hr)

(a) (b)

Figure 6-13: Cumulative enthalpy and volume of the reaction as lmeasured with
isothermal colichictiomi calorimetry on C:3 S-i cured at three different temperatures.

123
Ev (C3S-M-RD-W42-T25-Pl) Ev (C3S-M-RD-W42-T25-P1)
- Ev (C3S-M-RD-W42-T35-P1) - Ev (C3S-M-RD-W42-T35-P1)
_ Ev (C3S-M-RD-W42-T45-P1) Ev (C3S-M-RD-W42-T45-P1)
600 600
-

500 500
-

400 400
,

300 300
-

~L2-
200 200
-

100 100 I
-

0 0
0 100 200 300 400 500 0.35 0.55 0.75
Time (hr)

(a) (b)

Figure 6-14: Volumetric strain as measured in the RDC for C 3 S with mediimii particle
size (listribut ion (C3S-M\1) cured at different temperatures.

124
6.3 #-C 2S

Similar to C 3 S, the hydration reaction of C 2 S is well-known and produces the sample

hydration products; calcium hydroxide (CH) and C-S-H and it constitute one of the

major phases in CBM. The main differences between C 3 S and C 2 S are the rate of

hydration, the amount of CH produced during the hydration reaction, and the molar

volume of the anhydrous solids. These differences will help understand the role of

C-S-H and CH in the bulk volume changes. Similar to C 3 S, bulk volume changes of

C 2 S are studied for three different PSD cured under saturated conditions at 40'C.
Water-to-solid mass ratios are chosen to obtain reasonable workability of the slurry

and study the impact of w/s ratio on the bulk volume changes.

Isothermal conduction calorimetry is conducted on C2 S with three different parti-

cle size distributions mixed at water-to-solid ratio of 0.45 and cured at 40'C. Curing

temperature of 40'C is chosen for the hydration of C 2 S due to the very low reactivity

and to achieve higher degree of hydration in reasonable time. The volume of the

reaction is measured in the pressure vessel for three samples with fine, medium, and

coarse particle size distributions with w/c = 0.42, 0.38, and 0.34, respectively.

Figure 6-15 shows the rate of enthalpy and volume of reaction in function of

the degree of hydration, where the degree of hydration is calculated based on the

enthalpy of the reaction and volume of reaction assuming AHr., = 260 J/g. The

rate in function of time is shown in Figures 6-16 for the different PSD and w/c ratios

and the corresponding accumulative enthalpy and volume of reaction are shown in

Figure 6-17.

The resulting bulk volume changes as measured in the pressure vessel are shown

in Figure 6-18. All samples are tested in the rubber diaphragm configuration (RDC)

at 40'C curing temperature and 1 MPa curing pressure while the effective stress

delivered through the check valve is 60 kPa.

125
-- dHrxn/dt (mW/g) (C2S-F-W45-T40) dVrxn/dt (C2S-F-RD-W42-T40)
-- dHrxn/dt (mW/g) (C2S-M-W45-T40) -dVrxn/dt (C2S-M-RD-W38-T40)
dHrxn/dt (mW/g) (C2S-G-W45-T40) dVrxn/dt (C2S-G-RD-W34-T40)
0.9 1
0.9
0.8
0.8
0.7
0.7
0.6
0.6
0.5
0.5
0.4 0.4
x 0.3 x 0.3
0.2 0.2
0.1 0.1
0 0 01 02
0 0.1 0.2 0.3 0.4 0 0.1 0.2 0.3 0.4

(a) (b)

Figure 6-15: Enthalpy (a) and volume of the reaction (b) as a function of degree of
hydration reaction. Enthalpy of hydration as measured with isothermal conduction
calorimetry on C2 S with three different particle size distributions. Volume of reac-
tion is ieasured with the pressure vessel and the degree of hydration is calculated
assuming an enthalpy of reaction of 260 J/g for the complete hydration of CGS. The
degree of hydration as measured from the volume of reaction is also based on the
total enthalpy of 260 J/g. which corresponds to 35.0 2.0 J/pL as shown in Figure
9-1.

- dHrxn/dt (mW/g) (C2S-F-W45-T40) -- dVrxn/dt (C2S-F-RD-W42-T40)


- dHrxn/dt (mW/g) (C2S-M-W45-T40) - dVrxn/dt (C2S-M-RD-W38-T40)
dHrxn/dt (mW/g) (C2S-G-W45-T40) dVrxn/dt (C2S-G-RD-W34-T40)
0.9 1

0.8
0.7
- -_ 7tIi -_ 0.9
0.8

- i 0.7
0.6
0.6
0.5
0.5
--
--- - --
0.4
0.3
1) I
x
0.4
0.3
--I-
0.2
0.1
- --L 0.2
0.1
0 0
1 10 100 10 100
Time (hr) Time (hr)

(a) (h)

Figure 6-16: Rates of (a) Enthalpy an(l (b) Vohinne of the reaction measured during
the hydration of CS at, 40'C.

126

]
Hrxn (J) (C2S-F-W45-T40) -Vrxn (C2S-F-RD-W42-T40)
- Vrxn (C2S-M-RD-W38-T40)
-Hrxn (J) (C2S-M-W45-T40)
Hrxn (J) (C2S-G-W45-T40) Vrxn (C2S-G-RD-W34-T40)
100 _ 16

90 14
-

80 12
-

70
-

10
60
-

on
50 8
40 6
30
4
20
10 2

0 0 0
0 100 200 300 0 100 200 300
Time (hr) Time (hr)

(a) (b)

Figure 6-17: (a) Cumulative entlialpy of reaction as measured by isothermal calorine-


try on C 2 S pastes with different particle size distributions and w/c = 0.45. (b) Cu-
nulative volume of the reaction as measured in the pressure vessel on C2 S pastes with
different particle size distributions.

I
- Ev (C2S-G-RD-W34-T40) -- Ev (C2S-F-RD-W42-T40)
Ev (C2S-F-RD-W42-T40) -Ev (C2S-M-RD-W38-T40)
Ev (C2S-M-RD-W38-T40) Ev (C2S-G-RD-W34-T40)
350 350

300 300

250 -Ile, 250


-

200 200
-

150 150
-

100 100
-

50 50

0 0-
50 100 150 200 250 300 0.1 0.2 0.3 0.4
Time (hr)

(a) (b)

Figure 6-18: Bulk volume changes during the hydration of C.S at 40'(. Fig. (a)
shows the volumetric strain as a function of time. Fig. (b) Shows the vohunetric
strain as a function of the degree of hydration.

127
6.4 Class G cement

Unlike the hydration of pure phases, the hydration of Class-G cement involves a

number of simultaneous hydration reactions (see, for example, Table 10.2) with ap-

proximately a single transition state (activated complex). Despite the complexity of

these chemical reactions, it is usually C 3 S that is the dominant phase and the phase

with the largest contribution to the enthalpy and volume of the reaction. This section

presents the results of bulk volume change measurements and hydration kinetics of

Class-G cement tested in both the RDC test and the LBC test.

Bulk volume changes are measured for Class-G cement (CGC) samples with dif-

ferent water-to-cement ratios (w/c) and different particle size distributions. All CGC

samples are hydrated at 25'C in the RDC and LBC configurations to investigate the

effect the boundary conditions.

6.4.1 Hydration kinetics

Enthalpy H,,, and volume of the reaction Vrn have been measured for a number

of Class-G cement samples with a range of particle size distributions as detailed

in Section 3.3 with various water-to-solid ratios. Figures 6-19 to Figure 6-22 show

the measured enthalpy and volume of the reaction of these samples at 25 C. The

relationship between the enthalpy and volume of reaction (Fig. 6-23) is linear for the

most part of the hydration reaction except for the seeded materials where a significant

deviation is observed at about 16 hours due to the hydration of C 3 A and what seems

to be related to the formation of calcium aluminosulfates (see Figure 6-21 and Figure

6-22).
Relationship between the enthalpy and the volume of the reaction of Class G

cement samples with a range of particle sizes is shown in Figure 6-23. The average

ratio is 7.88 0.34 J/pL for all PSDs except for the coarse CGC-RDC-06 (CGC-C1-
RD-W32-T25) sample where the ratio was about 6.5 J/pL and for the seeded samples

the relationship between the enthalpy and volume of the reaction is non-linear, but is

approximately 7.8 J/pL. It is well-known that the addition of seeds to cement paste

128
will accelerate the hydration reaction in general and especially the hydration of C:3A

after the depletion of sulfate as also evident from the rate of volume of the reactioll.

In contrast with the rate of enthalpy shown in Figure 6-21. the measured volume of

reaction of the C 3 A is munch larger than that of the preceding C:3 S hydration, which

indicates a high V,, to H,,,, most likely related to the formation of nioio-sulfates

and hydrogarnet.

-- dHrxn/dt (CGC-F2-RD-W49-T25) -- Hrxn (CGC-F2-RD-W49-T25)


-dHrxn/dt (CGC-Fl-RD-W38-T25) - Hrxn (CGC-C1-RD-W32-T25)
-dHrxn/dt (CGC-Full-RD-W38-T25) - Hrxn (CGC-Full-RD-W38-T25)
-dHrxn/dt (CGC-Cl-RD-W32-T25) -Hrxn (CGC-Fl-RD-W38-T25)
4 400

3.5 - - - - - 350

3 -- -___-- --- l~
___ 300
2.5 250

E 2 - - 200 --...
......
-

1.5 I- 150

1 -- -- --- - - - 100

0.5

0
1 10 100 0 50 100 ISO 200
Time (hr) Time (hr)

(a) (b)

Figure 6-19: Isothermal calorinmetry results on Class-G cement pastes with different
particle size distributions with variable water-to-solid ratio. (a) Rate of the eintlialpy
of the reaction dHrx/dt as imeasured by the isothermal calorimetry. (b) Cumulative
enthalpy of reaction as measured by isothermal calorinetry oil Class-G cement. For
details of the PSD see Table 3.5 and for nomenclature see Figure 6-1.

6.4.2 Bulk volume changes

Bulk volume changes of Class G cement samples with two boundary conditions in
the LBC and RDC configurations were determined for various curing pressures and

particle size distributions. A list of the tested samples is shown iin Table 3.5. We start

by sllowiing that the evolution of permeability of samples prepared with w/c=0.42


caimot create significant pressure drop across the sample. To achieve that. samples

with w/c of 0.3. 0.35 and 0.42 are tested in the PVC pernieability configuration showmn

in Figure 5-1(b).

129
- dVrxn/dt (CGC-F1-RDW38-T25-Pl) Vrxn (CGC-C2-RD-W36-T25-P1)
- dVrxn/dt (CGC-FuII-RD-W38-T25-P1) Vrxn (CGC-F1-RDW38-T25-P1)
- dVrxn/dt (CGC-C1-RD-W32-T25-P1) - Vrxn (CGC-FulI-RD-W38-T25-P1)
dVrxn/dt (CGC-F2-RD-W49-T25-P1) - Vrxn (CGC-C1-RD-W32-T25-P1)
dVrxn/dt (CGC-C2-RD-W36-T25-P1) -Vrxn (CGC-F2-RD-W49-T25-P1)
2 50

-
40
Z

-
1.5 - _-
'Foil

-
011
30
1 - -- - I - --

-
20

0.5
--. 10

0 0
1 10 100 0 50 100 150 200
Time (hr) Time (hr)

(a) (b)

Figure 6-20: Volume of the reaction results on Class-G cement pastes with different
particle size distributions with variable water-to-solid ratio. (a) Rate of the volume
of the reaction dl TT/dt as measured in the pressure vessel. (b) Volume of reaction
as measured in the pressure cell on Class G cement pastes with different particle size
distributions with variable water-to-solid ratio. For details of the PSD see Table 3.5
and for nonenclature see Figure 6-1.

The rubber diaphragm tests are carried out at 1 MPa curing pressure with two

effective stresses. The sole purpose of the effective stress applied through the clieck
valve on these sanples is to guarantee contact between the sample and the rubber

diaphragn. The magnitude of the effective stress (cracking pressure of the check

valve) is chosen such that it cannot cause measurable creep and to reflect conditions

similar to the autogenous shrinkage test. Pressure drop due to the use of such check

valve is shown in Figure 6-26. Pressure at the bottom surface of the sample starts to

decline once the sample is set and all the bleed water is consumed due to the negative

vohne of the reaction.

Figures 6-25 and 6-26 show the bulk volume changes of Class G cement samples

with full particle size distributiou as a function of time and degree of hydration as

well as the evolution of the pressuri at the bottom surface of the sample. Pressure

drop starts after approximately 5 hours of hydration. which is (lose to the typical

setting time of cement.


Fioure 6-27 and Figure 6-28 show the bulk volune changes of Class G cement

130
- dHrxn/dt CGC Ref -Hrxn CGC Ref
dHrxn/dt CGC+20% Hrxn CGC+20%
dHi-rxn/dtCGC+100% ----Hrxn CGC+100%
5 400
4.5 350
4
300
3.5
3 -- - - - -- 250
2 2.5 -; ~ 200
2 150
1.5 ---
100
1 1
4
0.5 50

0 0 0
10 100 0 50 100 150 200
Time (hr) Time (ir)

(a) (b)

Figure 6-21: Isothermal calorimietry results on Class-G cement pastes with different
contents of silica flour seeds. (a) Rate of the Enthalpy of the reaction dH../dt as
imeasured by isothermal clorimetry on Class G cement pastes with different content
of silica flour. All samples are Mixed at water-to-solid mass ratio of 0.42 with 20%
and 100% replacement of silica flour by mass of cement. It is well-known that the
addition of seeds to cemeit paste will accelerate the hydration reaction in general
and especially the hydration of C 3 A after the depletion of sulfate as also evident fron
the rate of heat. (b) Cumulative enthalpy of reaction as measured by isothermal
clorimnetry on Class G cement lastes with different content of silica flour. All samples
are mixed at water-to-solid mass ratio of 0.42 with 20% and 100% replacement of silica
flour by mass of cement. For details of the PSD see Table 3.5 and for nomenclature
see Figure 6-1.

131

I
-dVrxn/dt (CGC-FuI-RD-W38-T25-P1) -Vrxn (CGC-Ful-RD-W38-T25-P1)
dVrxn/dt (CGC+20%) Vrxn (CGC+20%)
dVrxn/dt (CGC+100%) Vrxn (CGC+100%)
2.5 50
45
2 40
~Ij 35
~IJ 1.5 -_ 30
25
1 20
15
0.5 10
5
0 0
10 100 0 50 100 150 200
Time (hr) Title

(a) (b)

Figure 6-22: Volume of the reaction results on Class-G cement pastes with different
silica flour content. (a) Rate of the volume of the reaction dl 'i, /dt as measured in the
are
pressure vessel on class G cement with different content of silica flour. All samples
mixed at water-to-solid nass ratio of 0.42 with 20c and 100% replacement of silica
flour by mass of cement. It is well-known that the addition of seeds to cement paste
will accelerate the hydration reaction in general and especially the hydration of C:3A
after the depletion of sulfate as also evident from the rate of volume of the reaction.
In contrast with the rate of enthalpy shown in Figure 6-21. the measured volume of
reaction of the C:3A is much larger than that of the preceding C3 S hydration, which
indicates a high V, to H,,, most likely related to the formation of nmono-sulfates
and hydrogarnet. (b) Volume of reaction as measured in the pressure cell on Class G
ceimeit pastes with different particle size distributions with different content of silica
flour. All samples are imixed at water-to-solid mass ratio of 0.42 with 20% and 100/
replacement of silica flour by mass of cement. For details of the PSD see Table 3.5
and for nomenclature see Figure 6-1.

132

_____----I
400 -

-
350 -

-
300 -_-
9 250 -- _-__-

' 200
-

150 a __
15 - ClassG+20%
---
--- ClassG+100%
100 -__
CG-RDC-05
50 CGC-RDC-06
0 C0q,-RDC-07
0 10 20 30 40 50
Vrxn (pL/g)

Figure 6-23: Relationship between the enthalpy of the reaction and the volume of the
reaction of Class G cement samples with a range of particle sizes. The average ratio
is 7.88 0.34 J/piL for all PSDs except for the coarse CGC-RDC-06 sample where the
ratio was about 6.5 J/pL and for the seeded samples the relationship between the
enthalpy and volume of the reaction is non-linear.but is approximately 7.8 J/pL.

samples with Coarse particle size distribution as a function of time and degree of

hydration as well as the evolution of the pressure at the bottom surface of the sample.

The pressure drop at the bottom surface depends on the consumption of bleed water

and setting. In sample (CGC-C2-RD-W36-T25-P1) pressure drop starts at later age

due to significant sedimentation and low rate of hydration, while the pressure drops

immediately in sample (CGC-C1-RD-W31-T25) due to very low w'/c ratio.

Bulk volume changes in function of time and degree of hydration for Class-G

cement with fine particle sizes are shown in Figure 6-29. Initiation of bulk volume

change is based on pressure variation at the bottom surface. as showin in Figure 6-30.

Liquid barrier tests are carried out, at variable curing pressure ranging from 0.2

to 10 1Pa as listed in Table 3.5. In a typical liquid barrier test, the cement slurry

is placed inside the vessel directly on top of the filter and the barrier liquid is placed

directly on top of the slurry. Similar to other bulk volume change measurements,
complete contact between the barrier liquid and the onset of bulk volume change

measurement starts uponi setting and consumption of the bleed water to the reaction.

leaningful data from such test is only collected after the bleed water is completely

133
absorbed inside the sample. Bulk volume changes of Class-G cement are measured

in the liquid barrier configuration (LBC) under different curing pressures and sample

thicknesses. Figure 6-31 and Figure 6-32 show the bulk volume changes for Class-G

cement samples cured at 25'C in the LBC and the volume of the reaction as measured

simultaneously during the tests. Bulk volume changes of Class-G cement with variable

thicknesses are shown in Figure 6-33 and Figure 6-34.

5.5 45
-

40
5
-

35

-
4.5
30

4- 25

20
3.5 x -CGC-FuII-PVC-W30-T25
15 CGC-Fufl-PVC-W42-T25
3-
CGC-FuII-PVC-W35-T25
10
2.5 GC-Ful-PVC -W3-T2 5
CGC-FuIl-PVC-W35-T25
2- 0
0 100 200 300 0 100 200 300
Time (1br) Time (hr)

(a) (b)

Figure 6-24: Pressure development and volume of the reaction for Class G cement
with different w/c. Fig. (a) shows the pressure development at the bottom surface of
the sample cured at 25'C. The vessel is configured as shown in Figure 5-1(b). Fig.
(b) The measured volumie of the reaction for the same samples.

134
Ev (CGC-FuIl-RD-W38-T25) Ev (CGC-FuII-RD-W38-T25)
-Ev (CGC-FulI-RD-W38-T25) Ev (CGC-Full-RD-W38-T25)
Ev (CGC-Full-RD-W38-T25) Ev (CGC-FuIl-RD-W38-T25)
1400 1400

1200 1200

1000 1000

800 7- - > 800


600 600 ________ I
400 400

200 200 L-
0 0
0 50 100 150 200 0 0.2 0.4 0.6 0.8
Time (hr)

(a) (b)

Figure 6-25: Bulk volume changes of Class G cement samples in the rubber diaphragm
test. All samples experience fast expansion up to 40 hours followed by slower expan-
sion and shrinkage at about 100 hours. Data are initiated after consumption of the
bleed water or setting of the slurry as observed by pressure variation at the bottom
surface of the sample (see Figure 6-26).

I
Pc (CGC-Full-RD-W38-T25)
-Pc (CGC-FuII-RD-W38-T25)
--- Pc (CGC-FuII-RD-W38-T25)

1030
1020
1010
w
1000

w 990
980
E 970
C
C 960
950
940
0 20 40 60 80 100
Time (hr)

Figure 6-26: The evolution of pressure at the bottom surface of the sample during
hydration in the rubber diaphragm test, where the pressure drop is dIc to the check
valve. Sample CGC-RDC-01 is tested with a 14 kPa check valve while all the other
samples are tested with 60 kPa check valve.

135

I
- Ev (CGC-FuI-RD-W38-T25) --- Ev (CGC-Full-RD-W38-T25)
-Ev (CGC-C1-RD-W31-T25) -- Ev (CGC-C1-RD-W31-T25)

Ev (CGC-C2-RD-W36-T25) Ev (CGC-C2-RD-W36-T25)

12000 12000

-
-

10000 10000

-
-

8000 8000

-
-

6000 6000
-

4000 4000
-

2000 2000 - -0.


-

0- 0
0 50 100 150 200 250 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Time (hr)

(a) (b)

Figure 6-27: Bulk volume changes of Class G cement samples in the rubber diaphragn
test with coarse PSD. CGC-R.DC-05 is the reference sample with full PSD and w/c
0.38.

I
-Pc (CGC-C1-RD-W31-T25)
-Pc (CGC-Full-RD-W38-T25)
Pc (CGC-RD-C2-W36-T25-P1)
1040
1030
1020
1010
a) 1000
990
a)
980
E 970
C
960
C
950
940 -u-
0 20 40 60 80 100
Time (hr)

Figure 6-28: The evolution of )ressure at the bottom surface of the sample during
hydration in the rubber diaphragm test. where the pressure drop is due to the check
valve. Sample CGC-RDC-02 is teste(l with a 14 kPa check valve while all the other
samples are tested with 60 kPa check valve. It is evident here that sample CGC-
RDC-02 experienced significant sedimentation and hence the delayed pressure drop.
Pressure diro) at the bottom of sample CGC-RDC'-06 is innnediate due to the low
w/ c.

136

____________ I
- Ev (CGC-F1-RD-W38-T25) - Ev (CGC-F1-RD-W38-T25)
- Ev (CGC-F2-RD-W49-T25) - Ev (CGC-F2-RD-W49-T25)
Ev (CGC-Full-RD-W38-T25) Ev (CGC-Full-RD-W38-T25)
4000 4000
3500 3500
3000 3000
2500 2500
2000 2000

-
1500 1500
1000 1000
500 500
0 0
-500 -500
-1000 -1000
0 50 100 150 200 250 0 0.1 0.2 0.3
Time (hr) E

(a) (b)

Figure 6-29: Bulk volume changes of Class G cement samples in the rubber diaphragimi
test, with fine PSD. CGC-RDC-05 is the reference sample with full PSD and w/c=
0.38.

-- Pc (CGC-F2-RD-W49-T25)
_Pc (CGC-F1-RD-W38-T25)
Pc (CGC-Full-RD-W38-T25)
1040
1030
1020
1010
1000
Ln 990
980
970
E
960
950
940
0 20 40 60 80 100
Time (hr)

Figure 6-30: The evolution of pressure at the bottom surface of the sample during
hydration in the rubber diaphragm test. where the pressure drop is due to the check
valve. Sample CGC-RDC-03 has a 14 kPa check valve while all the other samples are
tested with 60 kPa check valve.

I
137
-Ev (CGC-LB-P1O-W42-T25-Full) - Ev (CGC-LB-P10-W42-T25-Full)

- Ev (CGC-LB-P1-W42-T25-Full) -- Ev (CGC-LB-P1-W42-T25-Full)
Ev (CGC-LB-P5-W42-T25-Full) Ev (CGC-LB-P5-W42-T25-Full)

0- 0-

-2- -2

-4- ~1~ - __ ..- ____ -______ -4-

E
-6 -6-
--_

-8 -8

-
-10 -10

-
-

-12 -12 -[
0 50 100 150 200 250 0.2 0.4 0.6 0.8
Time (hr)

(a) (b)

Figure 6-31: Bulk volume changes of Class G cement samples in the liquid barrier
test with full PSD with variable curing pressure.

I
-Vrxn (CGC-LB-P1O-W42-T25-Full) -( (CGC-LB-P1O-W42-T25-Full)
-Vrxn (CGC-LB-P1-W42-T25-Full) -- (CGC-LB-P1-W42-T25-Full)
Vrxn (CGC-LB-P5-W42-T25-Full) ( (CGC-LB-P5-W42-T25-FuLII)
45 0.8
-
-

40 0.7
-

35 0.6
-

30
0.5
I
-

25
X
/- 0.4
________ ________ _____

I
_____
-

_
20 ____________ _________ ___________ ________

15
0.3 1~
10
0.2
~1
5 0.1
,1
0 0
0 100 200 300 0 50 100 150 200 250
Time (hr) Time (hr)

(a) (b)

Figure 6-32: Volunie of reaction and degree of hydration of Class G cenent saniples
in the liquid barrier test with variable curing pressure.

138

i _----. - -a
MUM-w:

(CGC-LB-PO.2-W42-T25-H18.6)
-Ev Ev (CGC-LB-PO.2-W42-T25-H18.6)
Ev (CGC-LB-PO.2-W42-T25-H17.8) - Ev (CGC-LB-PO.2-W42-T25-H17.8)
Ev (CGC-LB-PO.2-W42-T25-H47.3) Ev (CGC-LB-P.2-W42-T25-H47.3)
0 0

-0.5 -0.5
-j
-1 -1
E
-

-
-1.5 -1.5
- - - --
-2 -2

-2.5 -2.5

-3 -3

-3.5 -3.5

-4 -4
0 50 100 150 200 250 0 0.2 0.4 0.6 0.8
Time (hr)

(a) (b)

Figure 6-33: Bulk volume changes of Class G cement samples in the liquid barrier
test with full PSD with variable thicknesses.

- Vrxn (CGC-LB-PO.2-W42-T25-H18.6) -- (CGC-LB-P.2-W42-T25-H 18.6)


- Vrxn (CGC-LB-PO.2-W42-T25-H17.8) -h (CGC-LB-PO.2-W42-T25-H17.8)
Vrxn (CGC-LB-PO.2-W42-T25-H47.3) ( (CGC-LB-P.2-W42-T25-H47.3)
45 0.8
-

40 0.7
-

35
0.6
-

IJ.
30 7/~~_
0.5
-

> 25

-K--~-~
0.4
-

20-
0.3
-

15
- -~. ______ _____
10 0.2
I
-
-

5 0.1
~1 ~
0
-/ -~ -~

0-
- -- 200 2
-

0 50 100 150 200 250 0 50 1()0 150 200 2 50


Time (hr) Time (hr)

(a) (b)

Figure 6-34: Volume of reaction and degree of hydration of Class G cement smles
in the liquid barrier test with variable thicknesses.

139
Part IV

Discussion

140
Chapter 7

Expansion During the Hydration


of Magnesium Oxide

The main purpose of including in our investigation the hydration of magnesium oxide

(MgO) is to calibrate and assess the performance of the main experimental setup (the

pressure vessel). MgO serves as a simple model material with well-known chemical

reactions, enthalpy of reaction, volume of reaction, and mechanical properties to study

the bulk volume changes of reactive porous solids. The hydration kinetics of MgO

also resembles the kinetics of cement-based materials, but through a single chemical

reaction with a single hydration product while the complete chemical reaction can be

achieved in one week at 25'C. The enthalpy of hydration of MgO is 930 J/g based on

thermochemical data tabulated in Ref. [35]. The volume of reaction can be calculated

based on the molar volumes of the different constituents as listed in Table (A.2) to

be -109 pL/g
The morphology and reactivity of MgO depends on the synthesis procedures and

mainly on the calcination temperature. There exists a number of MgO grades classi-

fied based on the calcination temperature and consequently their reactivity.

This chapter discusses the results obtained for the hydration of light-burn MgO

presented in Section 6.1. Hydration of MgO is investigated with multiple techniques

including SEM, nitrogen sorption, calorimetry, TGA, and indentations, as well as

volume of the reaction and bulk volume change measurements in the pressure vessel.

141
The development of the microstructure during the hydration of MgO is first investi-

gated with SEM, nitrogen sorption, calorimetry, TGA, and micro-indentations. The

resulting understanding of the microstructure of the material is then employed to

construct and validate a multiscale microporomechonics model. The second part of

the discussion focusses on the development and validation of the model followed by

interpretation of the bulk volume change measurements to relate the macroscopic

measurements to the origin of bulk volume changes.

7.1 Hydration kinetics of MgO

Reactivity of MgO depends on the calcination temperature. The hydration kinetics

of MgO resembles the kinetics of cementitious materials with an induction period

followed by acceleration and deceleration periods as shown in Figure 6-2 and Figure

6-3. Such behavior can be attributed to a boundary nucleation and growth accelera-

tion followed by a diffusion-limited deceleration. The close resemblance between the

hydration kinetics of MgO and cement can help elucidate the kinetics and controlling

mechanisms of cement hydration. For example; while the induction period observed

during the hydration of cement is sometimes attributed to the formation of a semi-

permeable membrane [22], it is not clear as to how a semi-permeable membrane of

Mg(OH) 2 can lead to the induction during the hydration of MgO.

Hydration kinetics is monitored with both volume and enthalpy of the reaction.

The tested light-burn MgO is initially partially hydrated as shown by the TGA results

and the measured enthalpy of 860 J/gMgo is necessarily an under estimation of the

known 930 J/gMgo. The volume of the reaction on the other hand was overestimated

when measured in the pressure vessel. Light-burn MgO consists of aggregates of

approximately 100 nm crystallites forming porous particles ranging in size from sub-

micrometers up to 10's of micrometers. It is possible that part of the MgO porosity is

occluded, which can result in overestimation of the measured volume of the reaction.

142
7.2 Microstructure of MgO and Mg(OH) 2

To build an appropriate micro-poro-mechanics model to characterize the material at

hand, we need first to establish an understanding of the microstructure of the material.

To build this understanding, the material is investigated under the scanning electron

microscope at different stages of the chemical reaction and under varying reaction

conditions. Sorption isotherms are also collected for these samples to investigate the

surface area and pore size distribution.

The specific surface area of light burn MgO for both batches was about 25 m 2 /g

within the range of values reported for the material in the material datasheet (Table

3.1). Assuming a mono-disperse particle size distribution of MgO and a density of

3.58 g/mL, the average radius is calculated to be 34 nm,which is close to the size

observed under the SEM (Figure 7-1).

The total pore volume of light-burn MgO is 0.23 cm 3 /g (Figure 7-3e). This pore

volume is mainly encapsulated inside the MgO clusters and part of this pore volume is

due to the contact between the grains. If this volume is assumed to be encapsulated

entirely inside the MgO clusters and taking the density of MgO to be 3.58g/cm 3

,
one can estimate the porosity of these clusters as 0 =0.23/(1/3.58+0.23)=0.45 and

a packing density of 0.55. This packing density is very close to the random loose
55
packing of a mechanically stable system TIRLP=0- [139]. Similarly, the pore volume

of the fully hydrated Mg(OH) 2 is 0.06 cm 3 /g corresponding to a packing density of

0.88. The packing density of 0.88 indicates that sintering of the Mg(OH) 2 has occured

during hydration.

The surface area of the fully hydrated Mg(OH) 2 was also measured with nitrogen

sorption to be about 10 m 2 /g. Taking into account that the molar volume of Mg(OH) 2

is 2.21 time the molar volume of MgO (see Table A.2), one can estimate a specific

surface area (SSA) of 28 m 2 /g. The small measured SSA indicates either merging

(sintering) of Mg(OH) 2 particles as observed in the SEM micrographs (see Figure 7-2)

or possibly, complete dissolution of the smaller MgO particles.

Hydration of light burned magnesium oxide produces a granular cohesive solid of

143
I
Ii
2 G 0 -N (b) 545X

It"

ALM Tk
Irv, W wo
Jlii-@ v-r

(c) 4.31kX (d) 45.25kX

Figure 7-1: MNlicrostructure of anhydrous MgO under the SEM. The material in dry
powder form was placed on a conductive adhesive carbon tape and carbon coated for
imaging. The basic building block of the material is crystallite with crystal size of
around 100 inn forming particles of agglomerates that can go up to 20 nm in size.

144
-tpm OC-S 100pm E SWW CM
83..POP

(a) 2.66kX(BSE) () 172X

2pm j CM -S
E 2pm I&O CMSE

(c) 3.41 kX (d) 8.78kX

WA.
I
200 n 200 nm

(e) 46.81kX (f) 84.66kX

Figure 7-2: SEM micrographs of the microstructure of Mg(OH) 2 at full hydration at


(liffereit magnifications. A fractured surface of fully hydrated sample of Mg(OH)2
was exposed and coated with carbon for the investigation. The basic building block of
the material is crystallite with crystal size of around 200 min. The original structure
of the anhydrous MgO is still observable at low magmification.

145
160 70
140 60
120 -4 (MgO-RD-07)
50
- (MgO-RD-08)
100

1----
40 J(MgOR-1t-9)
80
-o, 30
60
0z 20
cjz 40
20 10
:5
0 0
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
Relative Pressure (P/Po) Relative Pressure (P/Po)

(a) (b)
7 2
6
1.5 -L I
5
-A 17 -2

-
-
f
4
SU 1-
3 - -I-i-t
'-I
2 -o 0.5 (vi{RD&O
1 -4-( RDC4
~0 ( d RDC0
0 0-

.
1 10 100 1 10 100
Pore Radius (nm) Pore Radius (nm)

(c) (d)
0.25 -"-T
-4(g RDC-07i
0.06 -'RDC--:0
0.2 -

g IC-09
U
0.15 0.04

4----
C.)

S
0
0.05
0.1
7 0

C.)
0.02

C.) 0
0 0
1 10 100 10 100
Pore Radius (nm) Pore Radius (nm)

(e) (f)

Figure 7-3: Nitrogen sorption isotheris for Mgo an(d fully hydrated Mg(OH) 2 . Fig.
(a) and (1) are the niitrogen sorJptioli isotheruis for the alllydrous light burn MgO
and fully hlydrated Mg(OH) 2 respectively. Figs. (c) and (d) show the dl/dr pore
volume and Figs. (e) and (f) show the cumulative pore volume for MgO an Mg(OH) 2
.

respectively.
146
magnesiuni hydroxide. Here we investigate the properties of fully hydrated inagne-

sium oxide by means of indentations. The application of indentation analysis to

heterogeneous porous materials is made possible by an extensive development of

the grid indentation technique by Constantinides and Ulm [44, 196], Vandainne and

Uhin [201. 202], and Gathier and Uln [80].

Samples of MgO were mixed at high water-to-solid mass ratios to obtain reasonable

workability. Then water is squeezed out under different effective stresses to obtain

materials with different porosities (Table 6.1 shows the 6 tested samples). Six samples

were prepared with two different water-squeezing mechanisms in a rubber diaphragm

test and in a confinement cell loaded with a piston.

In the rubber diaphragm. samples with known mass and water-to-solid ratio are

subjected to controlled effective stress for a specific period of time to squeeze the wa-

ter out of the sample (see Table 6.1). The sample is then kept saturated until the tine

of testing (see Figure 7-4). Samples are then obtained for the indentation test. poros-

ity measurements by total evaporable water, SEM and TGA. For indentation tests,

the sample is cut down to the appropriate dimensions and left to equilibrate to the

ambient relative humidity which also is the relative humidity during the indentation

test.

(a) (b)

Figure 7-4: Sample sqlleezed in the rubber diaphragm: (a) Sample inside the rub-
ber diaphragm. (b) Sample outside the rubber diaphragm. Notice the corrugated
(buckled) edge due to squeezing.

147
In the confinement cell and the piston, samples with known mass and water-to-
solid ratio are subjected to controlled effective stress for a specific period of time.
Once the desired amount of water is squeezed out, the volume of the sample is held
constant while measuring the evolution of stress (see Figure 7-5). The initial drop in
stress shown in Figure (7-5) is due to stress relaxation after switching from constant
stress to constant displacement and the stress development afterwards is due to the
hydration and crystallization pressure of MgO.

6.5 0.5 3.5


6.3 - - 0.4 - 3
6.1 - - - 0.3 2.5
5.9 -- 0.2 2
5.7 0.1 1.5 -

-
5.5. 0 1
0 50 100 0 50 100 0 50 100
Time (hr) Time (hr) Time (hr)

(a) (b) (c)

Figure 7-5: Evolution of compressive stress in the samples cured under the piston.
Fig. (a) is sample MgO-Pis-01, Fig. (b) is sample MgO-Pis-02, and Fig. (c) is sample
MgO-Pis-03 (see Table 6.1). Samples with known mass and water-to-solid ratio are
subjected to controlled effective stress for a specific period of time. Once the desired
amount of water is squeezed out, the volume of the sample is held constant while
measuring the evolution of stress.

For the porosity measurement with the total evaporable water, a representative
saturated-surface-dry sample is oven dried at 105'C and the porosity was calculated
assuming a density of 2.34 for the magnesium hydroxide (see Table 6.1 for the mea-
sured porosity). TGA test were also conducted on each sample to guarantee the full
hydration which was achieved after 6 days.
It is necessary to establish an understanding of the material behavior for the
choice of indentation load profile and suitability of the micromechanics model. We
start by studying the structure of the material under the scanning electron microscope
followed by micro-indentations carried out at different loads and loading times.
The SEM images (Fig.7-2) show the granular nature of the material while the BSE
image shows the non-homogeneous porosity distribution were black represents pores

148
and bright is dense contamination by polishing pads. The particles seem to have a

single size of about 200 nm in diameter while the local porosity has a characteristic

length of about 5 micrometers as seen in the SEM images. Based on these charac-

teristics, the activated area in an indentation test should be approximately one order

of magnitude larger than the characteristic size of the crystallites, namely, about 1

micrometer. An indentation depth of about 1 micrometer should provide information

about the heterogeneous local porosity distribution. An indentation depth of more

than 10 microns is required to capture the homogeneous response of the material, but

not too large to cause cracking.

The loading profile of an indentation test is typically designed in such a way to

capture the desired material properties without interference from other undesired

properties. In this study, we measure the indentation modulus and indentation hard-

ness to extract the solid modulus in addition to the cohesion and friction angle. For

this purpose, we need to eliminate any possible interference by creep properties and/or

fracture. The effect of maximum load and hold time on the measured indentation

modulus and hardness are investigated here to help chose the appropriate values for

the loading profile.

Figure 7-6 shows the variation of mechanical properties as a function of the in-

dentation load on sample MgO-Pis-03. There seems to be a decrease followed by a

slight increase of the mechanical properties by increasing the maximum indentation

load.

Figure 7-7 shows the variation of the mechanical properties as a function of the

hold time on sample MgO-Pis-03. Hold time was varied from 5 to 600 seconds and

there was not a recognizable pattern. The hold time did not have any significant

impact on the measured values of the indentation modulus and hardness.

The compiled indentation results are shown in Figure 7-8 for the six tested sam-

ples. The model fit is shown in Figure 7-9 (see Section B.3 for the details of the model).

The resulting properties from the fit are m, = 73.00 0.40 GPa, c, = 0.2301 0.0024

GPa, a = 0.4185 0.0045, and h, = 3.157 GPa. The value of 73 GPa for the solid
modulus is very close to the reported values in the literature of 71GPa (calculated

149
I
640 30

-
-
29.5
620
~1

-
-F
-
29
4

-
600
28.5
-

-
580 28 -i
-

-
cjz e-1
560 0t 27.5

-
-

-) 27

-
540
-

26.5

-
520 26
-

-
500 25.5
0 2000 4000 6000 0 2000 4000 6000
Fn (mN) Fn (mN)

(a) (b)

Figure 7-6: Variation of indentation hardness (a) and indentation iodulus (b) as a
function of the max indentation load Fn on sainple MgO-Pis-03. Five indents were
carried out at each indentation load with 30 seconds loading, 30 seconds hold. and
30 seconds unloading.

580 28.5
-
-

560 28
-
-

T
540 27.5
-
-

520 I .-
-o
0
27 -A

500 26.5
-
-

-------- --
-

480 26

460 25.5
-
-

1 10 100 1000 1 10 100 1000


Hold time (s) Hold time (s)

(a) (b)

Figure 7-7: Variation of indentation hardness (a) and indentation niodulus (b) as
a function of the hold tie on sainple MgO-Pis-03. Five in(lents were carried out
at each hold time with 30 seconds loadig. and 30 seconds unloading and 2000 nIN
maxiiiniuin indentation load.

150

I
Bow - - - pw -- - -- - -=

40

35

30

25
----- M model (pa)
20 --- 4 Gpa}{MgG-RD-G9)
-

M (Gpa) (MgO-RD-08)
15 M (Gpa{MgO-RD-07)
M (Gpa) (MgO-Pis-03)
10
/

I
I M (Gpa) (MgO-Pis-02)
/

I
5 I.
I M (Gpa) (MgO-Pis-01)
I
I
I
0--
0 200 400 600 800 1000
H (MPa)

Figure 7-8: Indentation modulus M and indentation hardness H compiled from the
different specinens shown in Table 6.1.

ms=73 cs=0.2301 alpha=0.4185 hs=3.1574 error=0.00056795


80 3.5
70 3 -/
60 -/
2.5
-50
2
(40
1.5
30
/' 1 7
'

20
/

10 0.5

W. 0.6 0.7 0.8 0.9 1 T. 0.6 0.7 O 0.9 1


eta eta

Figure 7-9: Relationship between the packing density and indentation modulus and
hardness from1 the experime1nt and the nodel for the niiero indetations on all dry
Sam11ples.

I
151
from the elastic modulus of 67 GPa). The cohesion of 230 MPa seems to be related to
the cementing and capillary reinforcement at the contact points between the particles
as observed in the SEM images. The elastic properties of a single crystal of brucite
were measured by Jiang et al. [102] by means of brillouin scattering. They carried out
the experiments at pressure ranging from ambient pressure to 15 GPa in a diamond
anvil cell. The resulting elastic constants were C1 = 154.0(14) GPa, C33= 49.7(7)
GPa, C 12 = 42.1(17) GPa, C13 = 7.8(25) GPa, C14 = 1.3(10) GPa, C4 4 = 21.3(4)
GPa and aggregate moduli K, =36.4(9) GPa, G, = 31.3(2) GPa, E, = 73.0 GPa
from the Reuss bound and K = 43.8(8) GPa, G, = 35.2(3) GPa, E, = 83.3 GPa
from the Voigt-Reuss-Hill average. The elastic modulus of anhydrous polycrystaline
MgO is E = 300 GPa and the shear modulus G = 130 GPa [177].
The relationship between the indentation modulus and hardness of a pure porous
solid can be obtained by synthesis of that material at different packing densities
(porosities). This approach was particularly applicable to Mg(OH) 2 as obtained from
the hydration of light burn MgO due to the small and mono-disperse particle sizes
and due to the presence of agglomeration of MgO crystallites.
The origin of cohesion in Mg(OH) 2 compacted during hydration seems to be re-
lated to the observed reinforcement (Solid capillary bridges) between the particles in
the SEM images. Compacted slurries of previously-hydrated magnesium hydroxide
did not form a solid cohesive material. Spatial variation of packing density is due
to the original clusters of MgO crystallites. This variation allows the use of nano-
indentation to extract properties of the material over a wide range of porosities on
one sample.

7.3 Volume fractions and relative volume changes

This section presents the relationships and calculations of the evolution of volume
fractions during the hydration of MgO as well as the relative volume changes due to
the evolution of volume fractions. The evolution of volume fractions are first discussed
in light of the known hydration reaction and properties of the different components.

152
Then the relative volume changes are discussed for a single MgO grain based on pure

geometric considerations and the possible contribution of the reaction mechanisms.

This section also presents the calculations of the relative volume changes of the porous

MgO particles as a possible interpretation of the observed bulk volume changes.

7.3.1 Evolution of volume fractions during the hydration of


MgO

Evolution of the volume fractions of the system components can be related directly to

the enthalpy or volume of the reaction given a known chemical reaction with known

stoichiometric coefficients. To estimate the evolution of the volume fractions of the

different constituents of the system, we consider the following chemical reaction along

with material properties in Table A.2:

MgO(S) + H20(1) , Mg(OH) 2 (s) (7.1)

The reaction proceeds with the dissolution of MgO and precipitation of Mg(OH) 2 and

the resulting volume fractions of the different components is then:

frxn = (Rmg(OH) 2 ~ RH - 1)(1 - no) (7.2)

f(Mgo) = (1 - no)(1 - ) .3)

f Mw = no - (1 - no)RH (7.4)

f(Mg(OH) 2) = (1 - no)RMg(OH) 2 (7.5)

And the total porosity of the system in the absence of bulk volume changes is

#tot = 1 - f(MgO) - f(Mg(OH) 2 ) (7.6)


= no + (1 - no)(Rhp + 1)

153
where R,, = 2.21, f is the fractional volume of the reaction. j(NgO) is the volume

fraction occupied by the non-porous solid MgO. f... is the volume fraction occupied
by the remaining initial water used for mixing (initial water is treated as a sepa-

rate phase just like solids), f(Mg(OH) 2 ) is the volume fraction occupied by the solid

non-porous Mg(OH) 2 , 0O is the initial Porosity upon mixing. and ( is the fractional

conversion. A plot of the evolution of the volume fractions is shown in Figure 7-10.

For more details on the definition of each term, see Appendix A. As observed under

the SEM in conjunction with nitrogen sorption isotherms, the total porosity of the

system 4,t is divided into particle porosity (4,, and the macro porositv ), such

that:

Otot = OPar + Om (7.7)

The distribution of the porosity will be further detailed in the discussion of the mutis-

cale thought-model for the hydration of MgO.

1t
0.9

0.8

.00.7 fm-
00.6

10.5-

0.4
Mfg(OH)2

0.1 fJgo
0--
0 0.2 0.4 0.6 0.8 1

Figurie 7-10: E-volution of volume fraction of the various comnponents during tHe
hy. dratin of Mlgo with w/cl( = 1.

154
7.3.2 Geometry and fractional conversion of a composite sphere

For the purpose of this analysis, we consider a single non-porous crystal of reactive

solid (the core co) and the growing reaction products are also composed of non-porous

solids (the shell sh) as illustrated in Figure 7-11.

/a
(a) 0 KC.V=KC 0
S= 0 Core . G s=Gco

b
(b) a Kcs
Core Gs
0<4<1
Shell

bo KCs=Ksh
Gi=Gi0
Shell)

Figure 7-11: A schematic of the composite sphere .(a) at mixing time =0. (b)
during the hydration reaction > 0, (c) at complete hydration 1.

The core has a radius a = a (() and the growing sphere has a radius b b()., the
radils of the core at a fractional conversion is

a = ao( - )3 (7.8)

where u() is the radius of the core at 5 = 0 (see Figure 7-11). Equation (7.8) is based

on the assumption that the volume of the core is linearly related to the conversion as

follows
4
VCO = I" (I - () = 77al1 0) (7.9)

The ratio of the volume of the shell 1, to the volume of the core 1" can be calculated

155
based on stoichiometry of the reaction and molar volumes of the components:

dVsh _ dfh _ Vsh = - Rp; dV, < 0 (7.10)


dVeO dfrs Vco - VcOO

where dfrs > 0 is the incremental volume fraction of the reactive solid (MgO) con-

sumed in the reaction. The radius of the growing sphere is:

b = ao ((Rhp - 1 + 1)1/3 (7.11)

The volume fraction of the shell is:

Vsh _ b3 - a3 Rh(1
f V=C b3 (Rhp - 1) +1

where Vc, is the volume of the composite sphere. The volume fraction of the core is:

fCO= V _ a3 1 - (7.13)
* C8e b (Rhp- 1) + 1

The relative volume change is calculated relative to the original volume of the core

as:
E* = O = (Rhp - 1) (7.14)

It is important to note here that depending on the reaction kinetics and mecha-

nisms, the ratios can be drastically different. This analysis is valid if the reaction is a

surface reaction or topochemical reaction, but it should be modified at larger scales if

the reaction is a through-solution reaction. In a through solution reaction, not all the

dissolving core will contribute to the volumetric strain because part of the dissolving

ions will precipitate in the water-filled pore space. To account for the fact that not

the entire dissolving core contributes to the eigenstrain, Rhp should be multiplied by

a coefficient c accounting for the reaction mechanism and kinetics, and the relative

volume change becomes:


* = (cRhp - 1) (7.15)

156
The value of c depends on the degree of hydration and the thickness of the hydration
layer which in turn controls the reaction mechanism. In case of topochemical reaction
the entire volume difference between the reactive solid (MgO) and the hydration
products (Mg(OH) 2 ) is manifested as bulk volume change with c = 1. However, If
the chemical reaction proceeds through solution, c can range from 1 to 1 if expansion
is measured and it can be smaller than 1
Rhp
if shrinkage is possible. The thickness of
the hydration layer assuming radial growth only is:

t = b - a = ao ( (cR,- 1)+ 1- /1-) (7.16)

7.3.3 Relative volume change in the porous particles

Another possible way to view the relative volume change in the MgO particles is by
considering all the MgO grains in a single particle to maintain the relative positions
inside the particle. In this scenario, only the grains on the outer perimeter of the
particle contribute to the volume change as illustrated in Figure 7-12. The relative
radial deformations in this case can be described in terms of the crystal and particle
radii as follows:

b - ao a0 [((Rhp- 1) + 1)1/3 _ (1
R - Ro
E* = 0 =t o (7.17)

And the relative volume changes is

-3
3
, R3 R -R - (o -b -ao) - 3
3 RR (7.18)

where b is the radius of the composite sphere and ao is the original radius of the MgO
grains while RO is the initial radius of the porous particle. Taking the average radius
of the particle to range between 3 and 8 microns from Table 3.1, which is also in line
with the particle sizes observed under the SEM in Figure 7-1, and an approximate
grain size of MgO of 100 nm, the relative radial deformation can be calculated as
shown in Figure 7-13. The value obtained for volumetric strain obtained in this way

157
a-. a

is close to the vohne strain ieasured experimentally and is discussed further in the
following sections.

0.30.

RRe 00 ..

of HO~*

3o3~ ***~ (a.0

(b)
Figure 7-12: A schematic of the hydration of MgO particles where the volume change
is only attributed to the MgO grains on the outer boundary of the particles. (a) is
the original anhydrous MgO particle (b) The MgO particle when fully hydrated into
Mg(OH) 2 . It is assumed in this model that all the grains maintain the original posi-
tions due to the sintering of the grains. The sintering of MgO grains is a consequence
of the calcination process while sintering of Mg(OH) 2 is due to the progression of the
hydration reaction.

10

10

10

10

10
10 10 102
1? (pim)

Figure 7-13: Rate of eigenstrain in the MgO-Mg(OH) 2 particles as function of the


radius of the porous particles (1ie line). The dashed red line represents the strain
ineasurd experimentally of about 0.015 for w/.s = I and the corresponding initial
raldils of the porous Mg() particlc R ~ 3pmi.

158
7.4 Bulk volume changes and multiscale microp-
oromechanics for the hydration of MgO

A single MgO particle consists of a large number of agglomerated MgO grains as

observed in the SEM images in Figure 7-1. The packing density of these particles

is about ripar = 0.55 as shown in Section 7.2. All samples were initially mixed with

high water-to-solid mass ratio (w/s ~_ 1) due to the very high specific surface area

(~- 25m2 /g) of the hydrostatically neutral but hydrophilic surface of MgO (periclase).

Such high w/s ratio produced coherent, yet very weak porous solid of Mg(OH) 2

7.4.1 Multiscale Thought Model of MgO hydration

During the hydration of MgO and due to the fact that the volume of the solid reaction

products is larger than the volume of the solid reactants, the material experiences

a bulk volume change. The bulk volume change in this case is expansive, but the

extent of the expansion and the associated stress development are also related to

the microstructure of the material, the reaction mechanisms, and the deformation

behavior of the solid. The origin of the stress and the expansion during the hydration

of MgO is assumed to be related to the difference in volumes between the reactive

solids and the hydration products.

To estimate the elastic properties and the state of stress of the porous solid re-

sulting from the hydration of MgO, a 3-level multiscale microporomechanics model

is developed here as illustrated in Figure 7-14. Based on the microstructure investi-

gated previously, the multiscale microporomechanics model will consist of a three-level

model of the MgO-Mg(OH) 2 microstructure. The first level (level I) of this model

consists of nano scale crystals of hydrating MgO as a composite sphere where the

reacting MgO forms the core and the precipitating Mg(OH) 2 forms the shell (Figure

7-14(a)). The second level (level II) includes multiple grains (crystals) from level I

in addition to the interstitial space which will be referred to as the particle porosity

(Figure 7-14(b)). The second level (Level II) is analyzed using two methods: the first

159
method is the self-consistent method in which the aggregate of crystals is modeled as

larger scale particles. The second method is the Mori-Tanaka scheme in which the

continuous porous matrix is composed of the composite spheres. Finally, at the third

level (Level III), the aggregates of crystals are assumed to form a continuous porous

matrix composed of the porous particles (Fig. 7-14(c)) for which a Mori-Tanaka

scheme is ap)lied.

(a) (b) (b)


Composite Sphere Particle Continuum

N~~7C o
Core

Shell 0

Level (I) Level (II) Level (III)


Figure 7-14: Multiscale microporo mechanics model for the hydrating MgO.

7.4.2 LEVEL I: The composite sphere

A relevant model to the hydration of reactive solids like cement-based materials and

Magnesium Oxide is the shrinking-core model of a composite sphere. The composite

sphere is made up of the anhydrous MgO grains forming the core and the Mg(OH) 2

forming the shell. Here we develop the evolution of the elastic properties of a colm)os-

ite sphere with a reactive core followed by multiscale imicroporomechanics of granular

solids.

The first level of the nmltiscale though-nodel is composed of anlhydrous MgO and

Mg(OH), and the volume fraction occupied by these phases is shown in Figure 7-10

as LEVEL I:

J = flid = fAJIO + fAIf((O)I2 (7.19)

160
To estimate the elastic properties and the evolution of eigenstresses in a composite

sphere, we start by a geometrical description of the composite sphere.

Eigenstrain in a rigid composite sphere

Volume changes of the core and the shell can be considered individually by applying

the divergence theorem. A rigid composite sphere is considered here with a shrinking

core and swelling shell (see Figure 7-15). The volume change rate to which the

core is subjected is defined in terms of the velocity vector at the core-shell interface

V - n = V, = dur
dt
with n e, the outward unit normal to the shrinking core surface

(see Figure 7-15):

dveo V - n dS= divV dV (7.20)


S Vco '9Vco(t) VCo0 v"

Similarly, one can apply a similar approach to the shell assuming continuity of the

radial velocity at the core-shell interface

V - n dS = -Vo + V -n dS (7.21)

where 8O
s(t) = &Ve(t) - WVco(t), with &Ve(t) the outer surface of the composite

sphere oriented by the unit normal n = er, whereas the inner surface of the shell

is oriented by a unit normal n = -e, (see Figure 7-15). Normalizing the previous

expressions by the total volume of the composite sphere, we obtain:

V - n dS = c~t- c +Dv (7.22)


V osa Ves dt

where D, is the total volume strain rate of the composite sphere. If we then define

the average volumetric strain rate of the shell, dh, by:

Vsh (t)dsh = VhV - n dS (7.23)

161
we obtain the consistencv equation:

v;--((t ) CO V1,(t)' ('1J (7.24)


Dv = d +
S I Cs

no
nr
O3Vcob
no aa
Ves r VCS+aVCo aVcs

=0Vsh

Figure 7-15: A schematic of the coniposite sphere of shrinking M gO (core) and


swelling M\g(OH) 2 (shell).

Upscaling

Following the framework developed in Section 5.1. eigenstress can be incorporated in

the nuiltiscale inicromechanics model at the scale of the composite spheres in level

I (Figure 7-17). The assumption of the rigid sphere that underlines the geometrical

description above does not account for confinement effects the shell exerts onto the

core and vise versa. The focus of the inicroporoniechaiiics model is to account for

the deforiability of the core and shell in this process. Consider thus a distribution

of eigenstresses in incremental form:

d, (z) = k(z)dc(z) + iog, (z) (7.25)

with the distribution of the bulk modulus and the incremental pre-stress given by:

K(z)
Vz Ce KC{ ( dcCO
(7.26)
k ( z-=SI 11 EI
Vz e VA Kal, Vz

162
where Ke, = 160 GPa is the bulk modulus of the core (MgO), Keh = 43 GPa is the
bulk modulus of the shell (Mg(OH) 2 ), Vc0 is the volume of the core, Vh is the volume
of the shell, and do& is the incremental eigenstress that develops in the solid phase
due to the progression of the hydration reaction. dEc, = dc 0dt and dEch = dshdt stand

for the volumetric strain increments related to the volume change rates previously
defined for the rigid system. Homogenization of Eq. (7.25) with Eq. (7.26) entails
the macroscopic incremental stress state equation:

dE' = KsdE[ + dEPI (7.27)

where E, is the macroscopic mean stress,dEv, is the incremental macroscopic volu-


metric strain, fc, is the volume fraction of the core and fsh is the volume fraction
of the shell (fc, + fah = 1). K,, is the homogenized bulk modulus of the composite
sphere shown in Eq. (7.35) and dEP is the homogenized pre-stress of the composite
sphere. They are obtained in the following way:

* From strain localization d&(z) = A(z)dEv, with A' 0 = (A)vco and A"h (A)Vh

are the volume averages of the fourth order strain distribution tensors, we obtain
the homogenized bulk modulus of the composite sphere:

Kes = (k(z)A(z))VS - VCO M KeOAvO + V (t)KshAh (


)

The strain compatibility condition requires dEl = (dE(z))vcs = fco(de(z))vco


+

fsh(de(z))Vsh <- Vco(t)Av+Vh(t)ASh -Vs. The bulk modulus of the composite


sphere can be rewritten in the form:

Vs a(t)
Kes = Kco + V ) (Ksh - Keo) Ash (7.29)
Ves

where Avh is the strain localization factor of the shell. It can be readily deter-
mined from the known bulk modulus expression for the composite sphere in Eq.
(7.35) below.

163
9 Evoking Levine's theorem. the macroscopic eigenstress increment for the coin-

posite sphere is obtained

Vjeb(t)
V. 0(t) A()dk(,oA M KV' sa~ (7.30)
dEv = (Kdu(z)4(z)) K// d Co

+
-
If we now enmploy Eq. (7.24) in Eq. (7.30) to replace dess in function of dc.,,
and evoke the strain compatibility condition /(.o(t)Ag() + 1j4 (t)AS' =V, we

obtain:
d EiP' = KIx,,deco + K. A E - dSil

)
(7.31)
= (-Ke + KJ;',,RlJ)) <E

where dE,' = Ddt = (16* = (Rip - 1)c is the swelling increment of the rigid
composite sphere model with D, given by Eq. (7.24) and d* is given by Eq.

(7.14) and dc.O = -d


.

F Incremental mean stress in the coiposite sphere (Level I)

dZ'o KesdE' + dEP' (7.32)

dE=(-Kx, + K A& Rjh) dj (7.33)

dE' = Rl1p 1) < (7.34)


-

It is iteresting to notice here the resemblance between Eq. (7.34) and Eq. (7.15).

The coefficient c was introduced to account for the reactionimecianism and the poros-

ity of the system. while the coefficient K is a consequence of the poroilmechaiiCs

formulation. These coefficient will be discussed further at the second level of the

miiicroporomnechanics model.

Elastic properties of a composite sphere

In this analysis. we estiimate the evolution of the elastic properties of a composite

sphere. For this plurpose, we comisider a single coimiposite sphere coisistimig of a spher-
ical core and a concentric spherical shell (see Fig. 7-11(b)). The core has a bulk

164
modulus Kc, and a shear modulus G,, with a radius a while the concentric shell
has a thickness b - a with a bulk modulus Keh and a shear modulus Goh. The bulk
modulus of the composite sphere is then [41-43]

Kcs = Ksh + fco(Kco - Ksh) (7.35)


1+ (1 - fco) [(Kco - Ksh)/(Kh + 3G(sh)

And the shear modulus can be obtained by solving the quadratic equation [19-21]

A GS) 2 + B G) +C =0 (7.36)
(Gsh Gsh

where

A= 8(4 - 5Vh) ?l 4fo/ - 2 (63/2774 + 271773) fci 3 + 252T127 4 fj 3


12
-50 (7 - Vsh + 8V2}h) 772714 fco + 4 (7 - 1OVsh) r72r73
m774fc/ 3 + 4(63r274 + 2 /1) fco/3 -
B = -4(1 - 5Vsh) 504r/2r74fi 3
(7.37)
+150 (3 - "sh) Vsah72rq4fco - 3 (7 - 15Vsh) 772773

C = -4 (7 - 5Vsh) 714fco - 2 (63rq274 + 2r 1r 3 ) fT3 + 252r72774 fo/3


-25 (7 - 12h) 77274 fco - (7 + 51,h) p2773

And
7
71 = (7 + 5vco) (7 - 10vsh) T14 + 105 (1co - vsh) = 100.52

772 (7 + 5vco) T14 + 35 (1 - vco) = 51.92


(7.38)
773 = (8 - Ovsh) 774 + 15(1 - 'sh) = 27.86

774 = (Gco/Gsh - 1) = 2.94

For a composite sphere of Magnesium oxide hydrating to form magnesium hydrox-

ide, the evolution of the bulk modulus is shown in Figure 7-16(a) and the evolution

of the shear modulus based on Eq. (7.35) and (7.36) is shown in Figure 7-16(b).

At level I, the eigenstress results from the hydration of the MgO grains to form

the MgO-Mg(OH) 2 composite sphere and it is due to difference in molar volumes.

The driving force of eigenstress at this level is the driving force of the hydration

reaction, which depends on the difference in chemical potential between the ionic

165
Emmme. - - .19NOW- - -

-
160 - - - - - --- - -1
140-
140
120
140 -K-G

-isics

100 80

80 60

60L 40

40~ 20
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a) (b)

Figure 7-16: Evolution of the bulk modulus of the composite sphere, where I.s
is the bulk modulus of the composite sphere. (b) Evohition of the shear modulus
of the composite sphere, where G,.s is the bu)1lk modulus of the composite sphere.
Calculations are based on analysis in Section 7.4.2. Material properties are shown in
Table 3.1

species in the pore solution and the solid MgO. The magnitude of eigenstress in
the composite sphere depends on the reaction mechanism and the microstructure

of the material. The reaction mechanism determines the fraction of the hydratioin

products that contribute to the radial growth and the fraction that precipitates in the
water-filled pore space and does not contribute to the eigenstress. The contribution

of the microstructure of the material to the development of eigenstresses is mainly

due to water access for the progression of the hydration reaction and the possible

rearrangement of the hydrating grains in the porous particles.

In case of a confined composite sphere with the reaction miechamisms resembling

the topochemical or surface reaction with radial growth only with unrestricted water

access to the MgO core, the eigenstress is only dependent on the chemical equilib-

rium between the pore solution and MgO, which also determines the degree of super

saturation with respect to MIg(OH)2. In this hypothetical. yet, unlikely situation. the

relationship between the eigenstress and the eigenstrain can be assumed linear and

the strain can be linearly decomposed into eigelistrain and elastic strain.

dE = IKdE - Kd (7.39)

166
In general, eigenstrain as estimated from the reaction stoichiometry (Eq. (7.14)

cannot be directly related to the eigenstress and a correction such as the one shown

in Eq. (7.15) is necessary. The correction is necessary to consider the contribution of

the reaction mechansi and the microstructure of the material as discussed further

at LEVEL II next.

dEIn=KcsdEv +dP'

a\
Core

Shell

Figure 7-17: A schematic of the composite sphere of MgO (core) and Mg(OH) 2 (shell)
with the mean stress resulting from homogeneous stress boundary conditions.

7.4.3 LEVEL II: The porous MgO-Mg(OH) 2 particle

The second level (level II) of this model includes the composite spheres of MgO-

Mg(OH)2 and the interstitial space referred to as the particle porosity in addition to

the eigenstress due to the progression of the hydration reaction (Figure 7-22). The

heterogeneous stress distribution in a simple porous solid composed of a solid phase

and saturated pore space with eigenstress can be written in incremental form:

(Vz E A,) &d, m(z) = k (Z)d (1Cz)+ dup, (z) (7.40)

where z is the position vector. a,(z) is the incremental heterogeneous mean stress,
A(z) is the bulk modulus. dE(z) is the incremental volunietric strain, and dcT< (z) is
the incremental mean eigenstress. For a porous iiedium under drained conditions

167
AIM, =K"dEv" +dIP'I

\/00 0Y
>0 00 0'~
Composite sphere
0 0 0"-
>0
T1-A
Figure 7-18: A schematic of the porous MgO-Mg(OH) 2 particle with the mean stress
resulting from homogeneous stress boundary conditions. The initial packing density
as measured with nitrogen sorption is 0.55.

and in the presence of eigenstresses, we shall apply the following distribution:

Kcs Vz E eCs dEPI Vz E V


k (z) = (7.41)
0 Vz E VP dppar Vz E Vp

where K, is the effective bulk modulus of the composite sphere, V,, is the volume of

the composite sphere (solids), Vp is the volume of pores and p is the pore pressure.

Homogenization of Eq. (7.40) with Eq. (7.41) entails:

dzfJ + dppar = Ke 8 (1 - b + (1 - b")(dZE' + dppar)


1)dE{' (7.42)

= K" dEf + (1 - b)(dEP' + dppar) (7.43)

where E is the macroscopic mean stress at level II, par is the pore pressure in the
porous particles, (dEm, +dp) is the incremental effective stress, dEf' is the incremental
macroscopic volumetric strain at Level II. dE' = (de(z)) = 1(de(z)) " +#(Kde(z)) is

the incremental macroscopic volumetric strain, rj is the solid volume fraction occupied
by the composite spheres and # is the porosity (7"7 = 1 -- (A)' and (A) "s are
().

the volume averages of the fourth order strain distribution tensors, dc(z) = A(z)dE.

168
The Biot coefficient at level II reads:

K"
I='(A(1-r= A)'c= 1- (7.44)
A cs

The level II state equation defines changes in stresses and particle porosity at constant

hydration degree. and read as:

Incremental mean stress in the porous particles (Level II)

KEI
do,,,,(z)) =KI'dEf' + (I - b") dZ'P' - b, dppar (7.45)

where Ppar is the pressure prevailing in the particle porosity.

At, level II, the manifestation of the strain (dEf') is critically dependent on the

reaction mechanism and the nicrostructure of the material as introduced in the dis-

cussion of Level I. Porosity of the porous MgO-Mg(OH) 2 particles plays a critical

role in the bulk volume changes at this level due to the possible rearrangement of the

hydrating grains during the hydration reaction. And porosity also provides additional

space for the hydration products to form without contributing to the radial growth of
the hydrating grains if a through-solution reaction controls the reaction mechanism.

In this model, it is assumed that the relative positions of the hydrating con-

posite spheres only expands without rearrangement of the particles. The lack of

rearrangement of the grains requires that the eigenstrain of the porous particles must

equal exactly the eigenstrain in the composite sphere if the entire hydration products

contribute to the radial strain of the composite sphere. The initial packing density

as measured with nitrogen sorption is 0.55 and in the case where all the formed

IVg(OH) 2 contributing to the eigenstrain, then this packing density will remain con-

staiit throughout the hydration reaction.

Due to the precipitation of Mg(OH) 2 in the water-filled pore space. the packing

density of the particle will increase at a rate proportional to the contributing fraction

and the non-contributing fraction of Mg(OH) 2 . This conclusion is supported by the

decrease of the surface area of the MgO particles and the merging of the hydrated

169
I
particles observed under the SEM as well as the small impact of the w/c ratio on the
measured Imulk volume changes. The prefactor c in Eq. (7.15) and (7.34) is requiired
to correct the eigenstrains in the material to account for this phenomenon.

Ppar ~
MgO Core

Mg(OH) 2
d Contributing

Mg(OH) 2
- -s Non-Contributing

Fignre 7-19: An ilinstration of the porous MgO-Mg(OH) 2 particle with the part of
Mg(OH) 2 contributing to the measured eigenstrain and the part that precipitates
inside the water-filled pore space withont contribution to eigenstrain. The initial
packing density as measured with nitrogen sorption is 0.55 and in the case where
all the fornied Mg(OH) 2 contributing to the eigenstrain. then this packing density
will remain constant throughout the hydration reaction. Due to the precipitation
of Mg(OH) 2 in the water-filled pore space, the packing density of the particle will
increase at a rate pIroportional to the contributing fraction and the non-contributing
fraction of Mg(OH) 2
.

Evolution of the poroelastic properties of the porous MgO-Mg(OH) 2 par-

ticles (Level II)

Elastic properties of the 1)0r0us particles (Level II) is calculated with both self-

consistent and Mori-Tanaka schemes. It is imperative here to estimate the evohition

of the packing density by taking into account the impact of the reaction mechanism. If

we assume that all the measured eigenstrain originate at this level. then the packing

density of the particle should be corrected accordingly. Initially the volume frac-

tion occulpied by the porois parti(les can be estimated directly based on the volune

fractions of the colmponents as follows:

a AJ0 (7.46)
glOr

I~
(

170
where fpar is the volume fraction occupied by the porous particles. This volume

fraction includes the porosity of the particle; rII = rpa, is the packing density of the

particle, that is the volume of solids in the particle relative to the volume occupied

by the porous particle. The volumetric strain at level II is defined as:

I VP ar -YP-a r - fpar 1 (7.47)


par fpar

with the volume fraction occupied by the porous particles:

fpar - flgO + fMg(OH) 2 _ (" 1 (7.48)


ripar

Finally, we obtain the packing density of the porous particles:

'ar= II = fMgo + f.Mg(oH) (7.49)


77par = CII + If 0 H) 2
~ )p'ar

The packing density of the fully hydrated particles of Mg(OH) 2 is r' = 0.88:

=II =pa 0 + 2max _ 0ar g (7.50)

with er0 = 0.55 and n4"aax = 0.88 based on the results of the nitrogen sorption

isotherms.

Again, the prefactor c in Eq. (7.15) and (7.34) is considered here to estimate

the evolution of the packing density of the hydrating MgO-Mg(OH) 2 particles. The

poroelastic properties at this scale are calculated with the self-consistent scheme and

read as:
K"GI b" =
= 1 - b = (7.51)
Kcs 4GII/Ges+ 3 (1 -qrs) r(

where r' = KcS/Gcs = 2 (1 + vC) /3 (1 - 2vc"), with j = rII the packing density of

171
the MgO-Mg(OH) 2 porous particles. and:

GII
C( 1 5 - I) - 3-r' (2 + r7)
(7.52)

+1 1144 (1 - rs) - 4807 + 400 02 + 408rsr/ - 120rs (2 + 9 () (2

+
with r = r" the packing density of the MgO-Mg(OH) 2 porous particles.

The second method employed to estimate the elastic properties at Level II is


the Mori-Tanaka scheme. In this scheme, it is assumed that the second level can be
modeled as a continuous matrix with the particle porosity. Based on the Mori-Tanaka
scheme (mt), the shear modulus is [61]:

GI_ ( _ _ + 9rs)
_(8
t =9 r) (7.53)
GCS 8 + 9rs +6(1 - rII)(2+ rs)

And the effective bulk modulus is:

-m, ," 4 4n,,( 7.54


Kcs 3(1 - r/I)r8 + 4
)
A comparison of the elastic properties obtained by the two schemes is shown in Figure
7-20.

7.4.4 LEVEL III: The continuum scale

The third level of the multiscale microporomechanics model for the hydration of
MgO consists of the first two levels (the composite sphere and the particle porosity),
in addition to the macroscale porosity. The continuum scale (Level III) is modeled
with two methods similar to level II.

To estimate the packing density of Level III, it is assumed that part of the
Mg(OH) 2 precipitates inside the porous particle foH)2 and the other part precipi-
tates in the water-filled porosity outside the original boundary of the porous particle

fMg(OH). -The fraction of Mg(OH) 2 that precipitates inside the particle can be related

172
150 120

100

7- 80 -GeS

60

50 - 40

020.
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a) (b)

Figure 7-20: Evolution of the bulk modulus of the porous particles (Level II). where
A" is the bulk modulus of the porous sphere estimated based on the self-consistent
scheme, K'/ is the bulk modulus of the porous sphere calculated from the mori-
tanaka scheme. and K,., is the bulk modulus of the composite sphere. (b) Evolution
of the shear modulus of the composite sphere. where Gc, is the bulk modulus of the
composite sphere.

to the packing density of the particle:

H - I/fparx (7.55)

and the part that precipitates outside the boundary of the particle is:

.t'g(O>11)2 =Mg(O1) 2 - fjg(OH) 2 (7.56)

The fraction of Mg(OH). that precipitates in the macroporosity is assumed to have

the same packing density of the porous particles. The packing density of level III is

then:
f 7' + J'g(OH) 2 __ .fl\g(oH-I) 2 + 'fNg (757)
7/= or + IIIII 7.57

The method employed to estimate the elastic properties of Level III is the Mori-

Tanaka scheme. In this scheme. it is assumed that the third level can be modeled

as a continuous matrix with micro porosity. Based oi1 the Mori-Tanaka scheme. the

173
shear modulus is [61]:

GJ,[ I/7 H (8 + 9r1-) (7.58)


G" 8 + 9,r, + 6(1 - VIT")(2+ r,)

where r, = K''G'. The effective bulk modulus is:

KO (3 ( 14(i> (7.59)
K11 (3(1 -- )! s+ 4)

I
I '4 14 ~- ---
GI
-

14 12, -GS

~12
Z110

8 6

6 4
-KI!
0.2 0.4 06 0.8 1 0.2 04 0.6 0.8 1
0
(a) (b)

Figure 7-21: Evolution of the bulk modulus of the continuum (level III). where K"'
is the bulk modulus calculated based on the Mori-Tanaka scheme at level III and
the self-consistent scheme at level II. Kf/ is the bulk modulus calculated from the
mori-tanaka scheme at both level II and level III. (b) Evolution of the shear modulus
at Level III.

9
At this level, we consider a second pore space. the macroscale porosity. in which

another pressure prevails, denoted by p and the porous MgO-Mg(OH) 2 particles make

up the solid phase. The vohune occupied by the hydration products is designated

,r and the space occupied by the capillary porosity is ca.

The incremental stress state equation for the stress distribution at level III rcals:

d(IU, (z) = k(z)dc(z) + du')(z) (7.60)

174

- -- I
with:

k(z) { K"

0
Vz E

Vz E
Vpar

VCa
(7.61)

do") (z) { dB"=(1 b") dYZP' - b" dppar Vz E Vpar

-
(7.62)
-dp Vz E Vcap
The eigenstress at Level III is obtain by application of Levin's Theorem:

dEP"' = A(z)daP(z)) = (1 - b{") dEZ" - b'"dp (7.63)

Whence, the incremental equations of state:

dEf; = (doa, (z))

= KN',dEt + (1- b4" ) (1 - b") dEP' - (1 - b ") b(1(ipar


1 - bdp
!

dD' 1
(7.64)
where:

K", = (1I - b, 11) K'" (7.65)


1K" (7.66)
- <pfjAf" =- 1
-

K
(I - b"') (I - b") = S (7.67)

For identical pressure in the capillary porosity' and the particle porosity dpr = (1).
we have:

Incremental mean stress at the continuum scale (Level III)


I

11 (1) (7.68)
dEi, = do, (z)) = K'"dE " + (1 - b') dEP'
-

11
dEP

where
K"'
b'' - (1 - b(') b" + b'' = 1 T(7.69)

175
- - IMMEMIN--

2K4
----
L4L--
LJ---1":4--
K"

Ppar

Figure 7-22: A schematic of the continuum scale for the hydrating NgO-\lg(OH),
porous solid.

Under constant pore pressure. the volumetric strain is:

dEf"= - ( R- - 1d (7.70)

7.5 Comparison with experimental measurements

Bulk volume changes during the hydration of IgO is investigated in the Rubber

Diaphragm Configuration (1 DC) and the Liquid Barrier Configuration (LBC). This

section presents a discussion of the results obtained with both boundary conditioins.

7.5.1 Hydrostatic state of stress (RD test)

In the rubber diaphragmn test, the bulk volume changes of the composite systeni are

mleasured in a stress-free sample. We recall Eq. (5.104):

d~aD = (1 1)11) 1 1 K (771


d ERD (KI - _ KW R/\i]) - 1,d)7.

176
A comparison between the experimental results and the volumetric strain predicted

by the icroporonmechais model is shown in Figure 7-23. Considering. the corrected

eigenstrain c7* = (cR 1 , - 1) . the eigenstrain corrected for the effect of the reaction

mechanism and the microstructure of the material is defined in Eq. (7.15). Based on

the results presented in Section 6.1, the expansion rate in the rubber diaphragm test

was dc = 0.015d = (cR;,, - 1)d and the fitted value for coefficient c = 0.459. This

implies that only 46% of M~g(OH) 2 contribute to the radial growth of the composite

sphere as illustrated in Figure 7-19.

It is also possible to obtain the same expansion rate of MgO-Mg(OH) 2 considering

the expansion of the porous particles as discussed in Section 7.3.3. Adopting this

analysis, the eigenstrain at the second level (porous particles) can be related directly
to the measured bulk volume change at the macroscopic scale. The development of

eigenstress as calculated with Eq. (7.71) is shown in Figure 7-24.

- Ev (X) (MgO-LBC-0 5) - - - Corrected Model


Ev (X) (MgO-RDC-06)
12 - 10

10 - --

E 8

4 - - - - _

0.2 0.4 0.6 0.8 1

Figure 7-23: A typical vohunetric strain as measured in the rubber diaphragin test
(RDC) and the liquid barrier test (LBC) initiated at ( 0.3 with wc/c = 1 and the
corrected model with c = 0.46.

177
0 0.5 1
0

-200

-400

-600

* -800

-1000

-1200

Figure 7-24: Development of eigenstress during the hydration of MgO with w/s = 1.0
at 25 0 C

7.5.2 Oedometric state of stress (LB test)

Unlike the RDC test, the sample in the LBC test is constrained radially, and all

the bulk volume changes are manifested in the axial direction. Based on the elastic

solution of the volume changes one can calculate the volume changes as:

LB (1 - b( K"'I dZE'
(7.72)
zz o K+{GJ kK"' +GIII} Ke8
/ KIII \A/-K AV,
(7.73)
(K"'IK!!!
+ !G4 ~"" Rhp - 1)d
(KcARdR

1(+ v"1I dERD


(7.74)
3 1 i-v"

The ratio R shown in Figure 7-25 represents the ratio of the bulk volume change in
the liquid barrier test compared to the rubber diaphragm test. The results presented
in Section 6.1 and Eq. (7.74), where the ratio is measured to be 1/3, show that
the deformation of the hydrating MgO sample in the liquid barrier test may include
viscous deformations (creep). Experimentally:

R dELB 0.0048
, __ =__ = 0.32 (7.75)
dER 0.015

The main difference between the RDC and the LBC tests is the displacement

178
0.225 0.53

0.22

0.215 0.52-

0.21 ---
0.51
0.205-

0.2- 0.5
0.195/ RR
0.19 SI 0.49S
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a) (b)

Figure 7-25: (a)Evolution of the Poisson's ratio at level III. where VI4/ is the Poisson's
ratio calculated based on the Mori-Tanaka scheme at level III and the self-consistent
scheme at level II, ij is the Poisson's ratio calculated from the mori-tanaka scheime
at both level II and level III.. (b) Evolution of the ratio between the stress-free defor-
mations and the axial deformations in the oedometric boundary conditions assuming
elastic deformations only with R = 1+v
:3(1 -v)

boundary conditions. In the RDC test, the sample is maintained saturated under

hydrostatic state of stress with very small and constant effective stress of 3 kPa for

samples with iw/c = 1.0 and 64 kPa for samples with w/c < 1.0. The small values

of effective stress are required to maintain contact between the rubber diaphragm

and the sample but not large enough to cause significant creep. Consequently, the

measured strains can be attributed solely to the development of eigenstresses in the

samlple. The measured macroscopic strains are then related to the eigenstress through

the nmultiscale microporomechanics model with a correction of the eigenstrain to ac-

count for the contribution of the reaction mechanism.

In the LBC' test, on the other hand, the sample is radially constrained and the

bulk volumne changes are only allowed in the axial direction. The impact of the radial

Confinement on the measured strains in purely elastic deformations requires that the
measured strain be a fraction of the strain ieasured in the RDC test proportional

to the Poisson s ratio of the composite naterial. A Poisson's ratio of about 0.2 as

obtained from the nmultiscale microporomnechanics model corresponds to a ratio of


1+1 = 0.5 between the RDC amid the LBC volume changes. The ratio measured
3(1-
179
experimentally of about 0.3 corresponds thus to a Poisson's ratio of zero.

This discrepancy can be due to assumptions of the micromechanics model, espe-

cially, the assumed perfect contact between the components of the composite sphere

and the distribution of porosity. The distribution of porosity in the model is based

on the porosities calculated from nitrogen sorption isotherms, which does not distin-

guish between particle and macroscale porosity, while the porosity of the second level

is assumed to increase from the porosity of the MgO particles (0.45) to the porosity

at complete hydration of 0.88.

7.6 Summary

Bulk volume changes during the hydration of light-burn MgO has been investigated

with multiple techniques and the resulting understanding of the material was then

used to build a multiscale microporomechanics model to interpret the experimental

results. Hydration kinetics was monitored with both enthalpy and volume of the

reaction as well as thermogravimetric analysis. Back-scattered and secondary electron

images were collected for the MgO and Mg(OH) 2. Combined with sorption isotherms,

these results were used to construct a multiscale thought-model of the MgO-Mg(OH) 2

porous solid.

Expansion has been attributed to the hydration of MgO and the molar volume

difference between the reactive solids and the hydration product with a correction for

the reaction mechanism. The correction is based on the assumption that part of the

hydroxide precipitates in the water-filled pore space and does not contribute to the

measured eigenstrains and the other part precipitates between the hydrating particles

of MgO and contributes to the eigenstrain.

180
Chapter 8

Eigenstress Development During

C3 S Hydration

One of the materials investigated in this study is the monoclinic C 3 S with three
different particle size distributions designated C 3 S-F, -M, and -G for fine, medium,
and coarse PSDs as detailed in Section 3.2. The reason behind the choice of C 3 S
stems from the fact that it constitutes the main phase in cement-based materials and
the purity of the phase facilitates the interpretation of the measured quantities and
helps understand the more complex Class-G cement. Another advantage of the study

of C 3 S is the absence of sulfate and aluminate phases and the absence of hydrous
phases like AFt and AFm, to which expansion is usually attributed. This chapter
discusses the results obtained from the hydration of C3 S and details the development
of a multiscale microporomechanics model for the interpretation of the experimental
data.
The first part of this discussion focuses on the hydration kinetics and development
of the microstructure during the hydration of C 3 S with the goal of building the un-
derstanding necessary to construct a multiscale thought-model for the hydration of

C3 S. Hydration kinetics is monitored with both enthalpy and volume of the reaction
and the microstructure is investigated with SEM and sorption isotherms.
The second part of this discussion focuses on the development of a multiscale
thought-model for hydrating C 3S to extract the elastic properties and establish the

181
relationship between the measured macroscopic quantities and its microscopic origin.

The relationship between the macroscopic quantity and the driving force is then

discussed in relation to the colloidal forces in C 3 S paste and the associated kinetic

processes.

8.1 Analysis of kinetics of C 3 S hydration

The rates of the enthalpy of the reaction are reported in Figures 6-7 and 6-9 for three

particle size distributions as discussed in Section 3.2. A typical hydration curve is

characterized by the induction period followed by acceleration and later deceleration

periods from both the enthalpy and volume of the reaction. For different particle

sizes, the peaks of the hydration rate occur for all particle sizes at about 7.5 hours

into hydration and at degrees of hydration ranging from 0.09 to 0.18. Degree of

hydration is calculated assuming a total enthalpy of 520 J/g of C 3 S [182] starting

at 0.5 hr after mixing with water, where the 0.5 hr is required to establish thermal

equilibrium in order to collect meaningful data. The fact that the peak occurs at the

same time indicates the initiation of diffusion-limited reactions when the thickness of

the hydration products forming on the surface of the C3 S reaches a critical value. The

critical thickness of the hydration products seems to be independent of the particle

size distribution.

Unlike the effect of the particle size distribution, the curing temperature acceler-

ates the reactions and shifts the time of the main peak as observed in Figures 6-12 and

6-11 from both enthalpy and volume of the reaction. Such behavior is best captured

by Arrhenius-type equations to relate the rate of hydration d</dt at any temperature

T to a reference rate at a reference temperature. This is achieved by introducing the

hydration kinetics law [87,197-199]:

d< = A( )exp (- = 7W-1( , T) (8.1)

where A( ) is the normalized affinity that relates the rate of hydration to the degree

182
of hydration , Ea is the activation energy, R is the universal gas constant, and T

is the absolute temperature. The activation energy and the normalized affinity are

determined from a series of isothermal calorimetry tests on C 3 S paste with different

w/c ratios and curing temperatures as listed in Table 8.1. The normalized affinity is

fitted with a function of the form [87]:

1 - exp(-a) (8.2)
1 + b~c

where a, b, and c are fitting constants. Figure 8-1 shows the normalized affinities

determined from the isothermal calorymetry tests carried out at different temperature

for C 3 S paste with variable w/c ratios and PSD.

The calculated activation energy is in good agreement with published data (see

for example [56]).The small variation in normalized affinity in function of curing

temperature shows that A() depends primarily on the degree of hydration. It is

also evident from Figure 8-1 that the PSD can have a significant impact on the

measured activation energy. The effect of PSD is expected due to the differences in

the thickness of the diffusive hydration layer forming around the hydrating cement

grains. The thickness of the hydration layer can increase the activation energy of the

coarser PSD due to the diffusion control over the hydration kinetics.

Table 8.1: Summary of affinity fitting results.

T Ea
Sample a b C (oK) (J/mol)
C3 S-F-W49-T25 15.40 393.54 5.10 298.15 27661
C3 S-M-W42-T25 18.25 673.04 4.17 298.15 28329
C3 S-M-W42-T35 18.78 1069.69 4.16 308.15 28023
C3 S-M-W42-T45 21.54 1529.88 4.17 318.15 27880
C 3 S-G-W34-T25 32.04 4399.30 4.72 298.15 29363
C3 S-G-W34-T35 30.11 3386.85 4.29 308.15 29172
C3 S-G-W34-T45 37.69 5019.01 4.30 318.15 29091

As expected and due to purity of the material, the relationship between enthalpy
and volume of the reaction can be considered linear within the experimental error (see
Figure 8-2 (a)). In fact, the effect of the particle size distribution on the relationship

183
1
- -A(k) (C3S-F-W42-T25) - -A( ) (C3S-M-W42-T25)

A(k) (C3S-M-W42-T35) --- A( ) (C3S-M-W42-T45)


(C3S-G-W42-T25) A( ) (C3S-G-W34-T35)
A(k) (C3S-G-W34-T45)

1
0.9
0.8
%

0.7 -- -% V

0.6 -/\ ---


I" r
0.5
0.4
-

0.3
0.2
0.1
0
0 0.2 0.4 0.6 0.8

Figure 8-1: Normalized affinity determined from isothermal calorimetry of C:S pastes
carried out at different temperatures with different particle size distributions and w/c
ratios. The resulting fitting parameters for the affinity and the activation energy are
shown in Table 8.1.

184

____
between enthalpy and volume of the reaction is minimal and the relationship remains

approximately linear throughout the measurement as shown in fugure 8-2(a). The

effect of temperature on the other hand, deviates from linearity as the temperature

is increased as shown in Figure 8-2(b).

The reduction of the volume of the reaction at higher temperatures has been ob-

served experimentally (see Figure 8-3) by other researchers [82,132, 213], and it can

be attributed to changes in the surface area of the main constituent C-S-H as pro-

posed by Ulm et al. [195]. Acording to Ulm et al. part of the volume of the reaction

(chemical shrinkage) can be attributed to the difference in absolute volume between

the reactants and the products and he referred to this part as the stoichiometric sink

term. The other part of the volume of the reaction was attributed to the electrostric-

tion resulting from the adsorption of water to the surface of the solids and referred

to this part as the adsorption sink term. Such concept can also be found in the work

of Powers and Brownyard [157] and the work of Thomas et al. [185]. Powers assumed

a density for water in different categories of capillary and gel water based on isobars

and isotherms and estimated the density of the adsorbed and interlayer water at ap-

proximately 1.1 g/mL and Thomas et al. used similar density for the adsorbed and

confined water.

8.2 Microstructure of C 3S

The microstructure of C3 S is investigated with Scanning Electron Microscope (SEM)

and nitrogen sorption isotherms. The goal of the microstructure investigation is to

provide a better understanding of the microstructure of the material necessary to

interpret the experimental data and to build a proper multiscale thought-model for

the material and calculate the developing eigenstresses in the hydrating C3 S paste.

We start by developing an understanding of the pore size distribution and spatial

distribution of the solid phases during the hydration reaction and the relation to the

degree of hydration.

185
400 350
350 300
300
250 ---

-
250
200
200 x
150
150

100 M 100 (S-F-W49-T25)


(C3S-M.W42-T25) -(C3S-M-W42-T25)
-{C3S-G-W34-T25) 50
50 -(C3S-M-W42-T35)
(C3S-M-W42-T45)
0 0
0 10 20 30 40 50 0 10 20 30 40
Vrxn (gL/g) Vrxn (pL/g)

(a) PSD (b) Temperature

Figure 8-2: Relationship between volume and enthalpy of reaction as measured with
isothermal calorinetry and the pressure vessel. (a) For samples with different PSD.
the relationship is very close to linear with a, ratio of 8.89, 8.78, and 8.34 J//pL for
the fine, medium and coarse materials, respectively. (b) The relationship between
the volume and enthalpy of reaction as measured with isothermal calorinetry and
the pressure vessel with medium particle size distribution (C3S-M). The relationship
0 0
is very close to linear with a ratio of 8.78. 9.22. and 10.04 J/pL for the 25 C, 35 C
and 45'C curing temperatures, respectively.

+- 10"C, -AV/( = 0.1007

-0.005 -4- 25'C. V/ 0.07641


-- 40*C, -AV/u = 0.0668
-0.01
-- A-. 60*C, -AV/ = 0.0554
E
-0.015

0) -0.02

-0.025
.C
U A
-0.03
Class H, w/c=0.35
up to 3 days hydration
-0.035
0 0.1 0.2 0.3 0.4 0.5 0.6
Degree of Hydration, a,

Figure 8-3: Relationship betweeni volume of reaction (chemical shrinkage) and degree
of hydration as mneasured by TGA during the first 72 h of hydration of Class-H cement
with w/c=0.35 at T=10. 25. 40. and 60'C. From [213].

186
8.2.1 SEM

Figure 8-4 depicts backscattered electron images of C 3 5 paste made of three different

particle size distributions prepared at w/c=0.42 and hydrated in moist conditions


for 30 days at 25 C. The instrument used for this investigation is Zeiss MERLIN
high resolution scanning electron microscope. Backscattered electron micrographs
are produced at 15kV acceleration voltage for the three samples with different PSDs.
It can be seen in the micrographs of the fine and the medium-sized C3 5 pastes that
clusters of CH (light grey) form at few centers encapsulating some of the C 3 5 grains.
Such observation was made by Kjellsen et al. [109] who observed that these clusters
of CH tend to grow larger during the hydration reaction until they impinge to form
even larger CH masses. Kjellsen et al. attributed the growth of such masses in C3 5
paste (but not in OPC pastes) to the composition of the pore solution. The higher
pH found in OPC due to the common ion effect of alkalis. Centers of CH can also
be observed in the work of Thomas et al. [186] in C3 S paste mixed at w/c=0.5 and
hydrated for 28 days, while they disappear in samples seeded with 2% C-S-H seeds
under the same conditions. In this study, these centers of CH were not observed in
the C 3 S paste made with the coarse PSD. In all samples, the C-S-H (dark grey phase

in the BSE images) forms a continuous matrix in which CH and the residual C3 5
grains as well as the capillary porosity are embedded.

8.2.2 Sorption isotherms

Sorption isotherms were collected for a set of C 3 S samples with different PSD at
different ages cured at 25'C. All samples were prepared at w/c = 0.42 in moist
conditions in plastic tubes with open ends. The resulting sorption isotherms are
shown in Figures 8-5, 8-6, and 8-7.

The measured BET surface area (SSA) and pore volume (VP) are attributed
mainly to C-S-H. The normalized values are shown in Table 8.2, where the mass
fraction of C-S-H is calculated based on the volumetric relationships derived in Sec-
tion 8.2.3. The variation of the pore volume for pores up to 300 nm in diameter

187
Lfor

2Oj~m lOml
F-H__ ____ _ -4

CMSE
I'iF
C F 1 T
WI2 1) C $~-F-\V L~ T2~ *\~h1 :1

jIjfliCME _______________ ___ .~FF I~iiCMSE

WK\T NV W GT: A:*0( (1) CaS-\ -N 1? 7T? N 1 :t at 2ThK

rIT.
. L DN IIIii~cMsE
-

(e) C 1 :S-G-W42-T25-A3od at 2U0X (f) C: 1S-G-W42-T25-A3od at 50(X

F1igu11 S-4: SEM mi1crographs of the nIi(rOStr1.(tur of CE:S hydlr atd for 30 davs at
(lifflrenlt migli2 I( fict1ions with w/-(.42. White (Bright) is 1tle unhydiated( C'S. tight
rey l rprscnits porthanite, dark grey is the C-S-H gieI. ad blck is1 porosity.

158
indicate an increase in the pore volume as the mean particle size increases. The evo-

lution of the SSA and the pore volume tend to remain constant during the first 16

days followed by a sharp decrease at later ages. This finding agrees with the find-

ings of Jennings and Tennis [100], where the fraction of C-S-H accessible to nitrogen

decreases during the hydration of C 3 S paste.

The decrease of the pore volume and the SSA accessible by nitrogen can be a

result of the densification of the C-S-H gel. The densification is also discussed in the

following sections as a possible driving force of the bulk volume changes.

Table 8.2: Summary of sorption results for C 3 S samples.

SSA VP C-S-H C-S-H SSA C-S-H VP


Sample Age (d)
(m 2 /g) (cm 3 /g) mass fraction (m2 /g) (cm 3/g)
C 3 S-F-W42-T25-A8d 8 130.60 0.160 0.73 0.36 359 0.439
C 3 S-M-W42-T25-A8d 8 103.90 0.144 0.60 0.30 348 0.481
C 3 S-G-W42-T25-A8d 8 83.60 0.126 0.45 0.22 379 0.569
C 3 S-F-W42-T25-A16d 16 163.69 0.173 0.79 0.40 414 0.438
C 3 S-M-W42-T25-Al6d 16 132.10 0.164 0.68 0.34 390 0.483
C 3 S-G-W42-T25-Al6d 16 106.40 0.148 0.51 0.25 422 0.586
C 3 S-F-W42-T25-A32d 32 42.60 0.095 0.83 0.42 102 0.230
C 3 S-M-W42-T25-A32d 32 27.20 0.092 0.77 0.38 71 0.240
C 3 S-G-W42-T25-A32d 32 13.80 0.074 0.62 0.31 45 0.239
C 3 S-F-W42-T25-A80d 80 72.40 0.104 0.89 0.45 162 0.234
C 3 S-M-W42-T25-A80d 80 60.50 0.099 0.83 0.42 146 0.239
C 3 S-G-W42-T25-A80d 80 37.76 0.084 0.68 0.34 111 0.247

189
- Q- C3S-F-W42-T25-A8d - Q- C3S-F-W42-T25-Al6d
-Q- C3S-M-W42-T25-A8d -Q- C3S-M-W42-T25-Al6d
- Q- C3S-G-W42-T25-A8d Q- C3S-G-W42-T25-Al6d
450 450
400 400
H
CA
350 350
300 E 300

~0 250 250
-

200 200
0
~0 150 150
100 100

50 50
01
0 0
-

0 0.5 1 0 0.5 1
Relative Pressure (P/Po) Relative Pressure (P/Po)

(a) Age= 8 days (b) Age= 16 days

Figure 8-5: Nitrogen sorption isotherms of C:3S paste hydrated for (a) 8 days and
(b) 16 days in moist conditions with w/c = 0.42 and cured at 25'C. Sunnnary of
properties are given in Table 8.2.

- Q- C3S-F-W42-T25-A32d - Q- C3S-F-W42-T25-A80d
- Q- C3S-M-W42-T25-A32d -Q- C3S-M-W42-T25-A80d
Q- C3S-G-W42-T25-A32d Q- C3S-G-W42-T25-A80d
180 180
160 H1- 160

140 140
120 cu 120
-C

----- I- EM
100 - - 100
Ln
80 80
60 60
40 40
20 20
0 0
0 0.5 1 0 0.5 1
Relative Pressure (P/Po) Relative Pressure (P/Po)

(a) Age= 32 days (b) Age= 80 (lays

Figure 8-6: Nitrogen sorption isotherms of C:3S paste with differeilt particle size
distributions hydrated for (a) 32 days and (b) 80 daYs in moist conditions with u/c
0.42 and cured at 250 C. Sunnary of properties are given in Table 8.2.

I 90

_____
Q- C3S-F-W42-T25-A8d - Q- C3S-G-W42-T25-A8d
- Q- C3S-F-W42-T25-Al6d - Q- C3S-G-W42-T25-Al6d
Q- C3S-F-W42-T25-A32d Q- C3S-G-W42-T25-A32d
- Q- C3S-F-W42-T25-A80d - Q- C3S-G-W42-T25-A80d
350 - 450

-
H 400
H) 300

-
-

350

-
250 E
-

300

-
E 200 w 250
150 C 200 -/

-
-

V)
150
100

-
-

100-
50- C

01
0 0
0 0.5 1 0 0.5
Relative Pressure (P/Po) Relative Pressure (P/Po)

(a) Age= 32 days (b) Age= 80 days

Figure 8-7: Nitrogen sorption isotherms of C3 S paste made with (a) fine PSD and (b)
coarse PSD hydrated for different periods of time in moist conditions with w/c = 0.42
and cured at 25'C. Summary of properties are given in Table 8.2.

8.2.3 Volume fractions

At a given degree of hydration, the components making up the C 3 S paste include the

residual anhydrous C:3S, portlandite (CH), and C-S-H as well as the gel porosity and

the capillary porosity. Gel porosity is defined here as the part of the total porosity

of the system encapsulated inside the C-S-H gel with a maxiimuim characteristic size

of tens on ilanomleters. it is onlv indirectly affected by the w/c ratio and the curing

temperature as will be discussed later on. The capillary porosity is also assuied to
be encapsulated inside the C-S-H gel: but it has a significantly larger characteristic

size than the gel porosity and depends directly of the /tvcratio.
Unlike the hydration of MgO discussed in the previous chapter, the iolar volumes

of the iain hydration product (C-S-H) depends on the degree of saturation of the

C-S-H gel. As discussed in Section 8.1. the surface area has a significant impact, on

the imolar voluines and it should be considered when calculating the volume fractions.

WVhen calculating the evolution of volume fractions during the llydration of C:3S. it is

critical to consider the appropriate density for the C-S-H phase. The density of the

C-S-H gel as measured experimentally by many researchers (see Table 8.3) is shown

191
to depend on the degree of saturation of the gel with water. To calculate the density
of the gel, we start by classifying water in the C-S-H gel into two types, the first
type constitutes part of the gel water and has the same properties as the bulk water

(Pw ~ 1.0 g/mL) and this water is held inside the C-S-H gel by capillary condensation
and is referred to as the free or gel water [78].

The second type of water is the water that constitutes part of the C-S-H solid and
is referred to as the combined water. Fujii and Kondo [78] investigated the relationship
between the C/S ratio and H/S ratio and the combined water content was estimated as
the difference between the total water and the free water as determined by extraction
with ethyl methyl ketone. The composition of the C-S-H solid as determined by Fujii
and Kondo was CnSH(no.s), where n > 0.8 is the C/S ratio and it was calculated
from solubility data. For a C/S ratio of 1.7, the composition of C-S-H would be

C 1. 7SH 2 .5 , where 2.5 is the combined water. Sierra [174] examined the different forms
of the combined water in the C-S-H solid by molecular spectroscopy and thermal
analysis and suggested a distribution of water between hydroxylic, interlayer, and
surface water. The hydroxylic water is the part of the water that decomposed into
hydroxyl groups and became part of the C-S-H solid. The interlayer water is the
water confined inside the C-S-H solid layers and the surface water is adsorbed on the
surface of the C-S-H solid and both interlayer and surface water are characterized by
a density larger than 1 g/mL due to electrostriction of water.

Based on this classification of water in the gel and as relevant to calculation of


volume fractions and volume of the reaction, we distinguish three conditions of the
C-S-H gel based on the degree of saturation (water content). The first condition is
the completely saturated C-S-H gel where the gel porosity if filled completely with
water. The density of the gel in this condition is the density that is used later on
in the multiscale thought-model and it also defines the packing density of the C-S-H

gel.
The second condition is defined at the degree of saturation where part of the gel
water with bulk water properties (free water) is removed and only combined water
remains. The density of the C-S-H gel in this condition is related directly to the

192
specific surface area and surface charge of C-S-H. The importance of this condition

is related to the average density of the hydration products and the measured volume

of the reaction.

The third condition is achieved when free and adsorbed water are removed and

only the dry C-S-H solid remains with hydroxylic "water" and interlayer water only.

This condition defines the density of the C-S-H solid as used for the calculations of

packing density of the gel.

To estimate the composition of the C-S-H necessary for the calculation of the

molar volumes and water content of the C-S-H gel, we consider here two models for

the nanostructure of the C-S-H gel. Glasser et al. [83] suggested a simplified model

that was developed for the thermodynamic treatment and dissolution equilibria of

C-S-H and is represented by the formula:

CaH 6 -2xSi 2 O 7 - zCa(OH)2 -nH 2 0 (8.3)

The C/S ratio is given by:


Ca/Si = X (8.4)
2

This model is applicable to young C-S-H gels with C/S ratio ranging from 1.0 to 1.4.

The value 3+ z - x is related to the water bound as hydroxyl groups; and n refers to

the interlayer water and surface water. The requirement of the value of z is related

to the silicate chain length as will be seen next.

The second model considered here is the generalized nanostructure model devel-

oped by Richardson and Groves [165,166]. The model suggests a generalized formu-

lation for the nanostructure of C-S-H based on a modified Tobermorite/Portlandite

(T/CH) and Tobermorite/Jennite (T/J) structural viewpoints. Richardson and Groves

gave the formulation of the C-S-H structure as:

{Ca 2nHwSi( 3n-1)O(9n- 2 )} - (OH)w+n(y- 2) - Cay2 - mH 20 (8.5)

where w is the number of silanol groups, (3ni - 1) is the silicate chain length, and the

193
calcium to silicon ratio is given by

n(4 +y)
Ca/Si = n(4 + 1) (8.6)
2(3n - 1)

For a Ca/Si=1.7 and assuming that the dimeric silicate chains are the predominant

form in young C 3 S or OPC paste [83], n = 1, y = 2.8

The density of the nanoscale C-S-H gel particles ("solid" C-S-H) including the

hydroxilic water and the inter-layer water is assumed constant and independent of

the curing conditions and the degree of saturation of the C-S-H gel. Allen et al. [2]

combined measurements from small-angle neutron and X-ray scattering to measure

the average composition and density of the C-S-H solid. They considered C-S-H

particles including all the physically bound water in the internal structure of the

particles and excluding the free and adsorbed water. After Allen et al., the density

of C-S-H is taken as 2.60 g/mL for an average composition of C 1 .7 SH 1 .8

.
Young and Hansen [212] and Thomas et al. [184] estimated the density of the

C-S-H particles including the interlayer water and adsorbed water as measured with

pycnometry and neutron scattering as 2.18 g/cm 3 with an average composition of

C 1 .7 SH 2 . 1 . It is imperative to mention here that this density and composition include


the water adsorbed on the surface of the C-S-H solid and consequently, it depends on

the specific surface area of the solid. A compilation of C-S-H density measurements

are listed in Table 8.3 for different degrees of saturation and drying conditions.

In conclusion, the composition and density of the C-S-H phase used to calculate

the enthalpy and volume of the reaction must include, at least, all the combined water

(hydroxylic, interlayer, and adsorbed water) due to the contributions to enthalpy and

density of the hydration products. The minimum H/S ratio required for the balanced

chemical reaction is H/S=2.1. For the calculations of the packing density of the C-

S-H gel on the other hand, the composition and density must include the water that

defines the C-S-H solid, which in general is considered as the solid with the interlayer

only. An average density suitable for such calculation is found in the work of Allen
3
et al. [2] with an average density of 2.6 g/cm and a compositions of C 1 .7SH 1 .8
.

194
Before we embark on calculating the volume fractions of the different components

of the C 3 S paste, we start by introducing the necessary definitions of the volume

fractions and densities occupied by the C-S-H phase.

PCSH: is the density of the C-S-H solid including all the combined water and excluding

the free water. This density depends on the specific surface area, but does not

depend on the packing density of the C-S-H gel.

fCSH: is the volume fraction occupied by the C-S-H solid including the hydroxylic
water and interlayer water and excluding the adsorbed (surface) water and the

free water. This volume fraction is assumed to be independent of the surface

area and the packing density of the gel.

fCPsH: is the volume fraction occupied by the C-S-H solid including all the combined
water and excluding the free water. This volume fraction depends on the specific

surface area. But does not depend on the packing density of the C-S-H gel.

Pgel: is the average density of the C-S-H gel including all the combined water and

the free water inside the C-S-H gel (gel water). This density depends on the

packing density and specific surface area of the C-S-H phase:

Pgei = fipi = Tge1fCS + 'PgelPw (8.7)

where Pgei = 1 - 7lgel, and pw = 1.0 g/mL is the density of the free gel water.

fgel: is the volume fraction occupied by the C-S-H gel including the combined water

and the free water inside the gel porosity (the saturated gel). This volume

fraction depends on the packing density and specific surface area of the C-S-H

phase:

fgei = fCSH (8.8)


"7gel

To estimate the evolution of the volume fractions of the different components of

the system, we consider the chemical reaction below along with material properties

195
in Table A.2. The reaction is based on the formulation of Young et al. [212].

C 3 S + (3-x+y)H , - CxSHy + (3-x)CH (8.9)

where x is the C/S ratio and y is the H/S molar ratio.

The volume fractions of the different components are then:

frxn = (Rhp - RH - 1) (1 - no)( (8.10)

fc 3 s = (1 - no)(1 - (8.11)

)
fmw = no - (1 - no)RH (8.12)

fcsH = (1 - no)RcSH (8-13)

fcH = (1 - no)RCH (8.14)

fgel = (1 - no)Rei (8.15)

The total porosity of the system is:

di,=1 - f - fbp
(8.16)
= no - (1 - no)(Rhp - 1)

while the capillary porosity is:

Ocap = 1 - f(cs) - fgel - fCH (8-17)

Finally, the gel porosity is:

geI -- Otot - Ocap (8.18)

where Rh - (3= I H VCsH is the ratio between the molar volumes of the hydration
products and that of C 3 S, RH - (3-xy)V, is the ratio between the molar volumes

of water and the reactive solid (C 3 S), RCSH is the ratio between the molar volume
of C-S-H and C 3 5. frxn is the fractional volume of the reaction, f(c 3 s) is the volume

196
fraction occupied by the remaining solid non-porous CS, fm" is the volume fraction

occupied by the remaining initial water used for mixing (initial water is treated as

a separate phase just like solids), fCSH is the volume fraction occupied by C-S-H

as defined by RCSH, 0 is the initial porosity upon mixing, and is the fractional

conversion (degree of hydration). fhp is the volume fraction occupied by the hydration

products CH and C-S-H (see Figure 8-8). For more details on the definition of each

term, see Appendix A.

The volume fraction of the porous C-S-H can then be calculated as follows

fgei = fCSH (8.19)


qgel

where 'lgeI is the packing density of the C-S-H gel. It is one of the most critical

parameters in this study. fsH is the volume fraction occupied by the C-S-H solid

(no adsorbed water on the surface).

On the other hand, the density of the C-S-H gel and packing density thereof

is not an intrinsic property of the C-S-H gel and it varies spatially and temporally

depending mainly on the water-to-cement ratio, curing temperature, and particle size

distribution of the cement powder.

The derivation of the evolution of volume fractions during the hydration of Class-

G cement are discussed in the Class-G chapter below and relevant models in the

literature are also reviewed. The reviewed models include measurements other than

the enthalpy and volume of the reaction, such as ThermoGravimetric analysis (TGA)

[157], Quantitative X-Ray Diffraction (QXRD) [90], Nuclear Magnetic Resonance


(NMR) [133], QuasiElastic Neutron Scattering (QENS) [73], and Scanning Electron
Microscope (SEM) [71]. Other measurements can also be used in a similar fashion

without the resort to approximate chemical reactions.

197
Table 8.3: Variation of C-S-H density and molar volume

Formula p " Remarks Refs


(g/cm 3 ) (g/mol) (cm 3 /mol)
C1 .7 SH 4 (*) 1.85-1.90 227.5 121.3 [212]
1.90-2.10 227.5 113.7 [157]
2.12 229 108 [15]
C 1 .7SH 1 .2 2.86 177.0 61.9 D-dry [212]
C1. 7 SH(1. 3 - 1 .5 ) 2.44-2.50 180.6 73.1 D-dry [67,157]
2.70-2.86 180.6 65.0 [32,99,163]
C 1.7SH 1.8 2.604 187.8 72.1 [2]
C1.7SH2.1 2.18 193.2 88.6 11% RH [184, 212]
2.43-2.45 193.2 79.2 11% RH [67]
C 1 .6 7 SH 2 . 1 (t) 2.46 191.6 78 Jennite [114]
C 1 .7 5 SH 7 .5 1.73 293.3 170.0 Saturated/inner [55]
C1.75SH4.2 1.91 233.9 122.5 Saturated/outer [55]
C 1 .7 5 SH 5 .4 1.86 255.5 137.4 Saturated/inner [55]
C 1 .75SH 5 .3 1.96 253.7 129.4 Saturated/outer [55]
C 1 .7 5 SH 5 .2 1.83 251.9 137.8 Saturated [55]
* tandarrd enthailn of formntion i: AH = -1283 kJT/mol from reference r2121
.

t standard enthalpy of formation AHY - -2723 kJ/mol from reference [114].


Both enthalpies are equivalent if the difference in water content of the C-S-H gel has the standard
enthalpy of formation of water.

198
Average density of C 3 S paste

The total density of the C3 S paste during hydration can be calculated from the volume

fractions as:

Ppaste = (p) = fipi = fC3 spC3 s + fCsHPCSH + fCHPCH + totPw (8.20)

where fi and pi are the volume fraction and the density occupied by phase i, respec-
tively. The density of the paste can also be calculated from the initial density and
the fractional volume of the reaction as:

Ppaste = (P) = Po + frxnPrxn (8.21)

where po = (1 - no)pC3 s + nopw is the initial density of the mix and Prxn = Pw = 1.0
g/mL is the density of the fluid used to keep the sample saturated (water). The density
of the tested samples was measured at the end of each test and the w/c is corrected
to account for the water squeezed out of the sample during setting up the test and
the volume changes at early age before setting and upon complete consumption of

the bleed water. The corrected w/c are shown in Table 3.3.

Mass fraction of C-S-H

The mass fraction of the C-S-H solid is calculated here as a necessary quantity for

surface area calculations. The mass fraction can be calculated at a given degree of

hydration and water-to-solid mass ratio as follows:

PCSHfCSH __ PCsHfCSH
p) (1 - no)pcas + nopw + frxnPw

where (p) is the average density of the C3 S paste.

199
I

0.9 0.9

f. fH 0.8

0.7 0.7

0. 4 0. 4 --------------------- - ---- -

-
fCSH
0.303 ' ICS
0.2 '0.2
fc,3s kas
0.1 0.1

01 10 0.2 0.4 0.6 0.8


0 0.2 0.4 0.6 0.8
C

(a) w/c=0.34 (b) 'v/c 0.49

Figure 8-8: Evolution of volume fraction of the various components during the by-
dration of C 3 S with water-to-solid ratio of (a) iv/c = 0.34 and (b) w/c = 0.49. The
yellow spot marks the degree of hydration where all the remaining free water is inside
the C-S-H gel. f,,. is the volume fraction occupied by the free water remaining from
the original mixing water only (not including any water provided from the outside
during the hydration). fc,, is the volume fraction occupied by free water remaining
from the initial mixing water in the capillary porosity. fTsi is the volume fraction
of the solid C-S-H only. Pgl is tie volume fraction occupied by the gel porosity. The
capillary porosity d/c-, is represented by the areas shaded in blue and grey. The grey
- The area represents part of the water uptake from outside the sample .
total porosity of the system is 010t = cap +-,,I. The volume fraction occupied by
the C-S-H solids and the gel porosity is defined as the C-S-H gel f+l fCS - OgPel-

200
8.3 Eigenstress development during C 3S hydration
through a multiscale thought-model

A multiscale thought-model is developed here to interpret the bulk volume changes


and to calculate the eigenstress development during the hydration of C3 5 paste.
Bulk volume changes of C3 S are investigated for three different particle size distri-
butions, different water-to-solid ratios, and curing temperatures as shown in Figures
6-8 to 6-14. The water-to-solid ratio is chosen to obtain reasonable workability and
prevent excessive sedimentation for each particle size distribution.
In general, the bulk volume changes of C3 5 pastes include an expansion part during
the first 200 hours of hydration of saturated samples followed by shrinkage for the
remainder of the tests. The initiation of the bulk volume change measurements shown
in Figure 6-14 for different PSDs and in Figure 6-22b for variable curing temperatures
is related to the evolution of pressure at the bottom surface of the sample. Initially,
pressure at the bottom surface is identical to the pressure delivered through the top
port and the flow rate through the bottom port is completely restricted due to the
presence of the check valve in the flow network of the RDC test. Once the slurry is set
and the bleed water is completely consumed by the sample, pressure at the bottom
port starts dropping at a rate proportional to the rate of hydration and a maximum
drop equal to the cracking pressure of the check valve, as shown in Figure 8-9.
The measured bulk volume changes during the hydration of C 3 S are attributed
solely to the development of eigenstresses assuming that the 60kPa effective stress
does not contribute to the measured bulk volume changes in the set paste. Eigen-
stresses develop in cement-based materials due to a combination of possible driving
forces as discussed in Section 2.2. A multiscale microporomechanics is developed to
interpret the measured bulk volume changes of saturated hydrating C3 S paste. The
multiscale microporomechanics model consists of a three-level model of the cement
microstructure. In this model, CBM will be modeled as a matrix of C-S-H particles
in which all other solid phases and the pore space are embedded.
As shown schematically in Figure 8-10, the first level (level I) of the model includes

201
Pc (C3S-M-RD-W42-T25-P1)
- Pc (C3S-G-RD-W34-T25-P1)
-Pc (C3S-F-RD-W49-T25-P1)
1060 -

-
1040 - --

-
1020

-
1000

980

0 960 -~ ~rn
940
0 5 10 15 20
Time (hr)

Figure 8-9: Evolution of pressure at the bottom surface of the sample during the RDC
test with 60 kPa check valve. Pressure at the bottom port remains constant until the
slurry is set and the bleed water is consuned by the volume of the reaction. Pressure
at the bottom port starts dropping at a rate proportional to the rate of hydration
and a maximum drop equal to the cracking pressure of the check valve.

the C-S-H colloids and the interstitial space also known as gel porosity. The second

level (level II) includes level I in addition to the capillary porosity. Finally. residual

(un-reacted) cement grains and other non-reactive solid inclusions embedded inside

the porous matrix define the continuum scale (level III). Driving forces of bulk volume

changes are modeled as eigenstresses at the corresponding scales. The goal of the

nicromechanical modeling is to establish the relationship between the macroscopic

observations and their finer-scale origin by considering the intrinsic properties of the

material constituents and the driviNig forces of bulk volume changes.

8.3.1 LEVEL I (The C-S-H gel)

Following the framework developed in Section 5.1. the first level (level I in Figure

8-1() of this model includes the C-S-H colloids and the interstitial space also known
as gel porosity. Here, we consider a mli(roporomnechallics model of the C-S-H phmse.

which is conposed of the C-S-H colloids and the gel porosity. in additiol to the

202
OU,

o OC

C(CH
Level (I) Level (II) Level (III)
Figure 8-10: Three-Level thought model of hydrating C3 S paste. The first, level
(level I) of this model includes the C-S-H colloids and the interstitial space also
known as gel porosity. The second level (level II) includes level I in addition to the
capillary porosity. Finally. residual anhydrous cement grains and other non-reactive
solid inclusions are embedded inside the porous matrix define the continuum scale
(level III).

eigenstress due to the densification process. The volume fraction occupied by the

components at the first level is:

f = fgei = fCsH (8.23)


1/gel

where i1gel is the packing density of the C-S-H gel and fJJISH is the volume fraction

occupied by the C-S-H solids as defined by Eq. (9.7).

The level I state equations define changes in stresses and gelporosity at constant

hydration degree., and read as:

Incremental mean stress in the C-S-H gel (Level I)

(1 = do(z)) = KIdE' + (I - b') do-* - b' dp (8.24)

where p(; is the pressure prevailing in the gel porosity. The gel porosity change is

defined at constant hydration degree.

Applying the self-consistent scheme, the poroelastic properties at this scale read

203
as [193-195]:
K' 1 - bV
(8.25)
ks 4GI/gs + 3 (1 - r') rs
where rs = k/g8 = 2 (1 + v) /3 (1 - 2v') and:

- = - (1 - 1r9 (2 + 171) (8.26)

2
+ V"144 (1 - rs) - 480,q' + 400 (r7) + 408rsrI - 120r8 (71)2 + 9 (rs) 2 (2 + ql)2
16

with rTI the packing density of the C-S-H gel:

n= (/a)/ (8.27)

with 3 = In /In (7rn),and /a oo m= ./r ( o is the hydration degree


threshold at which the solid particles percolate r1o = 1/2). The packing density is
derived by fitting meso-scale simulations by Masoero et al. [121] as discussed in the
subsequent sections.

This densification is captured by a power relation of the form [195]:

(8.28)

where qlim is the C-S-H limit packing density (at complete hydration jim), which
depends on poly-dispersity of the C-S-H particles [121], (typically, qlim = 0.64 for Low-
density C-S-H; ulim = 0.74 for High-density C-S-H, [47,200]); While a is determined
from the percolation threshold:

In
a-= tm (8.29)
ln (-)-
lim

with /o = 0.5 the packing density beyond which the hydrating matter can support a
deviatoric loading.

204
8.3.2 LEVEL II (C-S-H gel with capillary porosity)

At level II, we consider a second pore space, the capillary porosity, in which another

pressure prevails, denoted by p and the C-S-H gel (C-S-H solid and the gel poros-

ity) makes up the solid phase. The volume occupied by the hydration products is

designated Vel and the space occupied by the capillary porosity is Vcap. The volume

fraction occupied by the second level is:

f= - fI + qcap = fgei + Ocap (8.30)

The porosity of the second level is then defined as:

II = ~ap (8.31)
-f 'I

The incremental stress state equation for the stress distribution at level II reads:

dum (z) = k(z)dc(z) + duP (z) (8.32)

with:
KI Vz (2 Vge
k(z)= (8.33)
0 Vz E Vcap

d PI = (1 b') do-* - b' dpG Vz G Vgle


d-P (z) = 8.34)
-dp Vz Vap

The eigenstress at Level II is obtained by the application of Levine's Theorem:

dEP1 = {A,(z)doP(z)) = (1 - W) dZPI - bIdp 8.35)


(

Whence, the incremental equation of state:

d E' = (do,,, (z)) = K'dEII + (1 - bI - bW) do-* - bI dpG - b,'dp (8.36)

205
wiere:

K" = (1- 2) K' (8.37)


K"
1)/i' - 7"Ai41 (8.38)

-
K'
K'\
i j'= = -bj) (8.39)

K1I
(1 - b/) (I - b') (8.40)

For identical pressure in the capillary porosity and the gel porosity dp; = dp. we

have:
Incremental mean stress at the second level (Level II)

d(EI = Kdo1 (z)) = KId Ef + (1 - b") du* - b" dp (8.41)


dZP'

where:

b"l = bbl=1b.?)bI +bW =I- (8.42)

The elastic properties at this level are calculated with a M\Jori-Talnaka scheme [193

195]:

K" 4 (1 - (8.43)
-

KI 3 , IIrhyd + 4

Gil (1 -") (8 + 9.h4d) (8.44)


GI 6H" (2 + rhlY() + 8 + 91r.Yd

where rhyd = K 1 /G.

8.3.3 LEVEL III. Reinforcement by Rigid (but Slippery) In-


clusions

The third level includes the reinforcing effect of the residual anhydrous cement grains

as a two-phase composite. One phase is the porous matrix formed of the hydration

206

Iw
products in addition to the gel and capillary porosity (Vm = Vgei + Vcap). The sec-
ond phase is the residual anhydrous cement grains and non-reactive solid additives

and portlandite. A convenient way to model the second phase is by considering the

anhydrous cement grains and the solid additives as rigid inclusions un-bonded to the

porous matrix (slippery inclusions). The assumption of discontinuous slippery inter-

face simplifies the strain localization tensor but does not affect the volume integral.

Inclusions like the residual cement grains can also undergo eigenstrains due to the

progression of the hydration reaction. Such contribution to the overall behavior of the

hydrating material will be included in this level of the multiscale microporomechanics

model.

Stress and Strain Localization

While the inclusions are rigid, it's worthwhile considering the strain localization in

the matrix. To do this, we consider an eigenstress-free situation according to which

d EII = KII'd EI

where K"'I is the homogenized stiffness tensor. Then consider:

dZEII = (dom(z)) = (1 - fi ) (K dEvI) + fcdo-m

where fc, is the volume fraction occupied by the residual anhydrous cement, and do-
is the mean stress in the inclusion:

fi douIn = K"'!dEI"1 - (1 - fci) (KIldE

Assuming that dorjc = KintdEn (with Kin' an interface stiffness, and dEv"nc
1M

f du dS ). Then,

feidE In V = KI"'
it dEI"'
v
- (1 - fci) Kint
KI" dE"
V

207
or, with strain localization factors (dEvflC = AincdEv!II, and dEf'' = AmdEtIII), while

considering strain compatibility, fcjAinc + (1 - fc) Am = 1:

-nc K"'I - KI
fiAi
Kint (1 - KII/Kint)

Thus, the inclusion stress:

Incw Kint inc 1l 1- KI'I - K "I


(lm K fci1 - KII/Kint d =1 1 - KII/KInt
"M
fel I - KIII/ int

and the total stress:

dY:Il
= ill 1 + KIIIK int EI
d !!=K" (I - KIIII/Kint d E"
m

For Kint -+ oc, and hence Ain, -+ 0, we naturally recover dEI' = KI'IdE7"', and
the stress localization:

KI
do Ic = BdE4I ; B, = 1 (8.45)
KIII

Finally, an application of a self-consistent model for rigid inclusions with slippery

boundary conditions gives:

K"' 3r' + 4 (G"'IG) fc


KII 3rm(1 - f) ( .'
)

with rin = GI"/K", and:

G"' 1 1
Gil 24 (2 - 3fjc)
(8 (3 - 2.fc) - (15 - 24fe) r-im (8.47)

+ 9(8fc - 5)2 (r7)2 + 48 (11f2 - 29fc + 15) (rm) + 64 (3 - 2fc)2

The model is restricted to f, <


.

208
Poroelastic State Equations

It is now appropriate to write the strain distribution in the system at level III in the
form:
dom(Z)
dc(z) = + deT(z) (8.48)
k(z)
with:dEZI - dEP" - KI'dEv

K" in V
k(z)= (8.49)
-+oo in Vine

dKP"I dp
deT(z) = KII =- (1- b') K" + bl'
2 K"I
in V
(8.50)
dEinc in Vine

Homogenization, using Levine's theorem yields:

dVI
dEvII= K + dET (8.51)

where macroscopic eigenstrain reads:

dE[ = (B(z)dE '(z)) = -K"' = -(1- foB) + fci BdE Te (8.52)


K"I
dEP"
(KIll - KI dcT
dine (8.53)
=KIII + KIII

This leads to the following incremental stress equation of state:

d EQI' = (dom(z)) = K"IdEv" +(I - b") doJ* - b' dpG - bl'dp - (K" - K") d4E n
dEPII inclusion

(8.54)

Last, when the two pressures are equal (dpG = dp), the previous equations simplify
as follows:

209
Incremental mean stress at the continuum level (Level III)

Kdu. &,,(z)) K" 'dEf"1 + (1 - b) de* - b dp1 + (K 11' - K"I) dET~

dEP' in clusion

(8.55)

The bulk volume changes of C 3 S have been measured for three different particle
size distributions. different water-to-solid ratios. and curing temperatures as shown in

Figures 6-8 to 6-14, can now be viewed through the multiscale microporomnechanics
nodel.

8.3.4 Eigenstress development

Based on the model microstructure shown in Figure 8-10, the elastic properties are

calculated at level I from Eq. (8.25) and (8.26). at level II from Eq. (8.43) and (8.44).
and at level III from Eq. (8.46) and (8.47). The resulting elastic properties are shown
in Figures 8-11, 8-12. and 8-13. Eq. (8.55) is then used in conjunction with the elastic

properties to estimate the evolution of eigenstresses in the hydrating material.


The resulting eigenstresses are shown as function of degree of hydration , packing

density of the C-S-H gel igel solid volume fraction SVF, and the inverse of the rate of

hydration 1/d in Figures 8-14 to Figure 8-17. It can]be seen that the onset of switching

from negative eigenstresses (expansion-inducing du* < 0) to positive eigenstresses

(shrinkage-inducing du* > 0) is related to the solid volume fraction and the rate of

hydration, which is expected to result from an out-of-equilibrium process. The solid

volume fraction is calculated based on the vohune fractions of the residual anhvdrous

C 3 S. CH and the C-S-H gel with constant packing density of 0.64:

SVF =tfoobI =fe.s + fo 0. +1


( (8.56)

Analogous to the expansion during the hydration of MgO. expansion of C1;S paste

210
at early ages is attributed to the difference in the average density of the reactive solids

and the hydration products. During the hydration of MgO, only Mg(OH) 2 is formed

and the contribution to the measured expansion was attributed mainly to the reaction

mechanism and the microstructure of the material. During the hydration of C 3 5 on

the other hand, it is assumed that CH precipitates only in the water-filled pore space

without significant contribution to the bulk volume changes. The main hydration

product of C 3 S hydration (C-S-H) is considered to be at the origin of both expansion

and shrinkage of C3 S paste. Similar to the precipitation of Mg(OH) 2 , part of the C-

S-H gel precipitates in the water-filled pore space without significant contribution to

the bulk volume changes. The other part precipitates within the original boundaries

of the C3 5 grains, possibly through a topochemical reaction, and plays the major role

in the measured expansion.

The effect of PSD and w/s ratio is shown in Figure 8-14 and Figure 8-15. The
rate of expansion and shrinkage seems to be independent of the PSD while the switch

from expansion to shrinkage is related to the development of solid volume fraction.

The effect of curing temperature on bulk volume changes is shown in Figure 8-16 and

Figure 8-17.

Effect of temperature on the hydration reaction includes the acceleration of the

reaction and changes in the nature of the hydration products. Increasing the curing

temperature seems to reduce the magnitude of the eigenstress developed during hy-

dration. The effect of the curing temperature on the rate of hydration is insignificant

after 30 hours of hydration (see Figure 6-12) and consequently, the variation of bulk

volume changes cannot be attributed to the change in the rate of hydration. The sec-

ond possible effect of curing temperature on the bulk volume change is the changes

in the nature of the hydration products. Increasing the curing temperature seems to

favor the formation of more stable (closer to equilibrium) structures of C-S-H which

is less susceptible to volume changes due to densification.

211
0.7 0.5
-

0.6 -Level I -Level I


I 0.4 -Level I1
- Level I III III
0.5 -Level -Level

0.4 0.3

0. 3 0.2
0.2
031.4 06 0.1
0.1 .

(a) (b)

Figure 8-11: Evolution of the shear modulus G/g" during the hydration of C3 S with
(a) w/c = 0.34 and (b) w/c = 0.49 as calculated by the multiscale microporomnechan-
ics model for the three levels in function of hydration degree.

0.7 0.4 - - _-__-_

0.6 - L vl] 0.35


-- Level 11
0.5 -L vlI 0.3 -Level III
0.25
0.4

().3~
0.151
0.2

0. 1

00 () - 0.- 04 0.6 0.8 1


0 0.2 0.4 0.6 0.8

Figure 8-12: Evolution of the bulk modulus K/k during the hydration of C:3 S with (a)
U/c = 0.34 and (b) w/c = 0.49 as calculated by the multiscale microporomechanics
model for the three levels in function of hydration degree.

212
0.95 0.95"
0.9 0.9!
0.85 0.851
0.8 0.8.
0.75 C 0.75
0.7r 0.71
-Level I -- Level I
0.65 -Level II 0.65 -Level I
--- Level III - Level III
0 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a) (b)

Figure 8-13: Evolution of the Biot coefficient b during the hydration of C 3 S with (a)
I/c = 0.34 and (b) w/c = 0.49 as calculated by the multiscale microporomechanics
model for the three levels in function of hydration degree.

-- *(C3-F-RD-W49-T25) --- a* (C3S-F-RD-W49-T25)


a*(C3S-M-RD-W42-T25) u* (C3S-M-RD-W42-T25)
-* (C3S-G-RD-W34-T25) -* (C3S-G-RD-W34-T25)
1
lgei
0.25 0. 45 0.65 0.85 0.55 0.6 0. 65 0.7
0- 0-
-
-

-1 01 - -1 0
-

0
-2 0 - _
1 -- -2 0

0
-3 0 -2 0 - _- -
-

-40 -40 T-
-

-50 -50
-

-60 -60
-
-

(a) (b)

Figure 8-14: Development of eigenstresses as calculated from the multiscale mi-


croporomechanics model for the three different, particle size distributions (F=fine,
M=miediuin, G=coarse). (a) as a function of the degree of hydration (b) as a func-
tion of the packing density of the C-S-H gel. where the packing density is calculated
based on Eq. (8.27).

213

'
* (C3S-F-RD-W49-T25) - 0* (C3S-F-RD-W49-T25)
0* (C3S-M-RD-W42-T25) 0*(C3S-M-RD-W42-T25)
-~ * (C3S-G-RD-W34-T25) - * (C3S-G-RD-W34-T25)

SVF 1/(di;/dt)
0.65 0.7 0.75 0.8 2000 4000 6000 8000
0.55 0.6
1 0
0

-10 -10
-

-20 -20
-

0
-30 -30
-

-40 -40
-

-50 -50
-

-60 -60

-
-

(a) (b)

Figure 8-15: Development of eigenstresses as calculated from the multiscale mi-


croporonieclianics niodel for the three different particle size distributions (F=fine.
\l=mediumn. G=coarse): (a) as a function of the solid volume fraction, where the
solid volume fraction is defined in Eq. (8.56): (b) as a function of the inverse of the
rate of hydration as measured by the volume of the reaction.

- 0* (C3S-M-RD-W42-T45) *- * (C3S-M-RD-W42-T45)
-- * (C3S-M-RD-W42-T35) -- ~
*(C3S-M-RD-W42-T35)
0* (C3S-M-RD-W42-T25)
0* (C3S-M-RD-W42-T25)

0.3 0.4 0.5 0.6 0.7 0.8 0. 6 0.62 0.64 0.66 0.68 0.7
0 0

-5 -5
-10 - ~ I-- - _
-10
-15 -15
-4-
-20 -20

-3
0 -25
-25
0
-30 -30
-

______ 'I -35


-35
-40
7 'I
-40

-45 -45

(a) (b)

Figure 8-16: Development of eigenstresses as calculated from the multiscale mi-


0 .
croporonmechanics model for three curing temperatures (T25=25 C, T35=35C
T45=450 C): (a) as a function of the degree of lydration: (b) as a function of the
calculated based oil
p)acking density of the C-S-H gel. wherc the packing density is
Eq. 8.27.

214

-I
-- - (C3S-M-RD-W42-T45) -- y* (C3S-M-RD-W42-T45)
(C3S-M-RD-W42-T35) - (C3S-M-RD-W42-T35)
T* (C3S-M-RD-W42-T25) a* (C3S-M-RD-W42-T25)
SVF 1/(d /dt)
0.6 0.65 0.7 0.75 0 .8 0 3000 6000 9000
n 0

-
-5 -5 -

-
-10 -10 - -

-
-15 -15
-20 -20
-25 -_______

-30 -30

-35 -- -35

-
--J
-40 - -40

-
-45 -_-_- -_-_ -45

-
(a) (b)

Figure 8-17: Development of eigenstre sses as calculated from the multiscale iii-
croporomechanics model for three cm ing temperatures (T25=25 0 C. T35=350 C.
T45=450 C). (a) as a function of the so lid volume fraction (1)) as a function of the
inverse of the rate of hydration as measi ired by the volume of the reaction.

8.4 Speculation about the origin of Eigenstresses

In the previous section, eigenstresses air assumed to originate at the scale of the C-

S-H grains and it is then up-scaled using the multiscale thought-model. This section

represents an attempt at relating the origin of such eigenstresses to known physical

phenomena.

A profound understanding of each of the driving forces of bulk volume changes is

required to separate the contribution of e ach of these forces and to develop appropriate

mathematical models and to design a r elevant experiment. The main focus of this

study is to investigate bulk volume ch anges during the hydration of CBM under

saturated conditions and this already ex cludes capillary pressure, disjoining pressnre,

and surface energy as possible driving fc rces from this investigation.

Another possible driving force of the bulk volume changes in addition to the forces

discussed in Section 2.2 is the densificati on and hardening of the hydration products.

This kind of force has never been expl( )red before but the recently published work

of Uhn et al [26] show that colloidal C- S-H particles precipitating inside the porous

215
C-S-H gel can lead to volume changes. In his analysis, Ulm suggested that the C-S-H

gel is initially formed with low packing density and that this density increases due

to the progression of the hydration reaction to reach a maximum of 0.72 depending

on the polydispersity of the system. Such assumption of densification can also be

found in other works related to the hydration kinetics models (see for example [23])

without reference to bulk volume changes. The physical origin of densification in

Ulm's model is based on meso-scale simulations and is resulting from insertion of

C-S-H colloidal particles inside the loosely packed C-S-H gel resulting in reduction

of the average separation distance between the colloids surfaces [121]. The reduction

in the separation distance in conjunction with the appropriate interaction potential

gives rise to eigenstresses in the system leading eventually to the observed shrinkage

of the material.

To understand the possible contribution of the densification of C-S-H to the bulk

volume changes, we start with a brief review of the structure and properties of the C-S-

H gel. Hydration of cement produces several hydration products among which, C-S-H

is the most abundant. C-S-H is the glue that binds all the other phases in cement paste

together and it is also the phase that predominantly controls the properties of the final

hardened product. Investigations of the structure and composition of C-S-H indicate

a granular nature with a wide range of pore sizes and chemical compositions. The

variation of the composition and structure of C-S-H is dependent upon the chemical

composition of the cement and pore solution as well as the mineral additives and

the curing conditions. C-S-H is generally viewed as an aggregate of nanoscale solid

particles packed randomly in porous structures with gel pores. This structure of the C-

S-H gel is mainly deduced from sorption isotherms, elastic properties, statistical nano-

indentations, thermal analysis, SANS [47,57,70,99,159]. A schematic illustrating the

representative structure of the C-S-H gel is shown in Figure 8-18 after [2].

8.4.1 Powers and Brownyard Model

A review of cement-based material models naturally starts with the pioneering work

of Powers and Brownyard [157,159,160]. In their model of hardened cement paste,

216
- Calcium silicate sheets
with OH- groups

mnerlayer space with


physically bound H 2 0

Adsorbed HO

Liquid H20 in
nanopores

5 nrn

Figure 8-18: Schematic diagrain of the nanoscale C-S -H particles from [2]. The black
lines indicate the boundaries of the solid C--S-H particles, with solid C-S-H in the
form of nanometre-scale particles with a structure based on infinite sheets separated
by layers of physically bound water.

Powers and Brownvard described the colloidal nature of cement paste based on water

sorption isotherms and thermal analysis where they assume that the volume of the

material is made up of unhydrated cement grains, cement gel, and capillary porosity

as shown in Figure 8-20. One of the main distinguishing features of this model is the

classification of water into three categories: the non-evaporable water wi,. gel water

wg, and capillary water wv. The non-evaporable water is arLbtrarjiydefined as the part

of "water" that has a vapor pressure of more than 6 x 10-' nnHg at 23 0 C. Although

it is referred to as non-evaporable "water", for the main part this "water" is present

as part of the solid phase in the paste mainly as OH groups and structural water.

Part of the non-evaporable water ealn also be chemically or physically combined with

the hydrous silicate and aluniinate phases. w, is quantified by loss on ignition of the
cement paste that has been dried to the previously specified pressure. All remaining

water that has vapor pressure above 6 x 10-' mmnHg is referred to as the evaporable

water.

Gel water. on the other hand, is defined as the water encapsulated inside the ce-

ment gel that includes all the hydration products (colloidal and non-colloidal). The

vapor pressure of the gel water ranges from 6 x 10-1 nunHg up to the vapor pressure
of the bulk water depending on the (legree of saturation of the gel. The weight of

the gel water in saturated mnaterial is approximately 41L, where L7, is the quantity

217
of adsorbate required to form a monomolecular layer on the surface of the solid. An-
other important characteristic of the gel water is a density of about 1.1 g/mL due to
confinement and its close proximity to solid surfaces. It is critical here to note that
"cement gel" refers to all the hydration products and it does not distinguish between
calcium silicate hydrates (C-S-H), portlandite (CH), calcium monosulfate alumino-
ferrite (AFm), Calcium trisulfate aluminoferrites (AFt) and all possible hydration
products.

The third type of water is the capillary water, which is defined as the pore solution
that occupies the remaining space in the material other than the solid phase and the
gel pores. Capillary water has the properties of bulk pore solution and it does not
exist at pressures below 0. 4 5 pS, where p, is the saturation pressure. The distinction
between the gel water and capillary water is arbitrary and the main distinction relies
on the fact that there is a fixed ratio between the gel water and the amount of gel in
the material.

These three classes of water can then be used to calculate the volume fractions of
the various components in the cement-base materials. The total volume of the cement
paste in Powers model is composed of unreacted cement and components derived from
the evaporable and non-evaporable water as related to the properties of the hydration
products

Vtot = CVc + WnVn + (wg + wC)Vd (8.57)


Vs VP
where c is the cement content in the paste in grams, w, is the weight of non-evaporable
water in grams, w, is the weight of capillary water in grams, w9 is the weight of gel
water in grams, V is the specific volume of cement in cm 3/g, Vd is the specific volume
of capillary and gel water - lcm 3 /g, and V is the specific volume of non-evaporable
water in cm 3 /g and here it is assumed to have constant empirical density of 1.22
g/mL regardless of its origin. wn, wg, and w, are measured for cement paste with
c grams of originally unreacted cement. In the absence of bulk volume changes, the
original volume is also the sum of the volume of the mixing water and the original
unreacted cement.

218
MWEEMMU-1- mm.- - -

-
Powers and Brownyard showed that the quantity of water adsorbed at relative

pressure below 0.45 is the same for all samples regardless of the water-to-solid ratio
and referred to this water as the gel water (see Figure8-19). They also attributed the

increase iin the adsorption above 0.45 relative pressure to capillary condensation ill

the capillary pores.

* Ref 9 -I4. /4 - 7
.50 7by.C
by wt.
6, W 0.439
S roble 29 0 Ref 9 -14
I I 7 Cement /5365
Cu'ed Wet /80 ayj
d
Age- 7 da.

-
- -/ ?8 doy>
4
"A/vm

180 deGY
3 --- - -
.

4 - - ------ --

2 -
-

) .1 i -3 .7 5 .9 .b I
PA
Fig. 3-8 (above)--Effect of original w/c
on w/V. curves
Fig. 3-9 (right)-Effect of wet-curing 0 .1 .2 .3 .4 .5 6 .7 b .9 0
on w/V curves p

Figure 8-19: Typical w/l to show the effect of w/c (left) and the effect of age (right)
on the adsorption behavior of cement-based materials, from [157].

8.4.2 Densification of C-S-H

One way to view the densification process of C-S-H gel is possible )v considering the

reaction mechanisms and the inner- and outer-C-S-H porducts. C-S-H as formned by

the hydration of Alite is typically classified into two types, the outer product C-S-H

(Op-CSH) and the inner product C-S-H (Ip-CSH) [179]. Inner product C-S-H (Ip-

CSH) is the C-S-H formed inside the original boundaries of the cement grains (alite

and belite particles). Ip-CSH is generally a high density C-S-H (HD-CSH); but Ip-

CSH from small cemnent grains is usually LD-CSH [164]. It is highly likely that the Ip-

CSH is formed through a topochemical reaction in a diffusion-inited reaction [179].

Outer product C-S-H (Op-CSH) is the C-S-H formed outside the boun(arv of the

219

I
A 5
Fig. 3-10

Figure 8-20: Schematic representing the model of cement paste iicrostructure as


depicted by Powers and Brownyard model from Ref. [157]. The shaded area represents
the porous cement gel with inclusions of non-colloidal (microcrystalline) material. (A)
is the cement paste with high water-to-solid ratio and the capillary lenses (capillary
bridges) represent the capillary water in the model. (B) is the model for the structure
of cement paste with low water-solid ratio where all the capillary pores are filled with
porous hydration products (cement gel).

4c

Fig. 1. Simpliied model $ pte structure. Gel particles are repre-


sented as needles or plates; C designates capillary cavities. Ca(OH2
crystals, unhydrated cement, and minor hydrates are not represented.

Figure 8-21: Schematic represe11tinlg the niodel of cement paste microstructure as


(lepicte(d iby Powers from Ref. [159].

I
220

--
original cement grains: and it usually has a low density and incorporates all kinds of
ions available in the solution in its structure.

Another classification of C-S-H call be found in the work of Tennis and Jen-
nings [183] and Constantinides and Ulin [46] in which C-S-H is classified based on

its packing density. High Density C-S-H (HD-CSH) is characterized by high density

and constitutes the major part of the Ip-CSH of the large cement particles and its

porosity is not accessible by Nitrogen. Low Density C-S-H (LD-CSH) is characterized


by the low density and constitutes most of the Op-CSH. Most of the porosity of the

LD-CSH is accessible by nitrogen [183].

Densification of C-S-H gel was also studied experimentally for different drying

conditions [163] and for saturated cement paste [133]. The desification of C-S-H in

the work of Muller et al. [133] seems to be dependent on the initial porosity (water-

to-solid ratio) with the ultimate packing density decreasing as the water-to-solid ratio

increases. Mikhail et al. [128] investigated the pore structure of fully hydrated cement

paste with w/c ranging from 0.35 to 0.70 and found that aln increase in w/c will result

in an increase in the average size of the gel pores and consequently, a lower density

C-S-H gel (see Figure 8-22).

120

140

120

0 0 0 0 0 0.2 0.4 0.6 0.8 1.0


S

Figure 8-22: Nitrogen sorption isotlierms on hardened portland cement paste. after
[128].

Whether thc densification of the C-S-H gel is a continuous process or simnply due

221
to the formation of inner and outer products, it seems that the initial porosity has a

significant impact on the densification. The fact that the densification of the C-S-H

gel is dependent upon the initial water-to-solid ratio indicates that the impingement

of the hydration products plays an important role in the process. The rate of the

hydration reaction during the early stage of hydration is driven by nucleation and

growth and nearly independent of the water-to-cement ratio [19]. Once the reaction

becomes diffusion-limited, the rate of hydration becomes strongly dependent on the

water-to-cement ratio [16]. The influence of the w/c is typically attributed to the

available space for the hydration products to form by linearly relating the rate of

hydration to both the volume fraction of water-filled porosity and the volume fraction

of residual anhydrous cement [16]. A closer look at the evolution of porosity (Figure

8-22) indicates that the impact of w/c on the reaction rate starts long before the w/c

becomes a determining parameter for the diffusion-limited reaction. That is, for a

large range of water-to-cement ratios, the rate of the reaction is affected even though

there is still enough empty space for the hydration products to form. This leads to the

conclusion that the rate determining parameter is not the availability of water-filled

pore space but rather the mean diffusion distance traversed by the both the solute

and the solvent.

The concept of the effect of the water-to-cement ratio on the mean diffusion dis-

tance (MDD) is illustrated in Figure (8-23). At higher w/c, the MDD remains low

due the high initial separation distance between the cement grains. At lower w/c

ratios, the impingement of hydration products occurs at earlier age and at a higher

rate leading to a drastic increase in the MDD. For higher water-to-cement ratios, the

mean separation distance between the cement grains is larger and the impingement

of the hydration products occurs at later ages and at a slower rate. Impingement

of the growing particles increases the MDD drastically depending on the initial w/c

ratio. The increase in the MDD leads to an increase in the degree of super saturation

with respect to the hydration products through the hydration shell which leads to

the precipitation and densification of the C-S-H gel. Berliner et al. [19] found an

exponential relationship between the effective diffusion coefficient and the w/c ratio

222
independent of the particle size distribution.
r Supersaturated
Bulk solution Bulk solution

Neighbor High
particle Shell supersaturation

Core
+ Concent ration

Core/Shell
Interface *-P4 Solute
*-- Solvent
Outer
-------- surface

(a)

-Complete acces sible Reduced accessible


sesurface area
- surface area

VNew outer
hydration products
New inner
hydration products

(b) (c)
Figure 8-23: A cartoon illustrating the impact of w/c on the mean diffusion distance
(MDD). For higher water-to-cement ratios, the mean separation distance between the
cement grains is larger and the impingement of the hydration products occurs at later
ages and at a slower rate. Impingement of the growing particles increases the MDD
drastically depending on the initial w/c ratio. Fig. (a) shows the effect of thickness
of the hydration shell on the concentration gradient of the solute ions of a partially
hydrated cement grain. Fig. (b) the MDD is exactly the thickness of the hydration
shell. Fig. (c) The MDD increased drastically due to the impingement of the growing
hydration products. The increase in the MDD leads to an increase in the degree of
super saturation with respect to the hydration products through the hydration shell
which leads to the precipitation and densification of the C-S-H gel.

8.4.3 Meso-scale simulations

Advances in miolecular modeling of cement-based materials provide a powerful and

versatile mieans to exploring the structure and properties of C-S-H [96.121,122,148].

223
Among all the possible driving forces of bulk volume changes in cement-based mate-
rials, we discuss here the possible contribution of stresses resulting from the precipi-
tation of the C-S-H gel. Precipitation and the subsequent densification of the C-S-H
gel seems to induce significant eigenstresses in the material as shown by the simula-
tions of Masoero et al. [121], Ioannidou et al. [96] , and micromechanical modelling
by Ulm et al. [195]. Stresses can be estimated in a simulation box using the known
expressions for the virial stress (at zero kelvins for simplicity) [135]

1 N
1 =E SS(x x)
- k1 (8.58)
k=1 j>k

where, k,l are colloids in the REV, VREV is the volume of the REV, x is the ith
component of position of colloid k, and J' is the JIh component of the force applied
on colloid k from colloid 1.

The idea is then to relate the measured stress to quantities that can be controlled
experimentally and can be measured at a macroscopic scale. This link is based on
the understanding of colloidal forces and the controlling parameters of these forces.
In general, these forces are dependent upon the properties of the interstitial liquid,
like pH, ionic strength I, and dielectric permittivity Er. These forces can also depend
on the surface charge density a, of the solid particles, which in turn depends on
the properties of the interstitial solution and the constitution of the solid particles.
Another important parameter that can control the colloidal forces is the temperature
T.
These forces can also depend on the chemical composition of the interstitial solu-
tion, which might dictate the type of counter ions available between the interacting
surfaces. Relating the stress developed during hydration and the aforementioned pa-
rameters can then be incorporated in a multiscale micromechanics model to study
the contribution of precipitation and densification forces on the bulk volume changes
in cement-based materials [193-195]. Development of such relations can be equally
important to study the contribution of other colloidal forces (other than precipitation
forces) in hydrating and mature cement-based materials. Colloidal forces in cement-

224
based materials can be modified by changing the chemical and physical properties of

the pores solution including the solvent exchange. From a simulation point of view,

this requires the development of interaction potentials that take into account these

variables and then use these potentials to estimate stresses.

In the following we discuss and clarify the significance of the tensile stresses de-

veloped inside the simulated C-S-H gel as modeled by Masoero et al. [121]. In these

simulations, the precipitation of the C-S-H gel is simulated by successively and ran-

domly inserting particles into a cubic periodic simulation cell. At early stages of the

simulations and due to the absence of mechanical percolation, all the inserted parti-

cles are at equilibrium and the overall stress in the system is zero. Once mechanical

percolation if reached, however, insertion of a new particle will induce stresses in the

system because the existing particles are not free to rearrange anymore. Stresses

induced by insertion in a percolated system are generally tensile stresses because the

long range interactions are attractive. This can simply be understood as increasing

the attractive interactions per unit volume of the material which will result in reduc-

tion of the volume (eigenstrain) in an NPT ensemble. Such effect can be seen in the

virial stress Eq. (8.58).

In Eq. 8.58, and due to the attractive potential and keeping VREV constant and

inserting a new particle will result in an increase in tensile stress if the resultant of

the interactions of the newly inserted particle is attractive. The magnitude of this

eigenstress will depend on the PSD and the interaction potential between the particles.

The interaction potential used in these simulations is a generalized Lennard-Jones

potential form:

Uij (rij) = 46(&ij) I( 2) (On) (8.59)


(rij rj

The potential is parameterized to describe the physical properties of C-S-H in mature

cement paste. Other potentials are possible. E.g. Ioannidou et al.'s potential [96]

for early hydration (eqn. 8.60), and can give different results for the volume changes,

where the potential is based on generalized Lennard-Jones attraction and Yukawa

225
repulsion and based on the experimental measurements by Plassard et al. [154].

U(r) 4E - 21 -
rI )r
+ AC I (8.60)

As a consequence of the insertion procedures in Masoero's simulations, the system is in


tension at the termination of insertion because of the strong inter-particle attraction

and because the structural rearrangements are carried out under pVT condition. An

example of the evolution of such stresses is shown in Figure 8-24.

900 1
POLY 10 llini = 0.60
800 - POLY 5 (S 0)
700 POLY 2
MONO
- 600
* 500 1
rm = 0.68
400 (6 = 0.19)

.RP 300
-

(b)
200

100 hin1 = 0.75


0 (6 = 0.47)

-100
0 0.2 0.4 0.6 0.8
(Vor
Packing density

Figure 8-24: Virial stress in the system resulting from the simulated precipitation
process. Tension due to jammning of "out-of-equilibrium" colloidal systems where
(lensification (increase in packiniig density) is due to progressioi of the hvdrationm re-
actions. Eigemnstress as calculated from Virial Stress at the end of Insertion Delay ill
'T enisemble. Delay niot log enough (precipitation too fast) to reach equilibrium
from the work of M\asoero et al. [121, 122]. The value of eigenstress and associated
volume camiges developed iin the simulated material depenid to some extent on the
mmechaiism of precipitation. the deep interaction well, the shape amid polydispersity
of the particles. and the plT ensemble.

The interaction potential sed in Masoero s silmllations was parameterized to sim-

ulate the interactionms iii mature cemeit paste, so its applicability to hydratimg cenmelit

226
at early ages is not obvious. A more relevant approach to simulate the volume changes

during the hydration of cement at early ages is using potentials based on the ionic

correlation forces as appears in the work of Ioannidou et al. [96]. In Ioannidou's sim-

ulations and due to the presence of a repulsive shoulder (based on AFM experiments

by Plassard et al. [154]), the net stresses developed in the system were compressive

and swelling was measured upon relaxation of the system.

A more realistic case would be an NPT ensemble with account for the aqueous

solution and transformation of part of the solution to form the particles. Thus far,

both the work of Masoero and Ioannidou have worked with implicit solution and ne-

glected the effect of the solution on the pressure. Hence, only after setting we can

impose a control of the pressure, alternating pVT (Grand Canonical) with NPT (iso-

baric), to have both precipitation and control over pressure to measure the associated

shrinkage.

8.5 Summary

Bulk volume changes during the hydration of C 3 S showed a general behavior of ex-

pansion during early age followed by shrinkage at later ages. The expansion part of

the bulk volume changes are interpreted in light of the findings during the hydration

of MgO; and is related to the difference in volume between the reactive solids and the

hydration products. The solid volume change in conjunction of the reaction mecha-

nisms are then employed to model the expansion. To interpret the shrinkage strain on

the other hand, a multiscale thought-model is developed for the saturated hydrating

C3 S paste. As shown schematically in Figure 8-10, the first level (level I) of the model
includes the C-S-H colloids and the interstitial space also known as gel porosity. The

second level (level II) includes level I in addition to the capillary porosity. Finally,

residual (un-reacted) cement grains and other non-reactive solid inclusions embed-

ded inside the porous matrix define the continuum scale (level III). Driving forces of

bulk volume changes are modeled as eigenstresses at the corresponding scales. It was

possible with this model to quantify the development of eigenstresses during the hy-

227
dration reaction and relate it to the main parameters such as the degree of hydration,

rate of reaction, packing density of the C-S-H gel, and the solid volume fractions.

The resulting eigenstresses are then attributed to the densification of the C-S-H gel

resulting from an out-of-equilibrium precipitation of the C-S-H particles.

228
Chapter 9

Expansion of #-C 2S
The main differences between C 3 S and C 2 S is the quantity of CH formed during the

hydration reaction and the rate of the reaction. The difference in the formed CH is
used to assess the contribution of CH to the measured bulk volume changes.

9.1 Kinetics of C 2 S hydration

The hydration kinetics of C 2 S is monitored with both the enthalpy of the reaction
in isothermal calorimeter and the volume of the reaction in the pressure vessel (see
figure 6-16 and 6-17). The relationship between the measurements is fairly linear
with a ratio of 7.60, 7.46, and 7.23 J/pL for the fine, medium and coarse materials,
respectively. Assuming the enthalpy of hydration of C 2 S to be Hrxn = 260 J/g [182],
the corresponding volume of the reaction is 35.00 t 0.88 /L/gc s.
2

9.2 Evolution of volume fractions during the hy-


dration of C 2 S

To estimate the evolution of the volume fractions of the different components of the
system, we consider the chemical reaction below along with material properties in

229
-Hrxn (J) (C2S-F-W45-T40)
- Hrxn (J) (C2S-M-W45-T40)
Hrxn (J) (C2S-G-W45-T40)
100

-
90

-
80

-
70

-
60

-
x 50

-
= 40

-
30

-
20
10 -
-
0
0 5 10 15
Vrxn (pL/g)

Figure 9-1: The relationship between the volume and enthalpy of reaction as measured
with isothermal calorimetry and the pressure vessel for C 2 S samples with different
PSD cured at 40'C with w/c = 0.45. The relationship is fairly linear with a ratio
of 7.60, 7.46, and 7.23 J/pL for the fine, medium and coarse materials, respectively.
Assuming the enthalpy of hydration of C 2 S to be H,,, = 260 J/g, the corresponding
volume of the reaction is 35.00 0.88 pIL/gc 2 s.

Table A.2. the reaction is based on the forntulation of Young et al. [212].

C2 S + (2-x+y)H , CSHY + (2-x)CH (9.1)

where x is the C/S ratio and y is the H/S molar ratio. For the chemical reaction

considered here x = 1.7 and y = 2.1. The resulting balanced chemical reaction is:

CS+ 2.4 H -" C. 7 SH2 .1 + 0.3 CH (9.2)

The volume fractious of the different components are then:

fr-n = (Rh, - RH 1)(1 -n) (9.3)

f(s= (1-no)(l - (9.4)


)

,I =- - (1 - O)R (9.5)

21 )RCsH (9.6)

230

I
fCH = (1 - no)RCH (9.7)

fgel - (1 - no)Rgei (9.8)

And the total porosity of the system is:

otot = 1 - fC 2 s - fAp

= no - (1 - no)(Rhp - 1)

while the capillary porosity is:

cap = 1 - f(C2 s) - fgel - fCH (9.10)

Finally, the gel porosity is:

bgel = Otot - #cap (9.11)

where Rh (2
-x)V H +VCSH is the ratio between the molar volumes of the hydration
products and that of C 2 S, RH = (2-x+y)VX
VU26
is the ratio between the molar volumes
of water and the reactive solid (C 2 S), RCSH is the ratio between the molar volume
of C-S-H and C 2S. frxn is the fractional volume of the reaction, f(C 2 s) is the volume
fraction occupied by the remaining non-porous solid C 2S, fmw is the volume fraction
occupied by the remaining initial water used for mixing (initial water is treated as
a separate phase just like solids), fCSH is the volume fraction occupied by C-S-H
as defined by RCSH, no is the initial porosity upon mixing, and is the fractional
conversion (degree of hydration). fhp is the volume fraction occupied by the hydration
products CH and C-S-H (see Figure 9-2).

Figure 9-2 presents the evolution of volume fraction of the various components
during the hydration of C 2S with water-to-solid ratio ofw/c = 0.34 and w/c = 0.42.
The yellow spot marks the degree of hydration where all the remaining free water is
inside the C-S-H gel. fm w is the volume fraction occupied by the free water remaining
from the original mixing water only (not including any water provided from the
outside during the hydration). fe, is the volume fraction occupied by free water

231
remaining from the initial mixing water in the capillary porosity. tCusH is the volume

fraction of the solid C-S-H only. < e1 is the volume fraction occupied by the gel

porosity. The capillary porosity $Qa,is represented by the areas shaded in blue and

grey. The grey area represents part of the water uptake from outside the sample

(#Oca - fc,). The total porosity of the system is #tot = Ocap + 0ge4. The volume

fraction occupied by the C-S-H solids and the gel porosity is defined as the C-S-H gel

.teg = 1Cs1 + OPeg.

For more details on the definition of each term. see Appendix A.

0.9 0.9

0.8 0.8 Sfill?-

0.7 0.7
fc~l, fillsu
0.6 0.6
fc s--
-0. 4
-: 0.4 0.4 CSSH

0.3 0.3

0.2 0.2
fC, S
0.1 0. 1

ol,
0 0.2 0.4 0.6 0.8 1 k0 0.2 0.4 0.6 0.8 1
c c

(a) ui/c =0.34 (b) in/c =0.49

Figure 9-2: Evolution of volume fraction of the various components during the liy-
dration of C2 S with water-to-solid ratio of (a) 't/c = 0.34 and (b) Iv/c = 0.42. The
yellow spot imarks the degree of hydration where all the remaining free water is inside
the C-S-H gel. fmw is the volume fraction occupied by the free water remaining from
the original mliximng water only (not including any water provided from the outside
(luring the hydration). fe, is the volume fraction occupied by free water remaining
from the initial mixing water in the capillary porosity. fcS' is the volume fraction
of the solid C-S-H only. qe,, is the volume fraction occupied by the gel porosity. The
capillary porosity #, is represented by the areas shaded in blue and grey. The grey
area represents part of the water uptake from outside the sample (.o(, - fe,). The
total porosity of the system is #+og=017) + #qc1. The volume fraction occupied by
the C-S-H solids and the gel porosity is defined as the C-S-H gel fe =1(5k] + 9c/.

232

--I
9.3 Eigenstress development during the hydration

of C 2S

The main difference between C 3 S and C 2 S pastes is the amount of CH formed during
the hydration as can be seen in the evolution of volume fractions shown in Figure 9-2
when compared to Figure 8-8. The similar magnitude of expansion in C3 S and C 2 S
pastes indicate that the observed expansion is due to the formation of C-S-H rather
than the formation of CH. This observation supports the conclusion that CH precip-
itates mainly in the water-filled space and does not contribute to the bulk volume
changes. The development of eigenstresses during the hydration of C 2 S is shown in
Figure 9-3 for the different particle sizes as function of the degree of hydration and the
solid volume fraction. Calculation of eigenstresses development during the hydration

of C 2 S is identical to that in C3 S with different evolution of volume fractions as shown


in Section 9.2.
Unlike C3 S and due to the very slow rate of hydration, the densification of C-S-H
does not seem to have been triggered as there is not enough solid volume fraction at
low degrees of hydration. The low degree of hydration and the resulting solid volume
fraction are the main reason that only expansion is observed during the hydration of

C 2 5.

9.4 Summary

The main difference between C 3 S and C 2 S pastes is the amount of CH formed during

the hydration. This chapter presents the discussion of the results obtained from the
hydration of C 2 S at 40'C with different particle size distributions and w/c ratios.
Unlike C 3 S and due to the very slow rate of hydration of C 2 S, the densification of
C-S-H does not seem to have been triggered as there is not enough solid volume
fraction at low degrees of hydration. The low degree of hydration and the resulting
solid volume fraction are the main reason that only expansion is observed during
the hydration of C 2 S. The similar magnitude of expansion in C 3 S and C 2 S pastes

233
-- * (C2S-F-RD-W42-T40) -* (C2S-F-RD-W42-T40)
-- * (C2S-M-RD-W38-T40) - * (C2S-M-RD-W38-T40)
a* (C2S-G-RD-W34-T40) a* (C2S-G-RD-W34-T40)
SVF
0.4 0.5 0 .55 0.6 0.65 0.7
0.1 0.2 0.3 -.
0
0
-5 -5

-10- -10

-15 -15
,_' -20
-20 - ___

-25 -25
N.
-30 -30

-35 -35

-40 -40
-

(a) (b)

Figure 9-3: Development of eigenstresses as calculated from the multiscale microp-


oromechanics model for for C2 S with three different particle size distributions. (a) as
a function of the degree of hydration; (b) as a function of the solid volume fraction
as defined in Eq. (8.56).

indicate that the observed expansion is due to the formation of C-S-H rather than

the formation of CH. This observation supports the conclusion that CH precipitates

mainly in the water-filled space and does not contribute to the bulk volume changes.

234

-
Chapter 10

Development of Eigenstresses
During the Hydration of Class-G
Cement

Hydration of Class G cement comprises a number of simultaneous hydration reactions.

A deconvolution of these reactions is necessary to understand the bulk volume changes

of such material. A simplified model for the deconvolution of the hydration reactions

of Class G cement is presented in Section 10.1. At early ages, the hydration reactions

involve the formation of ettringite, monosulfates and hydrogarnet from the hydration

of C 3 A and C 4 AF. These phases can have a significant effect on the measured bulk
volume changes during the first 50 hours of hydration of Class G cement in ambient

conditions. These reactions are marked by the sulfate depletion peak in the enthalpy

or volume of the reaction measurements.

The investigation of Class G cement starts by studying the impact of permeability

and the possibility of significant pressure drop across the sample. For this purpose,

samples with different w/c were investigated in the PVC configuration. For w/c of

0.42, the drop in pressure at the sealed surface was negligible while for lower w/c

the pressure drop was as high as a few MPa's as shown in Figure 6-24. Sample

dimensions and w/c ratios were chosen such that the contribution of permeability

would be assumed to be negligible.

235
We start this chapter by discussing the available models for the evolution of vol-
ume fractions followed by the simplified calculations used in this study. Analysis of
the bulk volume changes is then presented and discussed based on the multiscale mi-
croporomechanics model developed for the hydration of C 3 S and modified here for the
CGC. Measurements from both RDC and LBC are discussed here and a comparison
is presented to explain the discrepancy found between the two configurations.

10.1 Evolution of volume fractions during the hy-


dration of Class-G cement

Calculation of the evolution of volume fractions occupied by the various components


of the cement paste during the hydration is of critical importance to the multiscale
microporomechanics model. Of special interest in such calculations is the evolution of
the volume fractions of the main constituents of cement powder as well as the various
hydration products and porosities.

Powers and Brownyard model

A great example of volume fraction calculations can be found in the work of Powers
and Brownyard [157,159,160]. The calculation of volume fractions in this model is
mainly based on thermal analysis of hardened cement paste and the non-evaporable
water content and water vapor isotherms. In their model of hardened cement paste,
it is assumed that the volume of the material is made up of unreacted cement grains,
cement gel, and capillary porosity. The volume fractions were summarized by Hansen
[85] as follows:

The total volume of cement paste: The total volume of the cement paste in
powers model is composed of unreacted cement and components derived from the
evaporable and non-evaporable water as related to the properties of the hydration
products

Vot = cV + WnVn + (wg + wc)V (10.1)

236
where c is the weight of cement in the volume considered in grams, w" is the weight of
non-evaporable water in grams, we is the weight of capillary water in grams, wg is the
weight of gel water in grams, V is the specific volume of cement in cm 3 /g, Vd is the
specific volume of capillary and gel water ~ lcm 3 /g, and V, is the specific volume of
non-evaporable water in cm 3 /g. wn, we, and wg are measured for cement paste with
c grams of originally unreacted cement. In the absence of bulk volume changes, the
original volume is also the sum of the volume of the mixing water and the original
unreacted cement

Vot = cV + woV (10.2)

where wc is the weight of original mixing water in grams, and Vd is the specific volume
of mixing water - 1cm 3 /g.

Volume fraction of cement gel: The volume fraction of the "cement gel"

Vgei = cV + WnV + WgVd (10.3)

where Vgel is the total volume of cement gel in cm 3 /gcement and is the degree of
hydration or the mass fraction of cement consumed in the hydration reaction =

chIc = wn/wn, where ch is the mass of cement consumed in the reaction and wO is
the mass of non-evaporable water at complete hydration. w9 = akwn is the weight of
gel water, where the cement gel is assumed to form a colloidal structure encompassing
the gel pores. Part of the evaporable water is assumed to originate from the gel pores
as represented by the water vapor isotherm, while the rest of the evaporable water
originates from the capillary porosity. The weight of gel water wg is approximately
3-4 time the quantity of adsorbate required for the first adsorbed layer on the solid
surface Vm. The capillary water is defined as the remainder of the evaporable water.
A good approximation of the gel water mass found by Powers is given by:

W9 = aVm (10.4)

237
where a ~ 3.3 The quantity of the adsorbate required to form a single layer on the

solid surface, is

Vm = kw, (10.5)

where k depends of the nature of the hydration products and consequently on the

composition of the cement powder:

k = 0.230(C 3S) + 0.320(C 2S) + 0.317(C 3A) + 0.368(C 4 AF) (10.6)

Finally, the volume fraction occupied by the cement gel is

Vge mcV + (V + ak)w, mV + (V + ak)wn/c (10.7)


Vtot wo + cV (wo/c) +V

Volume fraction of gel water: The volume fraction occupied by the gel water

(gel porosity) can then be calculated based on the mass of gel water:

(,Ogel WgVd - akw - akw/c .0.28 (10.8)


Vtot WO + cVe (wo/c) +V

Volume fraction of capillary pores: The volume fraction occupied by the capil-

water:

WCVd _ wo - (V + ak)wn _ (wo/c) - (V, + ak(w,/c)) (10.9)


Scap wo+cV (wo/c)+V

where w, = wo - (V + ak)wn is derived from Eq. (10.1) and (10.2).

Volume fraction of the unhydrated cement: The volume fraction of unhy-

drated cement (remaining cement at degree of hydration ') can be expressed in terms

of the hydration degree and the original mass of cement as

fun cuhVc - ( 0+cVc - (I -) Vc (10.10)


tot WO+cVe (WOMc+Ve

238
where Cuh is the mass of unhydrated cement in grams. For the calculation of the
degree of hydration, Powers relied mainly on the amount of non-evaporable water as
defined previously with the weight of non-evaporable water at complete hydration
estimated relative to the mass of cement as

Wn = 0.187(C 3S) + 0.158(C 2 S) + 0.665(C A) + 0.213(C AF)


3 4 (10.11)
C

Powers model is also used to estimate other properties of the cement gel such as
porosity, surface area, density of the solid gel particles. It is also used to estimate
constraints on the degree of hydration as related to the initial water-to-cement ratio.
The cement gel in Power's model included all the hydration products (colloidal or non-
colloidal) whereas here, we need to make a distinction between the different hydration
products such as C-S-H, portlandite, ettringite, and hydrogarnet.

Parrot and Killoh model

Parrot and Killoh [142,143] developed a model to predict the degree of hydration of
each of the main cement phases, enthalpy of the reaction, and bound water content
during the hydration reaction of cement-based materials. The input parameters of the
model included the water-to-cement ratio, moisture history, composition of cement,
and the specific surface area of the cement powder. Similar to Powers and Brownyard
model, Parrot and Killoh do not distinguish between the different cement hydrates.

Jennings and Tennis

Another model for the calculations of volume fractions can be found in the work of
Jennings and Tennis [100], where they based their calculations on the composition of
cement, degree of hydration, and the initial water-to-cement ratio to estimate volume
fractions of AFm, CH, C-S-H, gel porosity and capillary porosity. In their model,
Jenning and Tennis, considered only four hydration reactions, namely, the hydration

of C3 S and C 2 S to form CH and C-S-H and the hydration reactions of C 3 A with water
and gypsum to form monosulfates and the reaction of C 4 AF with water and gypsum

239
to form trisulfates and hydroxides.
To estimate the degree of hydration of each of the main four phases, a set of
Avrami-type equations are used in the general form:

i= 1 - exp(-a(t - bj)ci) (10.12)

where j is the degree of hydration of phase i and t is the time of hydration. The
constants aj, bi, and ci are determined empirically based on the work of Taylor [180]
and listed in Table 10.1.

Table 10.1: Constants used in the Avrami Eq. 10.12

Compound a b c
C3 S 0.25 0.9 0.7
C2 S 0.46 0.0 0.12
C3 A 0.28 0.9 0.77
C 4 AF 0.26 0.9 0.55

The volume fractions are then obtained as follows:

Vunhydratedcement c(1 - 1
)

1 1
VCH - c(i-N1P1I9 -- -JO582P2)

VAFm c(0.84963P 3 + 0. 4 7 2 4P4)

VC-S-Hsolid c(0.278 ipi + 0. 3 6 9 2P2) (10.13)


Vcapillarypores (1 - c) - C Piii)

VC-S-Hpores = 0. 2 19VC-S-Hsolid

totalpores = VC-S-Hpores + Vcapillarypores

where V, is the volume of component x in 1 g of paste. c is the initial mass of cement.


j is the degree of hydration of phase i and pi is the mass fraction of phase i in cement,

where i = 1 refers to C 3 S, i = 2 for C 2 S, i = 3 for C 3A, and i = 4 for C 4 AF. Pcem

is the average density of cement and Aj is the volume of the reaction based on the

differences in molar volumes between the reactants and the products. The constant

0.219 in VC-S-Hpores is determined based on the difference in specific volume between

240
saturated C-S-H gel and the C-S-H in the D-dry state. The volume fraction of each
of the components can then be estimated by dividing the corresponding volume by
the total volume of the paste.
The model of Powers and Brownyard relies on sorption isotherms and thermal
analysis to calculate the evolution of the volume fractions as a function of the degree
of hydration, which in turn is calculated based on the non-evaporable water. On
the other hand, the model developed by Jennings and Tennis combined the premise
of Powers model with stoichiometric calculations based on approximate hydration
reactions to calculate the volume fractions. Similar to Jennings model, a combination
of thermal analysis and stoichiometry to estimate the evolution of volume fractions
can be found in the work of Young and Hansen [212]. Other attempts at calculating
the volume fractions can be found in the work of Bernard et al. [20], developed as
part of a multiscale microporomechanics model, which was also applied by Pichler et
al. [153] for the same purpose. Another model can be found in the work of Sanahuja
et al. [168].

Evolution of volume fractions in this study

A more accurate model would require a comprehensive understanding of the effect of


the different mix parameters on the properties of each phase present in cement-based
materials and the progression of each of the individual chemical reaction.
To estimate the volume fractions of the different components, the hydration reac-
tions are separated into silicate, aluminate and sulfate reactions. To deconvolute the
hydration reactions into sulfate, aluminate, and silicate reactions, we combine the ex-
perimental measurements of enthalpy and volume of the reaction in conjunction with
a set of known chemical reactions. The typical calorimetry curve of cement-based
materials includes the well-known events including the sulfate depletion, initiation
of the fast C 3 A hydration, and the silicate hydration peak (main peak). It is rele-
vant here to describe the typical enthalpy curve of cement-based materials and the
equally important, but almost absent from the cement literature, volume of reaction
(chemical shrinkage) curve.

241
4

3
C S+Gyu H-r-
Xf
-3 2 3

1 10 100
Time (hr)
Figure 10-1: Typical hydration curve of class G cement. The material used in this
test was class G passing 20pim sieve. Point 1 marks the depletion of sulfate and the
onset of C 3 A hydration marked by point 2. Point 3 is typically attributed to the
hydration of C 4 AF.

To be able to deconvolute the chemical reactions, a set of chemical reactions is re-

quired. Table 10.2 lists compiled chemical reactions that occur during the hydration

of CBM required mainly to relate the enthalpy of reaction to the volume of reac-

tion. A number of measurements (like enthalpy and volume of the reaction) equal to

the number of chemical reactions is required for the complete deconvolution of the

chemical reactions. Only enthalpy and volume of the reaction are employed in this

analysis. For this reason, we will deconvolute the reactions into aluminates and sili-

cates reactions only. This simplification of the chemical reactions will prove helpful in

quantifying the volume (and mass) fractions of the several phases in hydrating CBM.

242
Table 10.2: Typical hydration reactions in cement-based materials. Cement chemist notation is shown in Table A.1

Reaction Reference AHrx AVxn Reaction


Compound (J/g) (pL/g)
CSHO.5 + 1.5 H M CSH 2 CSH o 5. -110 -41.1 RXN1
C3 S + 5.3 H M C 1 .7SH 4 + 1.3 CH C3 S -520 -67.2 RXN2
C 2 S + 4.3 H M C1 .7SH 4 + 0.3 CH C2S -260 -67.4 RXN3
2 C 3A + 21 H T C 4 AH 13 +C 2 AH 8 C3 A -1062 -243 RXN4
C3 A + 6 H M C 3 AH6 C3 A -910 -174 RXN5
C 2 AH 8 +C 4 AH 1 3 M 2 C 3AH 6 +H C 2 AH 8 230 11 RXN6
C 3 A +3 CSH 2 + 26 H C6 AS 3 H 32 C3 A -1670 -166 RXN7
2 C 3A + C6 AS 3 H 32 + 4 H 3 C 4ASH 12 C 3A -882 -85.6 RXN8
C 4 AF +3 CSH 2 + 30H H C6 AS 3 H 32 + CH + FH 3 C 4 AF 804 -152 RXN9
2C 4 AF + C6 AS 3 H 3 2 +12H M 3C 4ASH 12 +2CH +2FH 3 C 4 AF 2504 -88 RXN1O
C 4 AF + 10H C 3 AH6 + CH + FH 3 C 4 AF 1254 -115 RXN11
M+H MH M -934 -109 RXN12
Enthalpy of the reaction (H,,n) and volume of the reaction (Vx) as measured

between the time of mixing and the complete consumption of gypsum also known as
sulfate depletion (point 1 in Figure 10-1) are due to the hydration of C 3 S to form CH
and C-S-H, and due to the hydration of C3 A and C 4 AF to form ettringite only. The
measured Hn and V_n during this stage are attributed to two categories of reactions,
the first is the silicate reaction, which is simply the hydration of C 3 S assuming minor
contribution of C 2 S according to reaction RXN2 in Table 10.2. The second category
is the sulfate reaction with C 3A and C 4 AF to form ettringite according to reactions
RXN7 and RXN9 in Table 10.2. The rate of hydration of C 4 AF is assumed to be 75%
that of C 3 A rate based on published data of reaction rates (see for example [8,91,182]).

Once the sulfate has been depleted, the hydration of C 4 AF is assumed to stop
completely and the hydration of C 3 A is assumed to go according to RXN4 in Table
10.2. The hydration of C 3 S proceeds independent of the sulfate depletion.

The sulfate (gypsum) depletion is marked by the beginning of the second peak in
the hydration rate curve (Hrxn or Vrxn) and it typically occurs at about 16 hours for

typical OPC at ambient conditions. The rate of gypsum hydration is assumed con-
stant and complete hydration is achieved during these 16 hours. The mass fraction
of gypsum is estimated based on the mass fraction of SO 3 according to the modi-
fied Bogue calculations [181] by dividing the sulfate into alkali sulfates and calcium
sulfates.

Composition of cement powder is typically reported in mass fractions of the dif-


ferent phases as shown in Table 3.4. To calculate the initial volume fractions of
these phases, we start by specifying the water-to-solid mass ratio (w/c) and the initial
porosity, where the initial water-to-solid mass ratio is

W mw _ fm9wPw (10.14)
mr fropc

where w/c, the initial water-to-solid mass ratio, is a pre-specified quantity, mo is the
initial mass fraction of the mixing water, mo is the initial mass fraction of the reactive
solids in the mix at time t = 0, and fl, is the initial volume fraction of the reactive

244
solids in the mix. The initial porosity of the system can then be estimated based

on the initial water-to-solid mass ratio and the densities of the initial components

(cement powder)

w/Peem (10.15)

The average density of cement powder can be calculated from:

1
Pcem = (P) =MI
(10.16)

The mass fraction is related to the volume fraction by:

mi = Pf ; (p) = pifi (10.17)


(P)

where pi and fi are the density and volume fraction of phase i, respectively. m is the
mass fraction of cement phase i relative to the cement powder. The average density
of Class G cement used in this study and based on the mass fractions in Table 3.4 is
3.25 g/mL.

For a polyphase material like CBM, the degree of hydration can be defined as the
weighted sum of the degrees of hydration of all the phases:

(i(10.18)
( = m

where j is the degree of hydration of phase i. Similarly we define the volume of


reaction and enthalpy of reaction as the weighted sum of the contributions of each
cement phase to the total measured quantity

Vrxn = mVrin and Hxn = sm|Hrxn (10.19)

The mass fractions of C3 S and C 2 S are known directly from Table 3.4, while the mass
fraction of C 3 A and C4 AF are corrected to account for reactions with CSH 2 . The
mass fraction of CSH 2 is considered separately and its contribution to the measured

245
volume and enthalpy of reaction is based on the reactions to form ettringite. The

volume fraction of C 3 A consumed through the reaction with CSH 2 is:

f C3A
f VSH2
Vt afC5H2 (10.20)
vV

where v is the stoichiometric coefficient of the corresponding component in the con-

sidered chemical reaction and a is the fraction of CSH 2 consumed in the reaction with

C 3 A. The same applies to C 4 AF:

efAF - VCAF bfCSH 2 (10.21)

where b is the fraction of CSH 2 consumed in the reaction with C 4 AF and a + b = 1.

The total volume fraction of ettringite can be calculated based on the gypsum

content from reactions RXN7 or RXN9. Hydration of the calcium sulfate hemi-

hydrate into gypsum is assumed to be fast enough such that it is not included in

the calculations. The remaining C 3A is assumed to react with water to form cal-

cium aluminate phases according to reaction RXN5. C4 AF reacts according to reac-

tion RXN11, with the remaining volume fractions given by f0T, = f 3A and

4F fC 4 AF - f 4AF RXN5 and RXN11 are assumed to proceed at a constant rate


until complete consumption of C 3 A and C 4 AF.

The mass fraction of C 3 A remaining after the reaction with CSH 2 is

Pc3A

Crr Pc(10.22)
Pcem

The same applies to C 4 AF

f rem2
PC 4 AF (f~F
mCorr PC4A 1-no ,I(10.23)
Pcem

The volume of the reaction and the enthalpy of the reaction measured during the

first stage is attributed to the hydration of C3 S and Gypsum. The contribution of

gypsum to the measured quantity is based on RXN7. The volume of the reaction

246
normalized with respect to the mass of C 3 A is -166 pL/gC3 A compared with gypsum

-
87 [zL/gC5H 2 . The enthalpy is -873 J/gCSH2 . The total volume of reaction attributed
to gypsum is - 8 7 mC5H
2, where mCSH2 is the mass fraction of gypsum. The total
enthalpy of hydration can then be calculated in J/g as follows (see for example [92,
146, 171]):

42
A Hrxn = 520mr 3 5 + 260mc2 s + 866meCor + 0mcoF sCrH2 (10.24)

Similarly, the volume of the reaction can be calculated in piL/g as follows

AVrxn = 58m' 3 5 + 35mc2 s + 174mcr + 115mcAF + CSH 2 (10.25)

Here we introduce the time at the sulfate depletion tdeHCSH 2 and the contribution of
CSH 2 to the rate of volume of the reaction during the first stage is

dVC5H2 -87Tn
r n deHCH2 (10.26)
dt
CSH 2
*

10.2 Bulk volume changes and multiscale microp-


oromechanics model for the hydration of Class-
G cement

Bulk volume changes of Class-G cement are measured for samples with different
particle size distributions and w/c under hydrostatic and oedometric boundary con-
ditions. In general, bulk volume changes of Class-G cement included an expansion
part at early age followed by shrinkage at later ages depending on w/c and PSD. The
choice of the w/c is mainly dictated by the PSD to prevent excessive sedimentation,
which can impede the measurement of bulk volume changes. A minimum w/c was
necessary to guarantee the workability of the slurry made of coarse cement powder.
Yet, this minimum proved too high for coarse samples due to shear thickening, which

247
0.9 0.9

0.8 fct fil fl, / 0.8 >1 fl 2I

0.7 fCH 0.7


+AFm
Eht

0.6 -------- ~0.6 fCH + .fEtt + AFm


11
-C--

0.5 CH

0.4 '-
-
0. 4 ------------------ --------------
fCS
0.3 0.3 .k'i-

0.1 J'' 0.1

0 0.2 0.4 0.6 0.8 10 0.2 0.4 0.6 0.8

(a) w/c 0.30 (b) w/c 0.42

Figure 10-2: Evolution of volume fraction of the various components during the
hydration of CGC with water-to-solid ratio of (a) w/c = 0.30 and (b) w/c = 0.42. The
yellow spot marks the degree of hydration where all the remaining free water is inside
the C-S-H gel. f 1 1, is the volume fraction occupied by the free water remaining from
the original mixing water only (not including any water provided from the outside
during the hydration). fe t is the volume fraction occupied by free water remaining
from the initial mixing water in the capillary porosity. fCSH is the volume fraction
of the solid C-S-H only. 9qd is the volume fraction occupied by the gel porosity. The
capillary porosity (/co is represented by the areas shaded in blue. The dark blue area
represents part of the water uptake from outside the sample (Oca, - ) The total
porosity of the system is cb, =- co + +ge. The volne fraction occupied by the
C-S-H solids and the gel porosity is defined as the C-S-H gel =fC[SH i Ogel-

248
resulted in large sedimentation.
To interpret the measured bulk volume changes during the hydration of saturated
Class-G cement samples, we employ a multiscale microporomechanics model similar
to the model developed for the hydration of C 3S in Section 8.3. The main difference
between the hydration of C3 S and Class-G cement is the evolution of the volume
fractions of the paste components as shown in Section 8.2.3 and Section 10.1. In
the simpler C 3 S hydration, the reactive solid is C 3 S and the hydration products are
CH and C-S-H gel only. In Class-G cement the reactive solids include C 3 S, C 2 S,

C3 A, C 4 AF, and CSH 2 and the hydration products include CH, C-S-H gel, Ettringite
C6 AS 3H 32 , monosulfate C 4ASH 12 , and other calcium aluminates like C 4 AH 13 and
C 2 AH8
.

As shown in Section 10.1, the hydration of gypsum and the formation of ettringite
occurs during the early age of hydration, mainly during the first 16 hours at ambient
conditions (T = 25'C and p = 1 atm). The depletion of CSH 2 is followed by the
hydration of C3 A to form mainly C 4 AH13 and C 2 AH8 during the first 30 hours. The
formation of ettringite and AFm phases can lead to the observed expansion during
the first 30 hours of Class-G hydration as shown in Section 6.4.2. The behavior of the
material after the first 30 hours depends on the w/c and the particle size distribution.
To model this behavior, we employ the multiscale microporomechanics model with
the following volume fractions:

1. Level I includes the C-S-H gel as composed of the C-S-H solid and the gel
porosity. The eigenstress that can lead to contraction in the paste is attributed
to the gel formed at this scale:

=frCHi+g = fCSH _ fCSH


+ge -Tgel 1 ~ Pge(

2. Level II includes the C-S-H gel (Level I) in addition to the capillary porosity
similar to the model of the C 3 S:

P = f' + Ocap (10.28)

249
3. Level III includes the residual anhydrous cement grains in addition to all the

hydration products that do not, contribute to the (leveloinent of eigenstresses

as well as gel and capillary porosity and non-reactive solids.

The evolution of elastic properties of Class-G cement during hydration with the

volume fractions discussed in this section are shown in Figures 10-3. 10-4, and 10-5.

- -- ----- - 0.5_

-Level I Level 1
0. 8~ -Lvel II 0.4 -- Level 1.1
-Level III -- Level III

0 .6 0.3

0 .4 0.2

0 .2
01
0o 0.2 0.4 0.6 0.8 1 0 0.2 0.4 0.6 0.8 1

(a)

Figure 10-3: Evolution of the shear modulus G/gs during the hydration of CGC with
(a) w/c =- 0.30 and (b) w/c = 0.42 as calculated by the mmultiscale nicroporomnechan-
ics model for the three levels in function of hydration degree

1.4- 0.5--
1.2 -Level I -Level I
-Level 11 0.4 -- eel II
S-Level III -Level III

0.3
0.8

0.6 0.2
0.4
0. 1
0.2

0
-

0 0.2 0.4 0.6 0. 0 0.2 0.4 0.6 0.8

(1)
(a)

Figure 10-4: Evolution of tie bulk modulus K/Ai during the hydration of CGC with
(a) wc = 0.30 and (1) c = 0.42 as calculated by tlhe multiscale microporomeehan-
ics model for the three levels in funiction of hydratioii degree
Af_- - - -

-
0.95 0.95
0.9 0.9-
-0.85
0.8
.0.75 50.75

0.7 0.7
-- Level I -- Level I
0.65 -Level 11 0.65r -Level II
--- Level III --- Level III
0 0.2 0.4
C
0.6 0.8 1 0 0.2 0.4 0.6 0.8 1
(.

(a) (b)

Figure 10-5: Evolution of the Biot coefficient b during the hydration of CGC with (a)
'w/c = 0.30 and (b) w/c = 0.42 as calculated by the multiscale inicroporomliechanics
model for the three levels in function of hydration degree

10.2.1 Bulk volume changes in the Rubber Diaphragm Con-


figuration

The experimental results of the bulk volume change study on Class-G cement (CGC)

were presented in Section 6.4.2. Depending on the PSD, degree of hydration, and

solid volume fractions, bulk volume change of Class-G cement displays high rate

of expansion at early ages due to the hydration of aluminate and ferrite phases to

form ettringite and monosulfates as well as calcium aluminate hydrates. The rate of

early expansion is observed to change its rate upon the depletion of sulfate while it

continues to expand at a comparable rate due to the hydration of C3A and formation

of calcium aluninate hydrates. After the first ~30 hours of hydration, the expansion

typically continues at a lower rate due to the formation of all hydration Products in

a processes reminiscent of the expansion of MgO and C 3 5.

Shrinkage of CGC is observed at later ages as shown in Figures 10-6 and 10-7

which indicates that large solid volume fraction is required for the densification of

C-S-H to occur. For all samples shrinkage starts at solid volume fraction larger than
0.8 except for sample (CGC-F2-RD-W45-T25) in Figure 10-6 where shrinkage starts

at solid volume fraction of 0.76. The solid volume fraction for Class-G cement is

251

__
defined as follows:

SVTF = CH + f-s-Hos 1
-- + fAFm + fAFt (10.29)

The development of eigenstresses during the hydration of Class-G cement are

then calculated using the multiscale incroporomechanics model. The resulting eigen-

stresses due to the observed expansion and shrinkage are shown in Figure 10-8 and

Figure 10-9 as function of the degree of hydration and the solid volume fraction.

Figure 10-8 shows the results of coarse CGC compared to the reference CGC and
extreme expansion is observed in the sieved coarse sample; probably due to the very

low solid volume fraction. Figure 10-9 shows the evolution of eigenstresses in Class-G

cement with fine particle sizes. It is evident from both sets of tests that the initial

1/c ratio and the evolution of solid volume fraction play an important role in the

observed bulk volume changes and eigenstress development.

--- Ev (CGC-F2-RD-W45-T25) -- Ev (CGC-F2-RD-W45-T25)


Ev (CGC-F1-RD-W33-T25) -Ev (CGC-F1-RD-W33-T25)
Ev (Ref-CGC-Full-RD-W31-T25) Ev (Ref-CGC-Full-RD-W31-T25)
4 4

3
j __________

I
3

__ __

4
2 2

____________ ______
1 1

0 0

-1 -1
0 0.2 0.4 0. 6 0.8 0.4 0.5 0.6 0.7 0.8 0.9
SVF

(a) (b)

Figure 10-6: Bulk volume changes as measured during the hydration of Class-G
cenmet with fine particle size distribution and the reference material with full PSD
(a) as function of degree of hydration (b) as function of the solid volume fraction
assuming a constant packing density of the C-S-H gel at 0.64. Sample nomenclature
is shown in Figure 6-1.

I
252
Ev (CGC-Cl-RD-W26-T25) Ev (CGC-C1-RD-W26-T25)
- Ev (CGC-C2-RD-W25-T25) Ev (CGC-C2-RD-W25-T25)
Ev (Ref-CGC-Full-RD-W31-T25) Ev (Ref-CGC-Full-RD-W31-T25)
12 12

-
10 d - -- _ - -_ _ 10

-
8 8-
E
6

4 4
-

2 2

-
-

0-1
0 0.2 0.4 0.6 0.8 0.5 0.6 0.7 0.8 0.9
SVF

(a) (b)

Figure 10-7: Bulk volume changes as measured during the hydration of Class-G
ceieit with Coarse particle size distribution and the reference material with full
PSD (a) as function of degree of hydration (b) as function of the solid volume fraction
assuming a constant packing density of the C-S-H gel at 0.64. Sample noInenclature
is shown iii Figure 6-1.

I
y* (CGC-C1-RD-W26-T25) a* (CGC-C1-RD-W26-T25)
a* (CGC-C2-RD-W25-T25) -c* (CGC-C2-RD-W25-T25)
a* (CGC-Ref-FuI-RD-W31-T25) cr* (CGC-Ref-FuIl-RD-W31-T25)
SVF
0 0.2 0.4 0.6 0.8 0.5 0.6 0.7 0.8 0.9
0- 0
-

-500 -500-

-1000 -c -1000
-
-

-1500 -1500
-

-2000 -2000
-

-2500 -2500
-
-

(a) (b)

Figure 10-8: Development of eigenstresses as calculated from the multiscale microp-


oroiechanics model for different PSD and v/c ratios of class G-Ceient (CGC). (a)
in function of degree of hydration and (b) as function of the solid volume fraction. A
reference material with (Full) PSD is colnpared to coarse CGC as obtained by sieving
(CI) and sedimentation (C2).

253

-I
- * (CGC-F2-RD-W45-T25) -- * (CGC-F2-RD-W45-T25)
--- * (CGC-F1-RD-W33-T25) -Y* (CGC-F1-RD-W33-T25)
(y* (CGC-Ref-FuIl-RD-W31-T25) a* (CGC-Ref-Full-RD-W31-T25)
SVF
0 0.2 0.4 0.6 0.8 1 0.4 0.5 0.6 0.7 0.8 0.9
100
100

-
0 0-

-100 -4--
-100

-200 -200

-
ci ci

-300 -300

-
-400 -400

-
-

-500 -500

-
-

(a) (b)

Figure 10-9: Development of eigenstresses as calculated from the nultiscale imicrop-


oromiechanics model for different PSD and w/c ratios of class G-Cement (CGC). (a)
in function of degree of hydration and (b) as function of the solid volume fraction. A
reference material with (Full) PSD is conipared to fine CGC as obtained by sieving
(Fl) and sedimentation (F2).

10.2.2 Bulk volume changes in the Liquid Barrier Configu-


ration

The second configuration employed to measure the bulk volume changes during the

hydration of CGC is the Liquid Barrier Configuration (LBC). A set of tests are carried

out on CGC with w/c = 0.42 at 25'C with different thicknesses and curing pressures

as listed in Table 3.5.

The liquid barrier test separates. by design. volumie change neasureients from the

water mass absorbed as discussed in Section 5.4. To achieve this separation, LBC test

is performed by placing the slurry directly inside the pressure vessel and covering with
approximately 10nun of the liquid barrier. The pressure and temperature (25'C) are

then applied and bulk volume changes and water uptake are monitored contiiously
for the duration of the test. The test starts with water access through the top port

only aid after a specific period of timie related to the setting of ceient. water is

allowed through the bottom port. Vith both top and bottom ports connected to the

saMe source of pressurized water. the effective stress developed inside the speciiimen

254
is negligible in the axial direction and depends on the volume changes in the radial
direction (see Section 5.4).

The liquid of choice for this test is the Poly[methyl(3,3,3-trifluoropropyl)siloxane],


Dow Corning FS-1265 chemically inert liquid. FS-1265 has a viscosity of 10000 cSt,
a specific gravity of 1.3, a surface tension of 28.7 mN/m, and a molecular weight
of 14000 g/mol. The values for the surface tension and contact angle of water in
cement are usually assumed to be similar to that of water and glass. The contact
angle between the siloxane, water, and cement is taken as 104' [104] (page 214).

As observed in Section 6.4, it was only possible to measure the bulk volume changes
after setting of the cement slurry and complete consumption of the bleed water. The
measured volumetric strain of the tested samples with different thicknesses and curing
pressures are shown in Figure (10-10). Similar to other bulk volume measurements,
the samples display large shrinkage that coincides with the known volume of the
reaction until setting and consumption of bleed water is complete. The results in
Figure 10-10 are initiated at this instance.

By comparing the strains measured in the RDC (Figure 10-6) and the LBC (Figure
10-10), a discrepancy is evident. The differences between the RDC and the LBC tests
are detailed here in an attempt to understand the discrepancy between the volumetric
strain measurements. The first difference is the dimension of the samples. In the RDC
all samples have similar size of about 46.7 mm in diameter and 25.4 mm in thickness
while the samples in the LBC have variable thickness with constant diameter of 53.9
mm (diameter of the pressure vessel). The impact of sample thickness may cause
some pressure distribution inside the sample due to the evolution of permeability and
the negative volume of the reaction. Shrinkage in thick samples is smaller than that
of thin samples prepared under identical conditions.

Another difference between LBC and RDC is the boundary conditions. In the
RDC test, the sample is exposed to a hydrostatic state of stress and controlled (very
small) effective stress. The liquid barrier test on the other hand, is constrained radially
with an effective stress in the axial direction that is dependent on the evolution of
permeability and the resulting drop in pore pressure due to the negative volume of

255
- Ev (CGC-LB-PO.2-W42-T25-H47.3) - Ev (CGC-LB-PO.2-W42-T25-H17.8)
Ev (CGC-LB-P.2-W42-T25-H18.6) - Ev (CGC-LB-P5-W42-T25-H12.4)
Ev (CGC-LB-P1-W42-T25-H44.1) Ev (CGC-LB-P10-W42-T25-H44.1)

0 0.2 0.4 0.6 0.8


0

-500 -- _ -___

-1000
-

-1500
-

5. -2000

-2500 __- -_

-3000

-3500 -- - -----

-4000
-

Figure 10-10: Volumetric strain (iii micro strains) as function of the degree of hydra-
tion of Class-G cement paste hydrated in the LBC test. Samples have different thick-
nesses (H in iii) and variable curing pressure (P in M\IPa) with constant w/c = 0.42
and constant temperature T =25'C.

256
(CGC-LB-PO.2-W42-T25-H47.3) -- a* (CGC-LB-PO.2-W42-T25-H17.8)
a* (CGC-LB-P.2-W42-T25-H 18.6) -- * (CGC-LB-P5-W42-T25-H 12.4)
0* (CGC-LB-P1-W42-T25-H 44.1) 0* (CGC-LB-P1O-W42-T25-H44.1)
600 1~~~~~~

500 i i -i__________ ________ ________________________ __________________________

400

/ /
-

0~
7
*

300 __ ____ ____/ __

200
____- ______ _____--- - -- I--- --7~- -~
-
/
/

100

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8

Figure 10-11: Development of eigenstresses (r* as function of the degree of hydration


of Class-G cement paste hydrated in the LBC test.

257
-- * (CGC-LB-PO.2-W42-T25-H47.3) --. ,* (CGC-LB-PO.2-W42-T25-H 17.8)
c* (CGC-LB-PO.2-W42-T25-H18.6) -* (CGC-LB-P5-W42-T25-H12.4)
a* (CGC-LB-P1-W42-T25-H44.1)

I-
* (CGC-LB-P1-W42-T25-H44.1)

600
-

500 I

400
-

300
*

200

100
-

0
0.55 0.6 0.65 0.7
1lgel

Figure 10-12: Development of eigenstresses as function of the C-S-H gel packing


density of Class-G cement paste hydrated in the LBC test.
the reaction. Radial stresses in the LBC depend on the volumetric strain as detailed
in Section 5.4. The ratio between the elastic deformations of the LBC to RDC strains
as described in Eq. (5.134) can range from 1/3 to 1/2 for Poisson's ratios of 0 and 0.2
respectively. That is, in all cases the strains measured in the LBC must be smaller
in magnitude that that measured in the RDC. Yet, this is not the case as seen in the
RDC (Figure 10-6) and the LBC (Figure 10-10).

One can speculate other possible causes of this discrepancy include possible for-
mation of water pockets between the barrier liquid and the top surface of the sample
and creep of the FS-1265 under pressure. The formation of water pockets was ob-
served in a test carried out in a separate container where such pockets formed by the
bleed water due to the difference in density of the liquids (see Figure 10-13). The
complication resulting from the formation of such pockets is mainly due to the small
fraction of sample surface exposed to water which takes longer time to be consumed
by the ongoing reaction and gives false shrinkage. To check for possible creep of the
FS-1265, a representative sample of FS-1265 was placed inside the pressure vessel
and the injected water volume was measured under a pressure of lOMP. Other than
the effect of temperature fluctuation, there was no significant volume change in the
system. FS-1265 does not creep (see Figure 10-14).
Another possible source of discrepancy between the LBC and RDC tests can be
speculated to be the ingression of the barrier liquid inside the specimen due to pressure
gradient resulting from low permeability and the negative volume of the reaction.
Evolution of the pressure for CGC samples with different w/c ratios is shown in Figure
6-24, where a pressure transducer is connected directly to the bottom port (port-3)
in Figure 5-1(b). The permeability of identical sample is measured directly with the
PVC configuration (Figure 5-1) by applying a pressure differential and measuring the
flow rate through the top and bottom ports.
To measure the volume ingress of FS-1265 inside the sample, we employ Eq.
(5.124) in Section 5.4, reproduced here for convenience:

VLB(t) = qAx(t) = A -_ J (Pw - Pt + PC dt) (10.30)

259
Water
pocket

Liquid barrier

Specimen

(a) (b)

Figure 10-13: Formation of water pockets between the liquid barrier and the top sur-
face of the specimen due to bleeding and the differences in densities. (a) a schematic
illustrating the geometry of the water pocket (b) water pockets as observed in similar
conditions to the LBC test. Water pockets can provide water to the specimen at a
very low rate resembling a false bulk volume change.

12
10

0
0 0 20 30 40 50 60 70
Time (hr)

Figure 10-14: Volume changes measulre( with only a sample of typical volunie of FS-
1265 inside the vessel. The measured volume is a result of temperature fluctuatiomis on
the syringe pumIp within 0jpL corresponding to alout less than 1% of the measured
bulk volume changes.

I
The cross sectional area of the specimen is A = 2.29 x 10-W2 , porosity # is esti-

mated from the evolution of volume fractions discussed previously, pw is the pressure

in water right underneath the liquid barrier, pt is the pressure on top of the liquid

barrier, pc is the capillary pressure, [1 is the viscosity of the liquid barrier (1 = 13

Pa.s), and K is the permeability of the specimen.

To measure permeability, two samples are investigated, the first sample was cured

under water for 6 days (144 hours) under 5 MPa pressure. Two pumps were used to

supply water to the top and bottom of the sample, the pressure at the bottom was

lowered to 1 MPa, resulting in nearly steady state flow over 22 hours with 4 MPa

differential pressure. Figure 10-15 shows flow into the sample through the top surface

(red) and flow out of the sample through the bottom surface (green) as measured
by the pumps. The difference in the measured flow rates coincides with the volume

of the reaction due to continuing hydration reaction. Using Darcy's law with the

average flow rates, the permeability was 3 x 10- 20 m2


.

Pressure in the pore solution right underneath the liquid barrier can be estimated

using Eq. 5.80, reproduced here:

h p dVxn Ms
PW =
W2 dt A8 + (10.31)

where A, is the area of the specimen, m, is the mass of the anhydrous cement, K is the

permeability of the specimen, p is the viscosity of the pore solution, h is the thickness

of the specimen, and PA is the curing pressure applied through the liquid barrier and

the bottom surface of the specimen. A sample calculation of the Ingression of the

liquid barrier inside the sample is shown in Figure 10-16, where the resulting strain

is in the order of the measured strains in the LBC. The 120pL at 200 hours for a 50

mL sample correspond to a strain of 0.0024 (2400 pE), which is comparable to the

strains shown in Figure 10-10.

261
I-
-

600 250

-
-'= 25.293t - 3602.7 200 11= 5.0004/- 956.34
400
150
200
100
-Vtop (pL-)
0 -_- Linrar (Vtop) E) 50
50 ) 9Vbot (pl-)
-- LiiI "ar ' )r
(VH Linear (Vtop L4)
0
-200 Linear (Vbot (pl))
-50
-400 -100 - -2to 2 2011+ 416.23
- 18.403t + 2605.2
-600 -150
140 150 160 170 190 200 210 220 230 240
Time (hr) Time (hr)

(a) (b)

Figure 10-15: Direct neasureinenit of permeability on Class-G ceient with w/c = 0.42
cured and 25'C curing temperature.(a) cured for 144 hr under 5 MPa curing pressure
with differential pressure of 4 MPa and thickness of 51.9 nun. the permeability is
3.4 x 10-20 (b)cured for 193 hr under 200 kPa curing pressure with pressure differential
2
of 0.8 IPa and thickness of 54.2 nn. the permeability was 3.0 x 10-201,

.
g
- phi=0.2 -phi=0.3 phi=0.4 --- phi=0.5

140
-

120

100

80 1 _---_-

60

40

20

0
0 50 1 00 150 200
Tim ? (hr)

Figure 10-16: Volnme of the liquid barrier ingressed isid( e the specimenildle to 50kPa
pressure dlifferential and 3 x 10-201m2 permeability. The measured vohunetric strain
is ablout 0.001. which is in the order of the llieasured strains in the LBC.

262
10.3 Summary

Bulk volume changes of Class-G cement have been investigated under two different
boundary conditions in the rubber diaphragm configuration (RDC) and the liquid
barrier configuration (LBC). In the RDC, expansion at early ages was observed fol-
lowed by shrinkage at later ages. Expansion is a result of the difference in volume
between the anhydrous solids and the hydration products. Eigenstresses are then
calculated based on the multiscale microporomechanics model. Eigenstresses seem to
be dependent on the solid volume fraction as is the case during the hydration of C 3 S.
Early expansion observed in CGC is attributed to the hydration of C3 A and C 4 AF
and the origin of this expansion is similar to the expansion of MgO and C3 S, and not
necessarily related to the crystallinity of the hydration products.
In the LBC on the other hand, only shrinkage is observed after setting and con-
sumption of the bleed water. Shrinkage is then modeled using the multiscale micro-
poromechanics model and the eigenstresses are calculated. The discrepancy between
the two boundary conditions is then discussed in light of the possible differences
between the RDC and the LBC, including sample dimensions, elastic stress field,
possible formation of water pockets, and the possible ingression of the liquid bar-
rier inside the sample. Thin samples in LBC showed larger volumetric strain than
thicker samples indicating that ingression of the liquid barrier might be a reason of
the observed shrinkage.

263
Part V

Conclusions and Perspectives

264
Chapter 11

Conclusions and Perspectives

This chapter presents a summary of the main findings of the study related to the
driving forces of bulk volume changes of cement-based materials in addition to the
main contribution of this thesis. Based on the findings of this study, the limitation
of the current study are presented and further research is suggested.

11.1 Summary of main findings

Bulk volume changes of cement-based materials are generally attributed to several


driving forces related to changes in relative humidity including capillary tension,
disjoining pressure, and surface energy (see Section 2.2 for more details). Such driving
forces require a drop in the degree of saturation (relative humidity), which is typically
attributed to self-desiccation (see for example [65]) or drying of the material due to
evaporation of the pore solution. In this study, samples are maintained saturated and
bulk shrinkage and expansion are still observed.
It is possible to monitor the hydration kinetics for all the investigated materials

(MgO, CA5, C 2 , CGC) with volume of reaction (chemical shrinkage) as well as


enthalpy of reaction. All material display a negative enthalpy and volume of reaction
as listed in Table 11.1. The negative volume of reaction entails reduction in the
absolute volume of the hydrating system associated with a significant increase in
the solid volume due to the difference in density between the reactive solids and

265
the hydration products. The increase in the solid volume during hydration can be

represented by the ratio of the total volume of the products relative to the volume of

the reactive solids, which is listed as Rhp in Table 11.1. Hydration of all investigated

materials is associated with more than 100% increase in the solid volume fraction.

The increase in the solid volume leads to the conclusion that an expansion should

be expected in all materials, yet, the experimental observations show otherwise. Dur-

ing the hydration of the simplest investigated material (MgO), expansion is observed

during the conversion of MgO to Mg(OH) 2 , representing only a small fraction of

the difference in volume between the reactive solid and the hydration product. On

the other hand, hydration of C 3 S displayed amore complex behavior starting with

expansion followed by shrinkage at later ages. The main difference between MgO

and C 3 S originates from the nature of cohesive forces in these materials. Cohesion

in the MgO-Mg(OH) 2 is attributed to development of ionic-covalent bonds between

the precipitating Mg(OH) 2 grains with low surface charge density in a process that

resembles sintering of ceramics. Cohesion in C 3 S, C 2 S, and Class-G cement, on the

other hand, is attributed to the strong electrostatic coupling between the colloidal

particles of C-S-H, which forms the binding phase in these materials.

The main difference between hydrating MgO and cement-based materials is the

presence of the colloidal C-S-H phase that binds all other phases through surface col-

loidal forces. A possible driving force of such bulk volume changes is the precipitation

and densification of the C-S-H gel as originally presented by Ulm et al. [195]. Densi-

fication of C-S-H is suggested to be driven mainly by the mean diffusion distance of

the solute and the solvent, which is controlled by the thickness of hydration products

surrounding the anhydrous cement grains and leads to the observed shrinkage. Thus,

the densification of C-S-H depends mainly on the w/c, PSD, and curing temperature.

It is found in this study that the main parameter controlling the densification of C-

S-H is the solid volume fraction which is a function of the w/c ratio and the degree

of hydration.

Densification of C-S-H and the resulting stresses are further supported by the

available meso-scale simulations of the precipitation and densification of C-S-H [96,

266
121,122,148] as discussed in Section 8.4.3 and Section 2.4. In these simulations, the

precipitation of the C-S-H gel is simulated by successively and randomly inserting

particles into a cubic periodic simulation cell. At early stages of the simulations and

due to the absence of mechanical percolation, all the inserted particles are close to

equilibrium and the overall stress in the system is close to zero. Once mechanical

percolation if reached, however, insertion of a new particle will induce stresses in the

system because the existing particles are not free to rearrange anymore. Stresses cal-

culated from the experimental results agree qualitatively with those from meso-scale

coarse-grained simulations of C-S-H precipitation originating from the electrostatic

coupling between charged C-S-H particles mediated by the electrolyte pore solution.

Expansion of the paste, on the other hand, is due to the fact that the volume

of the hydration products is larger that the original volume of cement powder. The

extent of the expansion depends on the mechanism of the hydration reaction as well as

the microstructure of the porous solid that is formed during the hydration reaction.

Depending on the reaction mechanism, only a fraction of the increase of the solid

volume can contribute to the observed bulk volume changes at the macroscopic scale.

Hydration of MgO is associated with an expansion that is linearly related to the

degree of hydration, indicating a constant fraction of the formed solid is contributing

to the expansion. The hydration of CBM, on the other hand, displayed a non-linear

relationship between the expansion and the degree of hydration due to the more

complex microstructure.

Table 11.1: A summary of the investigated materials and reaction properties.

AVn AHrxn Binder Cohesive


Material rnRhp
pL/g J/g R Phase Forces
MgO 109 930 2.21 Mg(OH) 2 Ionic-covalent
C3 S 58 520 2.16 C-S-H Colloidal
C2S 35 260 2.31 C-S-H Colloidal
CGC 60 413 2.13 C-S-H Colloidal

267
11.2 Research contribution

The contribution of this study can be classified in three parts relating to the main
experimental setup, the multiscale microporomechanics model, and the driving forces
of the bulk volume changes.

11.2.1 New experimental apparatus

For purpose of this study, a new experimental apparatus was designed and tested
on a number of materials with a range of chemical and physical properties. The
experimental apparatus was motivated by the lack of proper apparatuses capable of
simultaneous measurements of all the necessary poromechanical quantities; especially
the control of curing pressure, effective stress, and curing temperature. The main
advantage of the pressure control lies in the prevention of possible contribution of
permeability and self-desiccation due to "partially sealed" conditions. For this rea-
son, tests were carried out mainly under a pressure of 1 MPa to guarantee complete
saturation throughout the test and eliminate any possible contribution from hygral
forces like capillary pressure.

11.2.2 A multiscale microporomechanics model

A second contribution of this study is the implementation of a multiscale microp-


oromechanics model. This model is supported by extensive experimental investiga-
tions to predict the evolution of stress in cement-based materials during hydration
under saturated conditions.
A multiscale microporomechanics is developed to interpret the measured bulk vol-
ume changes of saturated hydrating C 3 S paste. The multiscale microporomechanics
model consists of a three-level model of the cement microstructure. In this model,
CBM is modeled as a matrix of C-S-H particles in which all other solid phases and
the pore space are embedded.
As shown schematically in Figure 8-10, the first level (level I) of the model includes
the C-S-H colloids and the interstitial space also known as gel porosity. The second

268
level (level II) includes level I in addition to the capillary porosity. Finally, residual

(un-reacted) cement grains and other non-reactive solid inclusions embedded inside
the porous matrix define the continuum scale (level III). Driving forces of bulk volume

changes are modeled as eigenstresses at the corresponding scales. The goal of the

micromechanical modeling is to establish the relationship between the macroscopic

observations and their finer-scale origin by considering the intrinsic properties of the

material constituents and the driving forces (eigenstresses) of bulk volume changes.

11.2.3 Driving forces of bulk volume changes in CBM

In saturated samples, the densification shrinkage starts due to the decrease of the

space available for hydration products to form and the initiation of precipitation

of C-S-H colloids inside the existing low density C-S-H. Precipitation of hydration

products can only occur on surface covered with pore solution or inside the pore

solution only and it cannot occur inside the gas (vapor) phase. Yet, the densification

of C-S-H that leads to shrinkage of the paste is dependent on the mean diffusion

distance, which in turn depends on the w/c, PSD, and curing temperature. The

densification of C-S-H resembles very closely the behavior observed for the sealed

samples as well as the saturated samples. Expansion on the other hand is attributed

to the difference in volume between the reactive solid and the hydration products.

11.3 Industrial benefits and impact

The main objectives of the primary cementing of an oil well is to support the casing

string and provide zonal isolation by filling and sealing the annular gap between the

casing string and the borehole [136]. Lack of understanding of the state of stress in the

cement sheath and how the sheath may fail can lead to loss of zonal isolation. Among

all the possible reasons that can lead to failure of cement sheath, bulk volume changes

that cement-based materials (CBM) experience during hydration was investigated in

this study.

When cement slurry develops its mechanical properties, it also undergoes volume

269
changes during the early age of hydration. If constrained, these volume changes will

lead to the development of stresses in the hydrating cement. Figure 11-1 shows the

development of macroscopic stress in a fully constrained material during hydration

of Class-G cement at 25'C with w/c = 0.42. To predict failure due to loading during

the service life of the well, one needs to understand initial stresses endured by these

volume changes. Prediction of stress and strain development in the cement sheath

at early ages is critical for performance evaluation. Yet, such a prediction calls for

a profound understanding of the chemical and physical mechanisms at early ages in

the cement sheath.

The main contribution of this study is a multiscale microporomechanics model.

This model is supported by extensive experimental investigations to predict the evolu-

tion of stress in cement-based materials during hydration under controlled conditions.

Besides predicting the evolution of stresses and strains in hydrating CBM, this model

also pins down the main parameters that control the evolution of the eigenstresses

and strains. Once the impact of these parameters on the bulk volume changes is

understood, only then it is possible to control the resulting stresses and performance

of the cement sheath or CBM in general. The implications of this model extend to

the development of an engineering model to predict the performance of the cement

sheath during the service life of the oil well.

11.4 Limitations and Perspectives

Results obtained with the rubber diaphragm configuration (RDC) and the liquid bar-

rier configuration (LBC) provided an apparent Poisson's ratio v = 0 for the hydrating

magnesium oxide if elastic deformations are assumed, yet, this is not the case during

the hydration of Class-G cement. The discrepancy in bulk volume changes of Class-

G cement as measured with RDC and LBC cannot bc explained with the Poisson's

ratio like is the case in the MgO. This discrepancy is attributed to several possible

differences between the RDC and LBC. Yet, a conclusive analysis was not possible.

Bulk volume changes have been studied at a macroscopic scale and it is found that

270
o* (MPa)
-250 -200 -150 -100 -50 0
5

-10

-15

Figure 11-1: Development of macroscopic (engineering) stress during the hydration


of Class-G cement of a fully constrained material in function of the solid eigenstress
determined from measurements.

the densification of C-S-H gel can be the driving force for the observed shrinkage in

saturated cement-based materials. Further investigation of the densification of the C-

S-H gel with nanoindentation in conjunction with water vapor and nitrogen sorption

isotherms can prove very helpful at understanding the densification and the associated

volume changes. It is of particular interest to investigate the water-to-cement ratio

on the mean diffusion distance and the resulting densification. The mean diffusion

distance can be studied by perneability measurements and impact of w/c ratio and

seeding on the rate of hydration.

At the mneso-scale simulations front, current studies indicate a possible contri-

bution of the simulated densification of C-S-H on the development of eigenstresses

during precipitation. The initial promising results of these simulations show the po-

tential of deeper understanding of the driving forces of bulk volume changes, yet a

systemiatic study of such contribution is not yet available.

271
272
Bibliography

[1] ABUHAIKAL, M., MUSSO, S., THOMAS, J. J., AND ULM, F.-J. An Apparatus
for Dissecting Volumetric Changes in Hydrating Cement Paste. In Mechanics
and Physics of Creep, Shrinkage, and Durability of Concrete (2013), pp. 316-
323.

[2] ALLEN, A. J., THOMAS, J. J., AND JENNINGS, H. M. Composition and


density of nanoscale calcium-silicate-hydrate in cement. Nature materials 6, 4
(apr 2007), 311-6.

[3] APFEL, R. E. The Role of Impurities in Cavitation Threshold Determination.


The Journal of the Acoustical Society of America 48, 5B (aug 1970), 1179.

[4] BABUSHKIN, V., MATVEEV, G., AND MCHEDLOV-PETROSJAN, 0. Thermo-


dynamics of silicates. 1985.

[5] BANGHAM, D. H., FAKHOURY, N., AND MOHAMED, A. F. The Swelling of


Charcoal. Part II. Some Factors Controlling the Expansion Caused by Water,
Benzene and Pyridine Vapours. Proceedings of the Royal Society A: Mathemat-
ical, Physical and Engineering Sciences 138, 834 (oct 1932), 162-183.

[6] BARCELO, L., BOIVIN, S., RIGAUD, S., AND ACKER, P. Linear vs. volu-
metric autogenous shrinkage measurement: material behaviour or experimental
artefact. Self-desiccation and its Importance in Concrete Technology (1999).

[7] BARCELO, L., MORANVILLE, M., AND CLAVAUD, B. Autogenous shrinkage of


concrete: a balance between autogenous swelling and self-desiccation. Cement
and Concrete Research 35, 1 (jan 2005), 177-183.

[8] BARNES, P., AND BENSTED, J. Structure and Performance of Cements, Sec-
ond Edition, vol. 2002. CRC Press, 2002.

[9] BAROGHEL-BOUNY, V., MOUNANGA, P., KHELIDJ, A., LOUKILI, A., AND
RAFAY, N. Autogenous deformations of cement pastes Part II. W/C effects,
micro-macro correlations, and threshold values. Cement and Concrete Research
36, 1 (jan 2006), 123-136.

[10] BELTZUNG, F., AND WITTMANN, F. H. Role of disjoining pressure in cement


based materials. Cement and Concrete Research 35, 12 (dec 2005), 2364-2370.

273
[11] BENTUR, A., IGARASHI, S.-I., AND KOVLER, K. Prevention of autogenous
shrinkage in high-strength concrete by internal curing using wet lightweight
aggregates. Cement and Concrete Research 31, 11 (nov 2001), 1587-1591.

[12] BENTZ, D., AND AITCIN, P. The Hidden Meaning of Water- Cement Ratio.
Concrete international30, 5 (2008), 51-54.

[13] BENTZ, D., GEIKER, M., AND HANSEN, K. Shrinkage-reducing admixtures


and early-age desiccation in cement pastes and mortars. Cement and Concrete
Research 31, 7 (jul 2001), 1075-1085.

[14] BENTZ, D., AND JENSEN, 0. Mitigation strategies for autogenous shrinkage
cracking. Cement and Concrete Composites 26, 6 (aug 2004), 677-685.

[15] BENTZ, D. P. A Three-Dimensional Cement Hydration and Microstructure


Program. I. Hydration Rate, Heat of Hydration, and Chemical Shrinkage, 1995.

[16] BENTZ, D. P. Influence of water-to-cement ratio on hydration kinetics: Simple


models based on spatial considerations. Cement and Concrete Research 36, 2
(feb 2006), 238-244.

[17] BENTZ, D. P., GARBOCZI, E. J., HAECKER, C. J., AND JENSEN, 0. M.


Effects of cement particle size distribution on performance properties of Port-
land cement-based materials. Cement and Concrete Research 29, 10 (oct 1999),
1663-1671.

[18] BENTZ, D. P., JENSEN, 0. M., HANSEN, K. K., OLESEN, J. F., STANG,
H., AND HAECKER, C.-J. Influence of Cement Particle-Size Distribution on
Early Age Autogenous Strains and Stresses in Cement-Based Materials. Journal
of the American Ceramic Society 84, 1 (jan 2001), 129-135.

[19] BERLINER, R., Popovici, M., HERWIG, K., BERLINER, M., JENNINGS, H.,
AND THOMAS, J. Quasielastic Neutron Scattering Study of the Effect of Water-
to-Cement Ratio on the Hydration Kinetics of Tricalcium Silicate. Cement and
Concrete Research 28, 2 (feb 1998), 231-243.

[20] BERNARD, 0., ULM, F.-J., AND LEMARCHAND, E. A multiscale


micromechanics-hydration model for the early-age elastic properties of cement-
based materials. Cement and Concrete Research 33, 9 (sep 2003), 1293-1309.

[21] B1oT, M. General theory of three dimensional consolidation. Journalof applied


physics 12, 2 (1941), 155-164.

[22] BIRCHALL, J., HOWARD, A. J., AND DOUBLE, D. D. Some general consider-
ations of amembrane/osmosis model for portland cement hydration. CEMENT
and CONCRETE RESEARCH 10, c (1980), 145-155.

274
[23] BISHNOI, S., AND SCRIVENER, K. L. Studying nucleation and growth kinetics
of alite hydration using pic. Cement and Concrete Research 39, 10 (oct 2009),
849-860.

[24] BJONTEGAARD, 0., HAMMER, T., AND SELLEVOLD, E. J. On the measure-


ment of free deformation of early age cement paste and concrete. Cement and
Concrete Composites 26, 5 (jul 2004), 427-435.

[25] BOCQUET, L., CHARLAIX, B., AND RESTAGNO, F. Physics of humid granular
media. Comptes Rendus Physique 3, 2 (jan 2002), 207-215.

[26] BORKENT, B., DAMMER, S., SCHONHERR, H., VANCSO, G., AND LOHSE,
D. Superstability of Surface Nanobubbles. Physical Review Letters 98, 20 (may
2007), 204502.

[27] BOSMA, M., CORNELISSEN, E., AND SCHWING, A. Improved Experimental


Characterisation of Cement/Rubber Zonal Isolation Materials. In SPE Asia
Pacific Oil and Gas Conference and Exhibition (apr 2013), Society of Petroleum
Engineers.

[28] BOSMA, M., CORNELISSEN, E., AND SCHWING, A. Improved Experimen-


tal Characterisation of Cement/Rubber Zonal Isolation Materials. In Interna-
tional Oil and Gas Conference and Exhibition in China (apr 2013), Society of
Petroleum Engineers.

[29] BOUZID, M., MERCURY, L., LASSIN, A., AND MATRAY, J.-M. Salt precipi-
tation and trapped liquid cavitation in micrometric capillary tubes. Journal of
colloid and interface science 360, 2 (aug 2011), 768-76.

[30] BOUZID, M., MERCURY, L., LASSIN, A., MATRAY, J.-M., AND AZAROUAL,
M. In-pore tensile stress by drying-induced capillary bridges inside porous
materials. Journal of colloid and interface science 355, 2 (mar 2011), 494-502.

[31] BOWERS, P. G., BAR-ELI, K., AND NOYES, R. M. Unstable supersaturated


solutions of gases in liquids and nucleation theory. Journal of the Chemical
Society, Faraday Transactions 92, 16 (jan 1996), 2843.

[32] BRUNAUER, S., AND GREENBERG, S. A. THE HYDRATION OF TRICAL-


CIUM SILICATE AND BETA DICALCIUM SILICATE AT ROOM TEMPER-
ATURE. Portland Cement Assoc R & D Lab Bull 1 (1970), 135-165.

[33] BULYCHEV S.I., ALEKHIN V.P., SHORSHOROV M.KH., TERNOVSKII A.P.,


SHNYREV, G. Determination of Young's modulus according to the indentation
diagram. Industrial Lab. 41 (1975), 1409-1412.

[34] BUNKIN, N. F., KISELEVA, 0. A., LOBEYEV, A. V., MOVCHAN, T. G.,


NINHAM, B. W., AND VINOGRADOVA, 0. I. Effect of Salts and Dissolved Gas
on Optical Cavitation near Hydrophobic and Hydrophilic Surfaces. Langmuir
13, 11 (may 1997), 3024-3028.

275
[35] BURGESS, D. R. NIST Standard Reference Database Number 69. Edited by
P. J. Linstrom and W. G. Mallard. National Institute of Standards and Tech-
nology.

[36] C01 COMMITTEE. Test Method for Chemical Shrinkage of Hydraulic Cement
Paste. Tech. rep., ASTM International, 2007.

[37] CARIOU, S., ULM, F.-J., AND DORMIEUX, L. Hardness-packing density scal-
ing relations for cohesive-frictional porous materials. Journal of the Mechanics
and Physics of Solids 56, 3 (mar 2008), 924-952.

[38] CARTE, A. E. Air Bubbles in Ice. Proceedings of the Physical Society 77, 3
(mar 1961), 757-768.
[39] CAUPIN, F., AND HERBERT, E. Cavitation in water: a review. Comptes
Rendus Physique 7, 9-10 (nov 2006), 1000-1017.

[40] CHIMENOS, J., FERNANDEZ, A., NADAL, R., AND ESPIELL, F. Short-term
natural weathering of MSWI bottom ash. Journal of Hazardous Materials 79,
3 (dec 2000), 287-299.

[41] CHRISTENSEN, R., AND Lo, K. Solutions for effective shear properties in
three phase sphere and cylinder models. Journal of the Mechanics and Physics
of Solids 27, 4 (aug 1979), 315-330.

[42] CHRISTENSEN, R., AND Lo, K. Erratum: Solutions for effective shear prop-
erties in three phase sphere and cylinder models. Journal of the Mechanics and
Physics of Solids 34, 6 (jan 1986), 639.

[43] CHRISTENSEN, R. M. Mechanics of Composite Materials. Courier Corporation,


2005.
[44] CONSTANTINIDES, G. Invariant Mechanical Properties of Calcium-Silicate-
Hydrates (C-S-H) in Cement-Based Materials: Instrumented Nanoindentation
and MicromechanicalModeling,. PhD thesis, Massatchusetts Institute of Tech-
nology, 2006.

[45] CONSTANTINIDES, G., RAVI CHANDRAN, K., ULM, F.-J., AND VAN VLIET,
K. Grid indentation analysis of composite microstructure and mechanics: Prin-
ciples and validation. Materials Science and Engineering: A 430, 1-2 (2006),
189-202.

[46] CONSTANTINIDES, G., AND ULM, F.-J. The effect of two types of C-S-H
on the elasticity of cement-based materials: Results from nanoindentation and
micromechanical modeling. Cement and Concrete Research 34, 1 (jan 2004),
67-80.

[47] CONSTANTINIDES, G., AND ULM, F.-J. The nanogranular nature of C-S-H.
Journal of the Mechanics and Physics of Solids 55, 1 (jan 2007), 64-90.

276
[48] CONSTANTINIDES, G., ULM,F.-J., AND VAN -VLIET, K. J. On the use of
nanoindentation for cementitious materials. Materials and Structures 36, 257
(apr 2003), 191-196.

[49] CONTRIBUTORS, W. Le Chatelier's principle, 2015.

[50] CORRENS, C. W. Growth and dissolution of crystals under linear pressure.


Discussions of the FaradaySociety 5 (jan 1949), 267.

[51] CORRENS, C. W., AND STEINBORN, W. Experimente zur Messung und


Erklirung der sogenannten Kristallisationskraft. Zeitschrift fur Kristallogra-
phie - Crystalline Materials 101, 1 (jan 1939).

[52] COUSSY, 0. Poromechanics. John Wiley & Sons, Ltd, 2004.

[53] Coussy, 0. Mechanics and physics of porous solids. John Wiley & Sons, Ltd,
2010.

[54] CRUM, L. A. Nucleation and stabilization of microbubbles in liquids. Applied


Scientific Research 38, 1 (1982), 101-115.

[55] DA SILVA, J. C., TRTIK, P., DIAZ, A., HOLLER, M., GUIZAR-SICAIROS, M.,
RAABE, J., BUNK, 0., AND MENZEL, A. Mass density and water content of
saturated never-dried calcium silicate hydrates. Langmuir : the A CS journal of
surfaces and colloids 31, 13 (apr 2015), 3779-83.

[56] D'ALOIA, L., AND CHANVILLARD, G. Determining the "apparent" activa-


tion energy of concrete: Ea-numerical simulations of the heat of hydration of
cement. Cement and Concrete Research 32, 8 (aug 2002), 1277-1289.

[57] DAVID L. KANTRO, STEPHEN BRUNAUER, AND CHARLES H. WEISE. Devel-


opment of Surface in the Hydration of Calcium Silicates, vol. 33 of Advances in
Chemistry. American Chemical Society, Washington, D. C., jun 1961.

[58] DE GENNES, P.-G., BROCHARD-WYART, F., AND QUERE, D. Capillarity


and Wetting Phenomena - Drops, Bubbles, Pearls, Waves, 2004.

[59] DELAGRAVE, A., PRINCE, J. C., MARCHAND, J., TOMA, G., PIGEON, M.,
AND BISSONNETTE, B. Equipment for the analysis of the behaviour of concrete
under restrained shrinkage at early ages. Magazine of Concrete Research 52, 4
(jan 2000), 297-302.

[60] DELVILLE, A., AND PELLENQ, R. J.-M. Electrostatic Attraction and/or


Repulsion Between Charged Colloids: A (NVT) Monte-Carlo Study. Molecular
Simulation 24, 1-3 (aug 2000), 1-24.

[61] DORMIEUX, L., KONDO, D., AND ULM, F.-J. Microporomechanics. John
Wiley & Sons, 2006.

277
[62] DUSSAN, E. B. On the Spreading of Liquids on Solid Surfaces: Static and
Dynamic Contact Lines. Annual Review of Fluid Mechanics 11, 1 (jan 1979),
371-400.

[63] EARLE WAGHORNE, W. Viscosities of electrolyte solutions. Philosophical


Transactions of the Royal Society A: Mathematical, Physical and Engineering
Sciences 359, 1785 (aug 2001), 1529-1543.

[64] EDWARDS, T. J., MAURER, G., NEWMAN, J., AND PRAUSNITZ, J. M.


Vapor-liquid equilibria in multicomponent aqueous solutions of volatile weak
electrolytes. AIChE Journal 24, 6 (1978), 966-976.

[65] FAGERLUND, G. Self-desiccation and its importance in concrete technology


proceedings of the Second International Research Seminar in Lund, June 18,
1999. Lund, 1999.

[66] FELDMAN, R. Sorption and length-change scanning isotherms of methanol and


water on hydrated Portland cement.

[67] FELDMAN, R. Helium flow and density measurement of the hydrated tricalcium
silicate - water system. Cement and Concrete Research 2, 1 (jan 1972), 123-136.

[68] FELDMAN, R. Diffusion measurements in cement paste by water replacement


using Propan-2-OL. Cement and Concrete Research 17, 4 (jul 1987), 602-612.

[69] FELDMAN, R. F., AND SEREDA, P. J. A model for hydrated Portland ce-
ment paste as deduced from sorption-length change and mechanical properties.
Materials and Structures 1, 6 (nov 1968), 509-520.

[70] FELDMAN, R. F., AND SEREDA, P. J. A New Model for Hydrated Portland
Cement and its Practical Implications. Engineering Journal 53, 8/9 (1970),
53-59.

[71] FENG, X., GARBOCZI, E., BENTZ, D., STUTZMAN, P., AND MASON, T.
Estimation of the degree of hydration of blended cement pastes by a scanning
electron microscope point-counting procedure. Cement and Concrete Research
34, 10 (oct 2004), 1787-1793.

[72] FERRARIS, C. F., AND WITTMANN, F. H. Shrinkage mechanisms of hardened


cement paste. Cement and Concrete Research 17 (1987), 453-464.

[73] FITZGERALD, S. A., NEUMANN, D. A., RUSH, J. J., BENTZ, D. P., AND
LIVINGSTON, R. A. In Situ Quasi-elastic Neutron Scattering Study of the
Hydration of Tricalcium Silicate. Chemistry of Materials 10, 1 (jan 1998), 397-
402.

[74] FLATT, R. J., CARUSO, F., SANCHEZ, A. M. A., AND SCHERER, G. W.


Chemo-mechanics of salt damage in stone. Nature communications 5 (jan 2014),
4823.

278
[75] FLATT, R. J., STEIGER, M., AND SCHERER, G. W. A commented translation
of the paper by C.W. Correns and W. Steinborn on crystallization pressure.
Environmental Geology 52, 2 (dec 2006), 187-203.

[76] FRENKEL, Y. I. Kinetic theory of liquids.

[77] FROMENT, G., BISCHOFF, K., AND WILDE, J. D. Chemical reactoranalysis


and design. 1990.

[78] Fuin, K., AND KONDO, W. Heterogeneous equilibrium of calcium silicate


hydrate in water at 30C. Journal of the Chemical Society, Dalton Transactions,
2 (jan 1981), 645.

[79] GALIN, L. A., Moss, H., AND SNEDDON, I. N. Contact problems in the
theory of elasticity.

[80] GATHIER, B. Multiscale strength homogenization - application to shale nanoin-


dentation. PhD thesis, Massatchusetts Institute of Technology, 2008.

[81] GAWIN, D., AND SANAVIA, L. Simulation of Cavitation in Water Saturated


Porous Media Considering Effects of Dissolved Air. Transport in Porous Media
81, 1 (apr 2009), 141-160.

[82] GEIKER, M., AND KNUDSEN, T. Chemical shrinkage of portland cement


pastes. Cement and Concrete Research 12, 5 (sep 1982), 603-610.

[83] GLASSER, F. P., LACHOWSKI, E. E., AND MACPHEE, D. E. Com-


positional Model for Calcium Silicate Hydrate (C-S-H) Gels, Their Solubilities,
and Free Energies of Formation. Journal of the American Ceramic Society 70,
7 (jul 1987), 481-485.

[84] HAMMER, T. Measurement Methods for Testing of Early Age Autogenous


Strains. Proceedings of the RILEM International Conference on Early Age
Cracking in Cementitious Sytems (2002).

[85] HANSEN, T. C. Physical structure of hardened cement paste. A classical ap-


proach. Materials and Structures 19, 6 (nov 1986), 423-436.

[86] HEATHMAN, J., AND RAVI, K. Methods and compositions for compensating
for cement hydration volume reduction, nov 2004.

[87] HELLMICH, C., ULM, F.-J., AND MANG, H. A. Consistent linearization


in Finite Element analysis of coupled chemo-thermal problems with exo- or
endothermal reactions. Computational Mechanics 24, 4 (oct 1999), 238-244.

[88] HENDERSON, S. J., AND SPEEDY, R. J. A Berthelot-Bourdon tube method


for studying water under tension. Journal of Physics E: Scientific Instruments
13, 7 (jul 1980), 778-782.

279
[89] HERBERT, E., BALIBAR, S., AND CAUPIN, F. Cavitation pressure in water.
Physical Review E 74, 4 (oct 2006), 041603.

[90] HESSE, C., GOETz-NEUNHOEFFER, F., AND NEUBAUER, J. A new approach


in quantitative in-situ XRD of cement pastes: Correlation of heat flow curves
with early hydration reactions. Cement and Concrete Research 41, 1 (jan 2011),
123-128.

[91] HEWLETT, P. Lea's Chemistry of Cement and Concrete, vol. 2003.


Butterworth-Heinemann, 2003.

[92] HOLT, E. Contribution of mixture design to chemical and autogenous shrinkage


of concrete at early ages. Cement and Concrete Research 35, 3 (mar 2005), 464-
472.

[93] Hu, Y.-F. Prediction of viscosity of mixed electrolyte solutions based on the
Eyring's absolute rate theory and the equations of Patwardhan and Kumar.
Chemical Engineering Science 59, 12 (jun 2004), 2457-2464.

[94] HUA, C., ACKER, P., AND EHRLACHER, A. Analyses and models of the
autogenous shrinkage of hardening cement paste: I. Modelling at macroscopic
scale. Cement and Concrete Research 25, 7 (oct 1995), 1457-1468.

[95] IGARASHI, S.-i., BENTUR, A., AND KOVLER, K. Autogenous shrinkage and
induced restraining stresses in high-strength concretes. Cement and Concrete
Research 30, 11 (nov 2000), 1701-1707.

[96] IOANNIDOU, K., PELLENQ, R. J.-M., AND DEL GADO, E. Controlling local
packing and growth in calcium-silicate-hydrate gels. Soft Mvatter 10, 8 (jan
2014), 1121.

[97] ISRAELACHVILI, J. N. Intermolecular and Surface Forces. Academic Press,


2011.

[98] JENNINGS, H. The Colloid/Nanogranular Nature of Cement Paste and Prop-


erties. In Nanotechnology in Construction 3 (2009).

[99] JENNINGS, H. M. A model for the microstructure of calcium silicate hydrate


in cement paste. Cement and Concrete Research 30, 1 (jan 2000), 101-116.

[100] JENNINGS, H. M., AND TENNIS, P. D. Model for the Developing Microstruc-
ture in Portland Cement Pastes. Journal of the American Ceramic Society 77,
12 (dec 1994), 3161-3172.

[101] JENSEN, 0. M., AND HANSEN, P. F. Water-entrained cement-based materials:


I. Principles and theoretical background. Cement and Concrete Research 31, 4
(apr 2001), 647-654.

280
[102] JIANG, F., SPEZIALE, S., AND DUFFY, T. S. Single-crystal elasticity of
brucite, Mg(OH)2, to 15 GPa by Brillouin scattering. American Mineralogist
91, 11-12 (nov 2006), 1893-1900.

[103] JONES, G., AND DOLE, M. The viscosity of aqueous solutions of strong elec-
trolytes with special reference to barium chloride. Journal of the American
Chemical Society 51, 10 (oct 1929), 2950-2964.

[104] JONES, R., ANDO, W., AND J., C. Silicon-ContainingPolymers: The Science
and Technology of Their Synthesis and Applications, vol. 30. Springer Science
& Business Media, 2000.

[105] JONES, S. Bubble nucleation from gas cavities - a review. Advances in Colloid
and Interface Science 80, 1 (feb 1999), 27-50.

[106] JONSSON, B., WENNERSTR6M, H., NONAT, A., AND CABANE, B. Onset of
cohesion in cement paste. Langmuir : the A CS journal of surfaces and colloids
20, 16 (aug 2004), 6702-9.

[107] JUSTNES, H., AND SELLEVOLD, E. The influence of cement characteristics on


chemical shrinkage. ... of Concrete, E&FN ... (1999).

[108] KIM, B. G., LEE, K. M., AND LEE, H. K. Autogenous shrinkage of high-
performance concrete containing fly ash. Magazine of Concrete Research 55, 6
(jan 2003), 507-515.

[109] KJELLSEN, K. 0., AND JUSTNES, H. Revisiting the microstructure of hydrated


tricalcium silicate-a comparison to Portland cement. Cement and Concrete
Composites 26, 8 (nov 2004), 947-956.

[110] KOFINA, A. N., AND KoUTSOUKOS, P. G. Spontaneous Precipitation of


Struvite from Synthetic Wastewater Solutions. Crystal Growth & Design 5, 2
(mar 2005), 489-496.

[111] KOHNO, K., OKAMOTO, T., ISIKAWA, Y., SIBATA, T., AND MORI, H.
Effects of artificial lightweight aggregate on autogenous shrinkage of concrete.
Cement and Concrete Research 29, 4 (apr 1999), 611-614.

[112] KUMAR, A., SANT, G., PATAPY, C., GIANOCCA, C., AND SCRIVENER,
K. L. The influence of sodium and potassium hydroxide on alite hydration:
Experiments and simulations. Cement and Concrete Research 42, 11 (nov 2012),
1513-1523.

[113] LIANG, Y., HILAL, N., LANGSTON, P., AND STAROV, V. Interaction forces
between colloidal particles in liquid: theory and experiment. Advances in colloid
and interface science 134-135 (oct 2007), 151-66.

281
[114] LOTHENBACH, B., MATSCHEI,T., M6SCHNER, G., AND GLASSER, F. P.
Thermodynamic modelling of the effect of temperature on the hydration and
porosity of Portland cement. Cement and Concrete Research 38, 1 (2008), 1-18.

[115] LoUKILI, A., CHOPIN,D., KHELIDJ, A., AND LE Touzo, J. Y. New ap-
proach to determine autogenous shrinkage of mortar at an early age considering
temperature history. Cement and Concrete Research 30, 6 (jun 2000), 915-922.

[116] LUBETKIN, S. D. Why Is It Much Easier To Nucleate Gas Bubbles than Theory
Predicts. Langmuir 19, 7 (apr 2003), 2575-2587.

[117] LURA, P., AND JENSEN, 0. M. Measuring techniques for autogenous strain
of cement paste. Materials and Structures 40, 4 (jul 2006), 431-440.

[118] MAENO, N. Air Bubble Formation in Ice Crystals, 1967.

[119] MARIS, H., AND BALIBAR, S. Negative Pressures and Cavitation in Liquid
Helium. Physics Today 53, 2 (feb 2000), 29-34.

[120] MARUYAMA, I. Origin of Drying Shrinkage of Hardened Cement Paste: Hy-


dration Pressure. Journal of Advanced Concrete Technology 8, 2 (jul 2010),
187-200.

[121] MASOERO, E., DEL GADO, E., PELLENQ, R. J. M., ULM, F. J., AND YIP,
S. Nanostructure and nanomechanics of cement: Polydisperse colloidal packing.
Physical Review Letters 109, October (oct 2012), 3-6.

[122] MASOERO, E., DEL GADO, E., PELLENQ, R. J.-M., Yip, S., AND ULM,
F.-J. Nano-scale mechanics of colloidal C-S-H gels. Soft matter 10, 3 (jan
2014), 491-9.

[123] MCGLASHAN, M. L. Manual of symbols and terminology for physicochemical


quantities and units. Pure and Applied Chemistry 21, 1 (jan 1970), 1-44.

[124] MCNAUGHT, A., AND WILKINSON, A. IUPAC Compendium of chemical ter-


minology. Blackwell Scientific Publications, Oxford, UK, 2000.

[125] MECHTCHERINE, V., AND REINHARDT, H. Application of super absorbent


polymers (SAP) in concrete construction: state-of-the-art report prepared by
Technical Committee 225-SAP. Springer Netherlands, Dordrecht, 2012.

[126] MEJLHEDE JENSEN, 0., AND FREIESLEBEN HANSEN, P. A dilatometer for


measuring autogenous deformation in hardening portland cement paste. Mate-
rials and Structures 28, 7 (aug 1995), 406-409.

[127] MIAO, B. A new method to measure the early-age deformation of cement based
materials. Proceedings of the InternationalRILEM Workshop on Shrinkage of
Concrete (2000).

282
[128] MIKHAIL, R. S., COPELAND, L. E., AND BRUNAUER, S. Pore structures
and surface areas of hardened portland cement pastes by nitrogen adsorption.
Canadian Journal of Chemistry 42, 2 (feb 1964), 426-438.

[129] MINDESS, S., YOUNG, J. F., AND DARWIN, D. Concrete, 2nd ed. Prentice
Hall, Upper Saddle River, NJ, 2003.

[130] MOHR, B., AND HOOD, K. Influence of bleed water reabsorption on cement
paste autogenous deformation. Cement and Concrete Research 40, 2 (feb 2010),
220-225.

[131] MORCH, K. A. Reflections on cavitation nuclei in water. Physics of Fluids 19,


7 (jul 2007), 072104.

[132] MOUNANGA, P., BAROGHEL-BOUNY, V., LoUKILI, A., AND KHELIDJ, A.


Autogenous deformations of cement pastes: Part I. Temperature effects at early
age and micro-macro correlations. Cement and Concrete Research 36, 1 (jan
2006), 110-122.

[133] MULLER, A. C. A., SCRIVENER, K. L., GAJEWICZ, A. M., AND MCDON-


ALD, P. J. Densification of C-S-H Measured by 1 H NMR Relaxometry. The
Journal of Physical Chemistry C 117, 1 (jan 2013), 403-412.

[134] MLLER, U., AND RfUBNER, K. The microstructure of concrete made with
municipal waste incinerator bottom ash as an aggregate component. Cement
and Concrete Research 36, 8 (aug 2006), 1434-1443.

[135] NAJAFABADI, R., AND YIP, S. Observation of finite-temperature bain trans-


formation (f.c.c. -+r b.c.c.) in Monte Carlo simulation of iron. Scripta Metal-
lurgica 17, 10 (oct 1983), 1199-1204.

[136] NELSON, E. B. Well Cementing. Newnes, 1990.

[137] NMAI, C., TOMITA, R., HONDO, F., AND BUFFENBARGER, J. Shrinkage-
Reducing Admixtures. Concrete International 20, 4 (apr 1998), 31-37.

[138] OLIVER, W., AND PHARR, G. An improved technique for determining hard-
ness and elastic modulus using load and displacement sensing indentation ex-
periments. Journal of Materials Research 7, 06 (jan 1992), 1564-1583.

[139] ONODA, G., AND LINIGER, E. Random loose packings of uniform spheres and
the dilatancy onset. Physical review letters 64, 22 (may 1990), 2727-2730.

[140] ONUKI, A. Surface tension of electrolytes: hydrophilic and hydrophobic ions


near an interface. The Journal of chemical physics 128, 22 (jun 2008), 224704.

[141] OR, D., AND TULLER, M. Cavitation during desaturation of porous media
under tension. Water Resources Research 38, 5 (may 2002), 19-1-19-14.

283
[142] PARROT, L., AND KILLOH, D. Prediction of cement hydration. Proc. Br.
Ceram. Soc. (1984).

[143] PARROTT, L. Measurement and Modeling of Porosity in Drying Cement Paste.


MRS Proceedings 85 (feb 1987), 91.

[144] PARROTT, L. L. J., GEIKER, M. M., GUTTERIDGE, W. A., AND KJLLOH,


D. Monitoring Portland cement hydration: Comparison of methods. Cement
and Concrete Research 20, 6 (nov 1990), 919-926.

[145] PATWARDHAN, V. S., AND KUMAR, A. A unified approach for prediction of


thermodynamic properties of aqueous mixed-electrolyte solutions. Part I: Vapor
pressure and heat of vaporization. AIChE Journal 32, 9 (sep 1986), 1419-1428.

[146] PAULINI, P. A weighing method for cement hydration. 9th Int. Cong. on the
Chemistry of cement, New Delhi (1992).

[147] PECQUEUR, G., CRIGNON, C., AND QUENEE, B. Waste Materials in Con-
struction Wascon 2000 - Proceedings of the International Conference on the
Science and Engineering of Recycling for Environmental Protection, Harrogate,
England 31 May, 1-2 June 2000, vol. 1 of Waste Management Series. Elsevier,
2000.

[148] PELLENQ, R. J.-M., KUSHIMA, A., SHAHSAVARI, R., VAN VLIET, K. J.,
BUEHLER, M. J., YIP, S., AND ULM, F.-J. A realistic molecular model of
cement hydrates. Proceedings of the National Academy of Sciences of the United
States of America 106, 38 (sep 2009), 16102-7.

[149] PELLENQ, R. J.-M., AND VAN DAMME, H. Why Does Concrete Set?: The
Nature of Cohesion Forces in Hardened Cement-Based Materials. MRS Bulletin
29, 05 (2004), 319-323.
[150] PENKO, M. Some early hydration processes in cement paste as monitored by
liquid phase composition measurements. PhD thesis, Purdue University, 1983.

[151] PERA, J., COUTAZ, L., AMBROISE, J., AND CHABABBET, M. Use of incin-
erator bottom ash in concrete. Cement and Concrete Research 27, 1 (jan 1997),
1-5.

[152] PETTERSEN, M., BALIBAR, S., AND MARIS, H. Experimental investigation


of cavitation in superfluid 4He.

[153] PICHLER, C., LACKNER, R., AND MANG, H. A. A multiscale micromechan-


ics model for the autogenous-shrinkage deformation of early-age cement-based
materials. Engineering Fracture Mechanics 74, 1-2 (jan 2007), 34-58.

[154] PLASSARD, C., LESNIEWSKA, E., POCHARD, I., AND NONAT, A. Nanoscale
experimental investigation of particle interactions at the origin of the cohesion
of cement. Langmuir : the ACS journal of surfaces and colloids 21, 16 (aug
2005), 7263-70.

284
[155] POWERS, T. Absorption of Water by Portland Cement Paste during the Hard-
ening Process. Ind. Eng. Chem. 27, 7 (1935), 790-794.

[156] POWERS, T. The properties of fresh concrete.

[157] POWERS, T., AND BROWNYARD, T. Studies of the physical properties of


hardened Portland cement paste. A CI Journal Proceedings 18, 2 (1947).

[158] POWERS, T. C. A discussion of cement hydration in relation to the curing of


concrete. Highway Research Board Proceedings 27, Bulletin No. 25, Portland
Cement Assn. (1947), 178.

[159] POWERS, T. C. Structure and Physical Properties of Hardened Portland Ce-


ment Paste. Journal of the American Ceramic Society 41, 1 (1958), 1-6.

[160] POWERS, T. C. Physical properties of cement paste. In Portland Cement


Assoc R & D Lab Bull (Washington, D. C. 1960, 1962), U. S. National Bureau
of Standards Monograph, pp. 577-613.

[161] RAJABIPOUR, F., SANT, G., AND WEISS, J. Interactions between shrinkage
reducing admixtures (SRA) and cement paste's pore solution. Cement and
Concrete Research 38, 5 (may 2008), 606-615.

[162] RAMACHANDRAN, V., AND BEAUDOIN, J. Handbook of Analytical Techniques


in Concrete Science and Technology. Elsevier, 2001.

[163] RELIS, M., AND SOROKA, I. Limiting values for density, expansion and in-
trinsic shrinkage in hydrated Portland cement. Cement and Concrete Research
10, 4 (jul 1980), 499-508.

[164] RICHARDSON, I. The nature of C-S-H in hardened cements. Cement and


Concrete Research 29, 8 (aug 1999), 1131-1147.

[165] RICHARDSON, I. Tobermorite/jennite- and tobermorite/calcium hydroxide-


based models for the structure of C-S-H: applicability to hardened pastes of
tricalcium silicate, f-dicalcium silicate, Portland cement, and blends of Portland
cement with blast-furnace slag, metakaol. Cement and Concrete Research 34,
9 (sep 2004), 1733-1777.

[166] RICHARDSON, I., AND GROVES, G. Models for the composition and structure
of calcium silicate hydrate (C-S-H) gel in hardened tricalcium silicate pastes.
Cement and Concrete Research 22, 6 (nov 1992), 1001-1010.

[167] ROTHSTEIN, D., AND THOMAS, J. Solubility behavior of Ca-, S-, Al-, and Si-
bearing solid phases in Portland cement pore solutions as a function of hydration
time. Cement and Concrete ... 32, 10 (oct 2002), 1663-1671.

[168] SANAHUJA, J., DORMIEUX, L., AND CHANVILLARD, G. Modelling elasticity


of a hydrating cement paste. Cement and Concrete Research 37, 10 (oct 2007),
1427-1439.

285
[169] SANT, G., LURA, P., AND WEISS, J. Measurement of volume change in
cementitious materials at early ages: Review of testing protocols and interpre-
tation of results. TransportationResearch Record: Journal of the Transportation
Research Board 1979, -1 (2006), 21-29.

[170] SCHERER, G. W. Crystallization in pores. Cement and Concrete Research 29,


8 (aug 1999), 1347-1358.

[171] SCHINDLER, A. K., AND FOLLIARD, K. J. Heat of hydration models for


cementitious materials. A CI Materials Journal 102, 1 (2005), 24-33.

[172] SELLEVOLD, E. J., VERBOVEN, F., VAN GEMERT, A., AND JUSTNES, H.
Total and external chemical shrinkage of low w/c ratio cement pastes. Advances
in Cement Research 8, 31 (oct 1996), 121-126.

[173] SEVANTO, S., HOLBROOK, N. M., AND BALL, M. C. Freeze/Thaw-induced


embolism: probability of critical bubble formation depends on speed of ice
formation. Frontiers in plant science 3 (jan 2012), 107.

[174] SIERRA, R. Distribution of different forms of water in the structure of pure


slurries of C3S and Portland cement. Proc. 7th Int. Congr. Chem. Cem (1980).

[175] SLAVCHOV, R. I., AND NOVEV, J. K. Surface tension of concentrated elec-


trolyte solutions. Journal of colloid and interface science 387, 1 (dec 2012),
234-43.

[176] SNEDDON, 1. Application of integral transforms in the theory of elasticity. 1975.

[177] SOGA, N. High Temperature Elastic Properties of Polycrystalline MgO and


AI2II oural f Amria
the Ccrbmi Soc1iety (1Ul).

[178] SUTTON, D. L., AND BURKHALTER, J. F. Gas generation retarded aluminum


powder for oil field cements, jan 1986.

[179] TAPLIN, J. A method for following the hydration reaction in portland cement
paste. Australian Journal of Applied Science (1959).

[180] TAYLOR, H. A method for predicting alkali ion concentrations in cement pore
solutions. Advances in Cement Research 1, 1 (oct 1987), 5-17.

[181] TAYLOR, H. F. Modification of the Bogue calculation. Advances in Cement


Research 2, 6 (1989), 73-7.

[182] TAYLOR, H. F. W. Cement chemistry, 2nd editio ed., vol. 20. Thomas Telford
Services Ltd, 1997.

[183] TENNIS, P. D., AND JENNINGS, H. M. A model for two types of calcium
silicate hydrate in the microstructure of Portland cement pastes. Cement and
Concrete Research 30, 6 (jun 2000), 855-863.

286
[184] THOMAS, J. J., JENNINGS, H. M., AND ALLEN, A. J. Determination of
the Neutron Scattering Contrast of Hydrated Portland Cement Paste using
H20/D20 Exchange. Advanced Cement Based Materials 7, 3-4 (apr 1998),
119-122.

[185] THOMAS, J. J., JENNINGS, H. M., AND ALLEN, A. J. Relationships between


Composition and Density of Tobermorite, Jennite, and Nanoscale CaO-SiO 2-H
2 0. The Journal of Physical Chemistry C 114, 17 (may 2010), 7594-7601.

[186] THOMAS, J. J., JENNINGS, H. M., AND CHEN, J. J. Influence of Nucleation


Seeding on the Hydration Mechanisms of Tricalcium Silicate and Cement. The
Journal of Physical Chemistry C 113, 11 (mar 2009), 4327-4334.

[187] THOMAS, J. J. J., AND JENNINGS, H. M. H. A colloidal interpretation of


chemical aging of the C-S-H gel and its effects on the properties of cement paste.
Cement and Concrete Research 36, 1 (jan 2006), 30-38.

[188] THOMAS, N., AND DOUBLE, D. Calcium and silicon concentrations in solution
during the early hydration of portland cement and tricalcium silicate. Cement
and Concrete Research 11, 5-6 (sep 1981), 675-687.

[189] THOMSON, J. XXXVII.-Theoretical Considerations on the Effect of Pressure


in Lowering the Freezing Point of Water. Transactions of the Royal Society of
Edinburgh 16, 05 (jan 1849), 575-580.

[190] THOMSON, W. XII. The effect of pressure in lowering the freezing-point of water
experimentally demonstrated. The London, Edinburgh, and Dublin Philosophi-
cal Magazine and Journal of Science 37, 248 (1850), 123-127.

[191] TUCKER, A. S., AND WARD, C. A. Critical state of bubbles in liquid-gas


solutions. Journal of Applied Physics 46, 11 (sep 1975), 4801.

[192] ULM, F., AND COUSSY, 0. Mechanics and durability of solids. 2003.

[193] ULM, F.-J. Microporomechanics for early-age behavior of hydrating cement.


Tech. rep., Massachusetts Institute of Technology, Cambridge, MA, 2012.

[194] ULM, F.-J., ABUHAIKAL, M., AND BASEL, K. Xcem report 2012, Chap-
ter 6, Microporomechanics modelling for early-age oilwell material testing and
borehole applications. Tech. rep., Massachusetts Institute of Technology, Cam-
bridge, MA, 2012.

[195] ULM, F.-J., ABUHAIKAL, M., PETERSEN, T., AND PELLENQ, R. Poro-
chemo-fracture-mechanics... bottom-up: application to risk of fracture design
of oil and gas cement sheath at early ages. CRC Press (2014), 61-70.

[196] ULM, F. J., CONSTANTINIDES, G., AND HEUKAMP, F. H. Is concrete a


poromechanics materials?-A multiscale investigation of poroelastic properties.
Materials and Structures 37, 1 (jan 2004), 43-58.

287
[197] ULM, F.-J., AND COUSSY, 0. Modeling of Thermochemomechanical Cou-
plings of Concrete at Early Ages. Journal of Engineering Mechanics 121, 7 (jul
1995), 785-794.

[198] ULM, F.-J., AND COUSSY, 0. Strength Growth as Chemo-Plastic Hardening


in Early Age Concrete. Journal of Engineering Mechanics 122, 12 (dec 1996),
1123-1132.

[199] ULM, F.-J., AND COUSSY, 0. What Is a "Massive" Concrete Structure at


Early Ages? Some Dimensional Arguments. Journal of Engineering Mechanics
127, 5 (may 2001), 512-522.

[200] ULM, F. J., VANDAMME, M., BOBKO, C., ALBERTO ORTEGA, J., TAi, K.,
AND ORTIZ, C. Statistical indentation techniques for hydrated nanocomposites:
concrete, bone, and shale. Journal of the American Ceramic Society 90, 9
(2007), 2677-2692.

[201] VANDAMME, M. The NanogranularOrigin of Concrete Creep: A Nanoinden-


tation Investigation of Microstructure and Fundamental Propertiesof Calcium-
Silicate-Hydrates. PhD thesis, Massachusetts Institute of Technology, 2008.

[202] VANDAMME, M., AND ULM, F.-J. Nanogranular origin of concrete creep.
Proceedingsof the NationalAcademy of Sciences of the United States of America
106, 26 (jun 2009), 10552-10557.

[203] WANG, P., AND ANDERKO, A. Computation of dielectric constants of solvent


mixtures and electrolyte solutions. Fluid Phase Equilibria 186, 1-2 (aug 2001),
103-122.

[204] WARD, C. A., BALAKRISHNAN, A., AND HOOPER, F. C. On the Ther-


modynamics of Nucleation in Weak Gas-Liquid Solutions. Journal of Basic
Engineering 92, 4 (dec 1970), 695.

[205] WEBER, S., AND REINHARDT, H. W. A New Generation of High Perfor-


mance Concrete: Concrete with Autogenous Curing. Advanced Cement Based
Materials 6, 2 (aug 1997), 59-68.

[206] WEYL, P. K. Pressure solution and the force of crystallization: a phenomeno-


logical theory. Journal of Geophysical Research 64, 11 (nov 1959), 2001-2025.

[207] WILCOX, W. R., AND Kuo, V. H. Gas bubble nucleation during crystalliza-
tion. Journal of Crystal Growth 19, 4 (jul 1973), 221-228.

[208] WITTMANN, F. Surface tension skrinkage and strength of hardened cement


paste. Matiriaux et Constructions 1, 6 (nov 1968), 547-552.

[209] WITTMANN, F. Interaction of hardened cement paste and water. Journal of


the American Ceramic Society 56, 6 (1973), 409-415.

288
[210] WITTMANN, F. Heresies on shrinkage and creep mechanisms. ... 8th Interna-
tional Conference on Creep, Shrinkage ... 8 (2008), 3-9.

[211] WOODS, H., STEINOUR, H. H., AND STARKE, H. R. Effect of Composition of


Portland Cement on Heat Evolved during hardening. Industrial& Engineering
Chemistry 24, 11 (nov 1932), 1207-1214.

[212] YOUNG, J. F., AND HANSEN, W. Volume Relationships for C-S-H Formation
Based on Hydration Stoichiometries. MRS Proceedings 85 (feb 1987), 313.

[213] ZHANG, J., WEISSINGER, E. A., PEETHAMPARAN, S., AND SCHERER, G. W.


Early hydration and setting of oil well cement. Cement and Concrete Research
40, 7 (jul 2010), 1023-1033.
[214] ZHANG, X. H., ZHANG, X. D., Lou, S. T., ZHANG, Z. X., SUN, J. L., AND
Hu, J. Degassing and Temperature Effects on the Formation of Nanobubbles
at the Mica-Water Interface. Langmuir 20, 9 (apr 2004), 3813-3815.

[215] ZHENG, Q., DURBEN, D. J., WOLF, G. H., AND ANGELL, C. A. Liquids at
large negative pressures: water at the homogeneous nucleation limit. Science
(New York, N. Y.) 254, 5033 (nov 1991), 829-32.

[216] ZIMA, P., MARSIK, F., AND SEDLAR, M. Cavitation rates in water with
dissolved gas and other impurities. Journal of Thermal Science 12, 2 (may
2003), 151-156.

289
290
Appendix A

Evolution of volume fractions


during hydration reactions

In this appendix, volumetric relationships are derived for the solids and liquids during
the hydration reactions of cement-based materials (CBM). Evolution of the volume
fractions are first derived for a single chemical reaction in which a reactant solid reacts
with water to form the solid hydration products.

A.1 Hydration kinetics and mechanisms

Mixing cement and water triggers a series of chemical reactions between the anhydrous
cement grains and water that lead eventually to the setting and hardening of cement-
based materials. The chemical reactions between water and cement are referred to
as the hydration reactions and the solids formed during the hydration reactions are
referred to as the hydration products [129]. Similar to other chemical reactions,
the hydration reactions of cement are characterized by the enthalpy of the reaction,
volume of the reaction, and reaction stoichiometries. This section reviews the concept
of the degree of hydration and relate the calorimetry data to the volume of the reaction
(chemical shrinkage).
Monitoring the kinetics and progression of the hydration reactions is of critical
importance for the interpretation of bulk volume change measurements and estima-

291
tion of the evolution of the mechanical properties and strength of CBM. Monitoring

the progression of the reaction is also necessary to study the impact of the different

mix variables, constituents, and curing conditions on the hydration kinetics. These

variables include the water-to-cement mass ratio (w/c), particle size distribution and

specific surface area, sample size, mixing procedures, the presence of seeds and admix-

tures, the chemical composition of cement, curing pressure, and curing temperature.

The quantity that is commonly used to describe the kinetics and progression of

the hydration reaction is the degree of hydration and the rate of hydration 9. The

degree of hydration is usually defined as the percent of mass (or amount in moles)

of the reactive part of the solid consumed in the hydration reaction. It is used to

describe as ratios how much of cement has reacted and it ranges between zero and

one [144]:
TO - m(t) _ m(t)
MO 1 - O (A. 1)

where m0 is the initial mass and m(t) is the mass of the reactive solid (cement) at

time t respectively. For a polyphase material like CBM, the degree of hydration can

be defined as the weighted sum of the degrees of hydration of all the phases:

= Z i/ (A.2)

where p, fi, and j are the density, volume fraction, and degree of hydration of phase

i, respectively; while the mass fraction mi of phase i is defined as:

m = fi and (p) = pifi (A.3)

The concept of the degree of hydration is closely related to the fractional con-

version Xi. It provides an approximate indication of the progression of the chemical

reaction. In fact, the degree of hydration is some sort of a mass average of the frac-

tional conversion of multiphase materials, where the fractional conversion is defined

as [77]:
Xi(t) = n - ni (t) (A.4)
n0

292
where Xi is the fractional conversion of reactant i, n? is the initial amount of reactant
i, and ni(t) is the amount of reactant i at time t.

The fractional conversion X and the degree of hydration are identical for a
single solid phase material when the solid is the limiting reactant. On the other hand,
there exists a quantity used by chemists to describe a concept similar to that of the
hydration degree and it is called the extent of the reaction. The extent of the reaction
a, also called the degree of advancement of the reaction, is defined as [123] [124]:

ni - no
ai i0 (A.5)
vi

where ni is the amount of substance (moles), vi is the stoichiometric number (dimen-


sionless) and a has units of moles. The extent of the reaction a will not be used in
this study and only the degree of hydration defined in eqn.A.1 and the conversion
Xi defined in Eq. (A.4) will be used in this study.

It is obvious from Eq. (A.1) that the hydration of CBM is treated as an elemen-
tary chemical reaction while it is known to be otherwise. The consequences of such
assumption depend on the use and interpretation of the degree of hydration or the
rate thereof. For example, determination of the volume fraction of C-S-H formed
during the hydration of CBM requires a direct determination of the conversion of
each of the cement minerals.

For purposes of this study, monitoring the hydration kinetics is necessary to model
the evolution of the microstructure of the material during hydration. Hence, it is
necessary to establish the relationship between the methods used to monitor the
hydration reactions and the evolution of the volume fractions of the various hydrates.

A number of methods are available to monitor the hydration kinetics of CBM.


These methods can be divided into direct and indirect methods. In the direct meth-
ods, the volume fractions of the different phases are measured directly. These meth-
ods include ThermoGravimetric analysis (TGA) [157], Quantitative X-Ray Diffraction
(QXRD) [90], Nuclear Magnetic Resonance (NMR) [133], QuasiElastic Neutron Scat-
tering (QENS) [73], and Scanning Electron Microscope (SEM) [71]. In the indirect

293
methods, the volume fractions of the various phases are inferred based on models.
This includes the enthalpy of the reaction as measured with conduction calorimetry
and the volume of the reaction (chemical shrinkage).

Calorimetry (Enthalpy of reaction (AZHrxn))

Enthalpy of reaction is the difference in enthalpy between the reactants and the prod-
ucts of a chemical reaction at specified temperature and pressure. The enthalpy of
reaction can be measured experimentally in isothermal calorimetry, or it can be calcu-
lated based on the known chemical reactions and the standard enthalpy of formation
for the reactants and the products:

AZHrxn = 13 AvpAH5 (products) - AvAZHf (reactants) (A.6)

where vp are the stoichiometric coefficients of the products from the known chemi-
cal reaction, v, are the stoichiometric coefficients of the reactants from the known
chemical reaction, LAH5 is the standard enthalpy of formation for the reactants or
the products.
Mixing of cement and water triggers an exothermic hydration reaction. This
makes iskotherm-"Al odut c-Alrieryr+- the mostq populaiIr teh ique us tf- Ynm;ni
LtL.LXLA'Q . JL U11iL.LI..k %I"i'LL U~I"Jii '"L L.1Ii1% U.LJy U11%, L.L.LJOU F e IW U%'%'JJ..iII' LLJ-%A U XIJ~_
4

tor the hydration kinetics of cement-based materials. Enthalpy of hydration is gener-


ally modeled as a linear function of the enthalpies of hydration of individual cement
phases assuming that the cement reactions proceed without interaction between the
compounds. Assuming a linear relationship between the phase composition of ce-
ment and its heat evolution, Woods et al. [211] estimated the enthalpy of reaction at
complete hydration as:
AHrxn = Z miAHi (A.7)

where AHrx, is the total heat of hydration per unit mass of cement at complete
hydration (typically measured in J/g), AHj is the enthalpy of reaction of phase i
per unit mass at complete hydration (J/g), and ni is the mass fraction of phase i,
respectively (i = C 3 S, C2 S, C 3A, C 4 AF, Gypsum, MgO, CaO, etc.). This formula

294
has been adopted by several authors and it seems to estimate the heat of reaction
with reasonable accuracy for the complete reaction (see for example [171]).
To estimate the heat of hydration for incomplete reaction, Eq. (A.7) requires the
determination of the degree of hydration of each phase individually. In other words,
the total heat of hydration cannot be related to the volume fractions of the hydration
products unless the degree of hydration of each phase is known.
The degree of hydration is typically defined as the ratio between the cumulative
enthalpy of hydration H,,, and the enthalpy of the reaction AHrx

= Hrxn
(A.8)

where the subscript H is used here to distinguish the degree of hydration as defined
in Eq. (A.1) and the degree of hydration as measured by isothermal calorimetry.

Volume of reaction (chemical shrinkage AVrx,,)

Most of volume change occurs due to decomposition of water to form hydroxyl groups
in the hydration products, like Ca(OH) 2 , Mg(OH) 2 , C-S-H, etc. Part of this volume
change is due to electrostriction of water in confined space like the interlayer water
in C-S-H or adsorbed water on the charged surface of the solids. Similar to the
enthalpy of reaction, the volume of reaction can also be used to monitor the hydration
kinetics. While enthalpy is monitored by means of isothermal calorimetry, the volume
of reaction is typically monitored by measuring the changes in the absolute volume
of a hydrating cement paste sample.
In addition to the enthalpy of reaction (AHrxn), hydration of cement is accom-
panied by an increase in solid volume (one volume of anhydrous material produces
on the average more than two volumes of hydrates) and, simultaneously, by a reduc-
tion of absolute volume due to difference in average densities between the reactants
(including water) and the products. This reduction in volume is referred to as the
volume of reaction (AVrn), also known as chemical shrinkage or Le Chatelier con-
traction. The volume of reaction (chemical shrinkage) has been directly quantified by

295
covering cement slurry with water (curing water) and measuring the volume change

of the entire system [155]. Although the fundamental technique has not changed over

time, other researchers have measured the change in the pressure at the bottom of a

burette placed on the top of the cement container [213]. Another method consists of

measuring the change in the buoyancy force on an immersed specimen in direct con-

tact with water [169]. A standard test method is also available for the measurement

of chemical shrinkage; ASTM COI [36].


The volume of reaction is another property of the chemical reaction similar to the

enthalpy of the reaction. Consequently, it can be used to monitor the reaction kinetics.

Unlike heat of hydration and due to hardware limitations, the specimen size required

for the chemical shrinkage measurement is much larger than that required for the

measurement of the enthalpy of reaction. The large size of the specimen introduces

new complications to the measurement. Heat capacity and thermal conductivity can

interfere with the enthalpy measurement but the small size of the specimen will make

this interference negligible. On the other hand, compressibility of the cement paste

constituents in conjunction with its permeability will influence the chemical shrinkage

measurement and the extent of this influence will depend on the specimen size.

Similar to the enthalpy of reaction, we define the volume of reaction for multiple

simultaneous chemical reactions as

AVrxn = miAVi (A.9)

And the degree of hydration as

= AVrxn
A=rxn (A.1O)

where the subscript V is used here to distinguish the degree of hydration as defined

by Eq. (A. 1) and the degree of hydration as measured by the volume of the reaction.

The volume of the reaction (Chemical shrinkage) can cause changes in the reaction

environment, which in turn can activate a number of bulk volume change mechanisms.

Despite this, there is no direct relationship between the chemical shrinkage and bulk

volume change and it is only the bulk volume change that can lead to development

296
of stresses in cement. Development of stresses in CBM at early ages is a consequence

of a combination of curing conditions, evolution of mechanical properties, types and

degree of constrains and the volume changes. The first step in the attempt to estimate

the evolution of stresses and straints in hydrating cement-based materials (H-CBM)

is to understand the mechanisms of volume changes.

A.2 Evolution of volume fractions

To derive the evolution of volume fractions for a single reaction, we consider a single

reactive solid (RS) reacting with water (RW) to form the hydration product (HP).

VrsRS(s) + VrwRW() 7- vhpHP(s)

where vrs,vrwand vhp are the stoichiometric coefficients of the reactive solid, water

consumed in the reaction, and the solid hydration products, respectively. A compila-

tion of possible chemical reactions in cement based materials are shown in Table 10.2

with material properties in Table A.2 and the cement chemist notation in Table A.1

by. We define the fractional conversion of substance i (degree of hydration) by:

ni0 - ni(t) _ m? - mrc(t) foc - fTc frs (A.11)

where no is the initial amount of reactant i at time t = 0, ni(t) is the amount of

reactant i at time t, mrc(t) is the mass of reactant consumed in the reaction at time

t, and m% is the initial mass of the reactant. fr, is the volume fraction of the reactive

solids consumed in the reaction, fre is the initial volume fraction of the reactive solids,
and frc is the volume fraction of the residual un-reacted solid.

We will assume that RS is the limiting reactant in this reaction and it will be used as

the reference to calculate the fractional conversion. In terms of measured quantities

in this study, the degree of hydration can be approximated as:

Vrxn(t) _ Hrxn(t) (A.12)


AVrxn A Hrxn

297
where A~ n.AHe,, are the volume of the reaction and the enthalpy of the reaction,

respectively. For the hy dration of C3 S- AV -58pL/gcs AH, = -520k J/g.:Ss,

V%'n (t) and H,.x., (t) being the volumne of the reaction and the enthalpy of the reaction
as measured at time t. respectively.

Afrxn extra water = Afl

mixing ~ nwil
water

no
Constant
mass-2
-_Constant
mass-]
residual
cement
1-no
frc

0
0 4ma 1

Figure A-1: Evolution of volume fractions where the red lines represent the constant
volume. The entire system here has a constant mass but the absolute volume is
decreasing while the volume of the paste is considered constant (red lines). fr is
the volume fraction occupied by the anhydrous cement, powder. f;, is the volume
fraction occupied by the hydration products. f, is the volume fraction occupied by
the remaining water used for mixing, and r is the porosity of the system. The yellow
circle represents the degree of hydration when all the initial mixing water is consumied
in the reaction. It can be seen in the figure that the total volunie of the system is
decreasilng due to the negative volume of the reaction (chemical shrinkage) while the
volunie of the solids is increasing. Most importantly,. the total mass of the system is
constant (Conservation of M\ass). Note that the extra water added to maintain the
systen saturated is exactly Afr,, which is the mininum amount of water required
to maintain the system saturated. 1/ is the volume of the paste and it is assumed
constant in the derivations of volume fractions. "Constant mass-1" is the mass of the
seahed system and the reaction stops at "PWN when all the initial water is consuied
ili the reaction (water is the limiting reactant). "Constant nmass-2" is the mass of the
paste in additioi to the mnininni amount of curing fluid required to maintain the
pore pressure constant.

Now consider the reference volunie to be composed of the voluinme fraction of

I
residual reactive solids frc, hydration products fhp (in the absence of air and non-
reactive solids). Then as illustrated in Figure A-1, we have:

#+ frc + fp = 1 (A.13)

where 4 is the total porosity of the system (at all scales). For a saturated material,
the porosity is filled with the mixing water fm w and water provided from the outside,

fmw is the volume fraction of water that remained from the initial water used for
mixing. This definition is important because under constant pressure in a saturated
system, the volume fraction of water in the system will be fmw + cfrxn, where c < 1 is

a factor depending on the initial w/c. Also, in sealed specimens, if the initial amount
of water available for the reaction is small, then water can be the limiting reactant.
That is, during the hydration reaction:

fm W f< (A.14)

where f0 = no is the initial volume fraction of water in the mix at t 0. This leads
us to define the initial water-to-solid ratio:

W/C = m _ (A.15)
mrc fcpPc

where w/c, the initial water-to-solid mass ratio is a pre-specified quantity, mm is the
initial mass of the mixing water, m' is the initial mass of the reactive solids in the
mix at time t = 0, and frc is the initial volume fraction of the reactive solids in the
mix. Assuming that the initial volume is composed of water and solids only (not
considering entrapped air of non-reactive inclusions), we have:

f +f0
0 =1 (A.16)

To satisfy the conservation of mass under specific conditions including constant pres-
sure and temperature we need to consider a system composed of the mass of residual

299
cement powder M, the initial mass of cement powder M,?c, the mass of hydration
products Mhp, the mass of mixing water Mmw, the initial mass of mixing water M'MW,
and an additional mass of a fluid (water) necessary to maintain the system under con-
stant pressure, which we call the mass of reaction Mxn. Then the mass conservation
reads:

Mrc + Mhp + Mmw + Mrxn = Mo = constant (A.17)

And in terms of mass fractions (see Figure A-1):

mc + mhp + mwm + mrxn 1 (A.18)

where mi is the mass fraction of phase i. The initial water-to-cement ratio is w/c =
M". The relationship between the mass fraction and the volume fraction is:

ni = fi (A.19)

where (p) is the average density of the system, which is increasing over time. pi is
the density of phase i, and fi is the volume fraction occupied by phase i. The mass
fraction mi is the mass of phase i relative to the total mass of the system Mo,which is
ii'.L 0UUL UI VVYLILL%- ULi1V- V %JIUIIIV>k LI (AAU I kii J % 10J Uli%, V VJiLLIII%, A"UJIA. LJ1}~iA , 1~ I-" V~ ~I V V- U") LII.'

total volume V which does not correspond to the total mass mentioned previously.

Mass conservation can now be written in terms of volume fractions, where the
mass of the reactants is equal the mass of the products:

Mrs + Mrw = Mhp (A.20)

or, in terms of volume fractions:

Prs frs + Prwfrw = Phpfhp (A.21)

where the mass is constant but the volume is not (see figure A-1). frw = f m2 - fmw is

the volume fraction of water that participated in the chemical reaction; frs = f2, - frc

300
is the volume fraction of the reactive solid that participated in the chemical reaction,
and fhp is the volume fraction occupied by the hydration products. Note that to
satisfy Eq. (A.21), the volume of the system must change due to differences in the
average densities of the reactants and the products.

Considering a constant volume of the paste (V in figure A-1), during the time
interval dt the reaction proceeds with the amount d, the increment of the fractional
volume of the reaction is:

dfrxn = dfhp - dfrw - dfrs (A.22)

where frxn is not a real volume fraction occupied by a material (water), but rather the
volume of water required to maintain the system under constant pore pressure. dfrw,
dfrs are the increment of volume fractions of the water and solid that reacted to form
the incremental volume fraction of the reaction products dfhp. For a typical reaction
considered here, dfrw > Odfrxn > ,dfrs > O,and dfhp > 0. Another important ratio is

the volume ratio of the increment of reacted water to the increment of reacted solid:

RH = dfrw - VrwVrw (A.23)


dfrs VrsVrS

where VnW and Vrs are the molar volumes of the water and the reactive solids from
which we have:

fr= RHfrs; fr (A.24)

The ratio of the incremental volume of the solid products to the incremental volume
of the solid reactants is:
Rhp = dfhp _ vhp > 0 (A.25)
rs V dfrs V

where V is the molar volume of the solid hydration products, with:

fhp = Rhpfrs; 0 0 (A.26)

The values of RH and Rhp can be estimated based on the known chemical reaction

301
stoichiometries and molar volumes of the different constituents. The initial porosity

of the system can then be estimated based on the initial water-to-solid mass ratio

and the densities of the initial components

no=f"i0 = j (A.27)
Pc

From Eq (A.22), (A.23), and (A.25), we thus obtain:

dfrxn = Rhpdfrs - RHdfrs - dfrs dfrs(Rhp - RH - 1) (A.28)

dfrx
dfre = fx (A.29)
f Rhp - RH - 1

Based on the definition of the degree of hydration (A.11) and the initial porosity

(A.16), we also have:

frc= (1 - no); frs = (1 - no) (A.30)

The volume of the reaction Vrxn is typically reported as volume per unit of the initial

mass of the reactive solids, and the dimensionless fractional volume of the reaction

is:

frxn = VrxnfrcPc = ip - frs - frw (A.31)

The fractional conversion can be rewritten as:

V _rxn _frS (A.32)


AVrxn frsc

where Vrxn is measured in pL/g and at complete hydration V=/x = AVrxn. AHrxn
can be used in the same manner with AVrxn = V zAHrxn. By definition:

AVrxn - dfhp - dfrw - dfrs (A.33)


Pcdfrs

Introducing Eq. (A.23) and (A.25) into Eq. (A.33) gives

Avrxn = Rhp - RH I (A.34)


Pc

302
Finally, assuming that enough space is available for the hydration products to form

f( 1) 1 or in terms of initial porosity no > 1 -


-Rhp
with cement being the

limiting reactant, fIW(( 1) = 0 or in terms of initial porosity no > R + the

volume fractions in termns of the known and measured quantities are obtained:

I
Evolution of volume fractions in a single chemical reaction

Fractional volume of the reaction (negative)

frn =(RhJ - R - 1)(1 - no) (A.35)

Volume fraction of the residual un-reacted cemlenlt (decreasing)

fre = (1 - no)(1 (A.36)


-

Volume fraction of the remaining mixing water (decreasing)

fm = no - (1 - no)RH ;> 0 (A.37)

Volume fraction of the solid non-porous hydration products (increasing)

fil (1 - nO)Rjyp (A.38)

The total porosity of the system in the absence of bulk volume changes is

= 1 -fc - flp
(A.39)
= - (1 - no)(Rlp - 1)

In the presence of non-reactive solids with a volune fraction of Jur all quantities
must be multiplied by (1 - f,,).

303
A.2.1 Minimum water content and initial porosity for com-
plete reaction

For the reactive solids in the system to be the limiting reactant, a minimum amount
of water must be provided to the system. This water can be introduced as the
initial amount of water in the system or provided to the system through its outer
boundaries during the reaction. If the system is sealed and water exchange with the
surroundings is prohibited, then there will be a minimum amount of water for the
reaction to proceed to completion. The minimum amount of water required such that
the solids are the limiting reactants can be introduced as the minimum water-to-solid
mass ratio w/C~mi

w/C > w/cmin = RH PW (A.40)


Pc

where w/c"i is calculated when the fractional conversion as calculated from water or
solids is identical (that is, both the reactive solids and the initial amount of mixing
water will be completely consumed at complete hydration):

fmw( = 1) = 0 = no - (1 - no)RH (A.41)

or
f oc m frWfrc - (A.42)

If the system is not sealed and water is provided continuously from the surroundings,
there will be a limit on the initial porosity that would allow for complete reaction (in
the absence of bulk volume changes):

no= fhp - frc = (no - 1)(Rhp - 1) (A.43)

The minimum water-to-solid ratio necessary to provide enough space for the hydration
products to form without stress development, W/co, is:

w/c0 = (Rhp - 1) P (A.44)


Pc

304
Still, for reasonable workability, the minimum water content will be influenced by the
specific surface area of the solids and its surface properties. The fractional conversion
at which the initial volume is occupied completely by the non-porous reactive solids
and non-porous reaction products, can be calculated by setting:

frc + fAp = 1 (A.45)

The fractional conversion is then:

S< 1 (A.46)
(no - 1)(Rhp - 1)

-
A.2.2 Porosity of the individual solid phases

In the previous analysis, both the reactants and the products are assumed to be non-
porous solids with constant density. Although changes in density of the pure phases
will not be considered in this analysis, porosity of these phases can play a critical role
in the evolution of the volume fractions and the associated stresses in the system. In
the following analysis, the total porosity will be decomposed into fraction of porosity
encapsulated inside the reacting solids, p8 , fraction of porosity encapsulated inside
the reaction products, Sorp, and the remaining porosity (macro porosity), Sm, such
that the total porosity is:

S= (Prc+ Ohp + Om (A.47)

This decomposition is based on the characteristic size of the pores in the different
phases and will prove necessary for the microporomechanics modelling of the individ-
ual phases and the composite material. The volume fraction occupied by the porous
solids can then be defined in a way such that:

f h+'f, + m 1 (A.48)

305
where f7 e and fo are the volume fractions occupied by the porous reactive solids and
the porous reaction products, respectively. According to this definition, we have:

4 fr- (1 - no)(1 - (A)9


re?lrc(A.49
ire qrc

f fr_ (1 - no)Rhp (A.50)


p7rp 77hp

OM = - - (A.51)

(;rc= frc4rc = frc rc , and Tph = fhp9hp = h (A.52)


hpP fhp

where 9i and ri are the porosity and packing fraction of phase i and i+ r7 = 1. Now
the fractional conversion (hydration degree) at which the porous solids can fill the
initial volume is calculated by setting:

f4 + f hp = 1 or #m = 0 (A.53)

The degree of reaction required for this condition is:

-0 hp (IJc+ro- 1)_
4 (T hp- TjrcRhp)(no - 1) 1 (A.54)

Or, in terms of the initial rater-to-solid ratio

_ nlhpL[rc(w/cprc + Pw) - Pw] < 1 (A.55)


Co (Thp - rqrcRhp)pw

For the hydration of C 3 S (see Table A. 1 for chemist notation), the degree of hydration
at which the initial volume is filled with porous hydration products is calculated by
setting:

fGel + fCH + frc = 1 (A.56)

where fGe= fCSH , or

fmw - (fGelOGel) = 0 (A.57)

306
where fGei - fcsH is the volume fraction occupied by the porous C-S-H gel. With
??CSH

jo =0.5 and Tulim = 0.72, the degree of hydration when all the free water in the
system is inside the C-S-H gel is:

no (A.58)
(1 - no) * (RH 0.3889RcSH)

The maximum water-to-solid ratio required to fill the initial volume at complete
reaction is
= (Rhp - Thp) pw (A.59)
7hrpPc

For two simultaneous reactions, the enthalpy of the reaction and the volume of
the reaction are sufficient to calculate the evolution of the volume fractions of the
different phases in the system. To estimate the evolution of the volume fractions of
all phases in multiple chemical reactions like that in ordinary cement, we will need
a number of independent measurements equal to the number of chemical reactions.
In general, relevant measurements can be divided into two main quantities: the first
quantity is the enthalpy of the reaction as measured in isothermal calorimetry. The
second quantity is the direct and indirect measurement of the volume (or mass)
fractions of the different system constituents. The second quantity can be measured
in tests like the volume of the reaction (chemical shrinkage), TGA [8], QXRD [9],
NMR [10], QENS [11], SEM [12] etc. An example of additional measurements can be
obtained from TGA, where the mass fractions of some phases like Portlandite can be
obtained . Once a quantity like Portlandite is obtained, it can then be related to the
corresponding reactions.
Table A.1: Standard cement chemistry notation as assumed throughout this docu-
ment.

Oxide Composition Molar mass Oxide Composition Molar mass


C CaO 56.08 S SO 3 80.06
S Si0 2 60.08 M MgO 40.30
H H20 18.02 C CO 2 44.01
A A1 2 0 3 101.96 K K2 0 94.20
F Fe 2 O 3 159.69 N Na 2 0 61.98

307
Table A.2: Physical properties of cement compounds

Compound name Compound Specific Molar mass Molar3 volume AH Refs


formula gravity (g/mol) (cm /mol) (kJ/mol)
water H 1 18.015 18 -285.83 [35
Tricalcium silicate C3 S 3.21 227.91 71 -2927.82 [129]
Dicalcium silicate C2S 3.28 170.56 52 -2311.6 [129]
Tricalcium aluminate C3 A 3.03 269.97 89.1 -3587.8 [129]
Tetracalcium aluminoferrite C4 AF 3.73 477.44 128 -5090.3 [129]
Anhydrite CS 2.96 136.14 45.84 -1433
Hemihydrate CSHo.5 2.74 145.15 52.9 -1577.9 [129]
Gypsum CH 2 2.32 172.17 74.2 -2022.6 [129]
Calcium hydroxide CH 2.24 74.1 33.2 -986.1 [129]
Ettringite C6 AS 3 H 3 2 1.78 1255 715 -17539 [129,182]
Monosulfate C 4 ASH 1 2 2.02 623 313 -8778 [129]
Hydrogarnet C3 AH 6 2.52 378 150 -5540 [114]
AFm C 4 AH1 3 2.02 260 -8318 [129,182]
2.04 560.3 274 -8302 [114]
C 2 AH8 1.95 358.1 165 -5436 [4,129]
1.95 358.2 184 -5433 [114]
Iron hydroxide FH3 3 209 69.8 -823.9 [129]
Magnesium oxide M 3.58 40.3 11 -601.24 [35]
Magnesium hydroxide MH 2.37 58.3 24.2 -924.66 [35]
Calcium oxide C 3.34 56.1 16.79 -635.09 [35]
Silicon dioxide (Silica) S 2.648 60.08 22.69 -910.06 [35]
Appendix B

Elementary concepts and tools for


micro-poro-mechanics modelling

B.1 Representative volume and scale separability

It is critical to establish the concept of a representative elementary volume (REV) to

understand the behavior and average properties of multiscale heterogeneous materials.

The choice of the size of the REV is dependent upon the porosity, particle size dis-

tribution of the solids, and the scale of observation [13]. The goal of homogenization

techniques is to replace the complex heterogeneous material by a fictitious homoge-

neous material that behaves macroscopically in the same way. This implies that the

fictitious homogeneous material exhibits the same response to macroscopic mechani-

cal loading as the actual heterogeneous material. This requires that the macroscopic

stress and strain fields Ej and Eij derived at the macroscopic scale represent the

average values of microscopic heterogeneous stress and strain fields o-sj(z) and eij(z)

derived at the microscopic scale [14].

The size of the REV is chosen such that the characteristic length 1 of the REV

is much smaller than the characteristic length of the structure L to guarantee the

relevance of the continuum description [15]. At the same time, the size of the REV

should be large enough to represent the average properties of the heterogeneous ma-

terial. This requires that the size of the heterogeneity d should be much smaller than

309
the characteristic length of the REV to guarantee that the REV is representative of
the physical and geometric properties of the heterogeneous material. In the end, we
must have:
(B.1)

B.2 Stress and strain field averages

For homogeneous stress boundary conditions, the components of the surface traction
prescribed on the boundary are:
T = Eijnj (B.2)

where nj is the unit normal to the surface. Similarly, for homogeneous boundary
conditions the displacement components at the boundary are given by

ui = Eijxj (B.3)

From Eq. (B.2), and the divergence theorem one can prove that:

E7-4 = (os z)= -1 f tT' (z)dV (B.4)~


V Jv

Similarly, from Eq. (B.3) one can prove that:

Eij = (Eij (z)) =


V IV Eij(z)dV (B.5)

In multiphase materials,

Eij = (o7j(z)) = E fr( (z))' (B.6)

where f, is the volume fraction of phase r. And similarly

Eij= (Eij (Z)) =


r
I frK(-ij (Z)' (B.7)

310
These tools will be used to formulate the multiscale micro-poro-mechanics model

B.3 Statistical indentation

The application of indentation analysis to heterogeneous porous materials is made

possible by an extensive development of the grid indentation technique by Constan-

tinides and Ulm [44], Vandamme and Ulm [201], and Gathier and Ulm [80].

The aim of the indentation test is to extract the elastic modulus and hardness

of a material. An indentation test consists of driving an indenter tip with known

properties orthogonally into a flat surface of the tested material. During an inden-

tation test, both load and displacement are continuously recorded and the resulting

load-displacement curve (P-h curve) defines the characteristic mechanical properties

of the material. The main parametrs obtained from an indentation test are the inden-

tation modulus M and the indentation hardness H, where the indentation modulus

is defined as

M = S (B.8)

where S = dP/dh is the slope of the unloading branch of the P-h curve, and A,

is the projected contact area between the indenter tip and the indented material.

This definition of the indentation modulus is attributed to Bulychev, Alekhin and

Shorshorov [33]. the contact area A. can be determined indirectly in terms of the

maximum indentation depth measured in the indentation experiment. One of the

methods used to determine the contact area A. is the Oliver and Pharr method [138].

For isotropic materials, the indentation modulus can be related to the plane-stress

modulus through the following formula [176] [79]:

E
M = 2(B.9)
1 --v 2

where E is the Young's modulus, and v is the Poisson's ratio of the indented material.

The indentation hardness, on the other hand, is defined as the average stress below

311
the indenter
P
H =c (B.10)
HAc
where P is the maximum load applied by the indenter. For the case of cohesive-
frictional materials, Cariou and Ulm [37] and Gathier and Ulm [80] developed an
approach that allows the determination of cohesion and friction of a material gov-
erned by a Drucker Prager strength model. This methods establishes the relationship
between hardness H, packing density q, and the cohesion and fiction properties (CS,
a). The scaling relations have the form [80]:

H = h 8 (c', a) x HH (a, T) (B.11)

where HH(a, TI) is a dimensionless function, and h8 (c', a) = lim, 1 H is the asymp-
totic hardness of the cohesive-frictional solid governed by the Drucker-Prager strength
criterion [201]:
h8 = cS x A (1 + Ba + (Ca)3 + (Da)10 ) (B.12)

Herein, A = 4.76438, B = 2.5934, C = 2.8160, and D = 1.6777. The second


dimensionless function reads:

IH(a,n) = l1(l) + a(1 - 'q)12 (a,n) (B.13)

where H1 and 1 2 are given for a polycrystal morphology-by:

2(2T - 1) - (2n - 1)
ig)= sqt
sqrt2 - 1(1
1 + g(1 - 9j) + h(1 - g)2 + j(1 - rg) 3 ) (B.14)

2i - 1
r2(a, q) = 2 (k + m(1 - q) + p(l - q)a + qa) (B.15)

where g = -5.3678, h = 12.1933, j = -10.3071, k = 6.7374, m = -39.5893,


p = 34.3216, q = -21.2053. The previous relations are valid for porous granular
material with homogeneous distribution of porosity. For heterogeneous materials,
Ulm and coworkers [45-48,196,200] developed the grid indentation technique. The

312
grid indentation technique consists of performing a large number of indents on a flat

and statistically representative surface of a heterogeneous material. Provided that

the scale separability is satisfied, each indentation experiment represents a single

porous solid phase (constant porosity at each indent). The grid indentation provides

the porosity distribution of the heterogeneous material, the mechanical properties of

multi-phase materials as well as the corresponding volume fraction of each phase.

For a porous solid, the indentation modulus is representative of the elastic prop-

erties, which is considered constant for the rev subjected to the indentation load. In

other words, the indentation modulus depends on the elastic properties in, v. and

the porosity of the solid = (1 -):

M=HA(v., #)
(B.16)
ins

For granular material with spherical particles [44], the following holds:

lsc _ MSC _ sc (9 y7 + 4Msc + 3-y,)(3-y + 4) (B.17)


M ms 4(4MSC + 37s)(3ys + 1)

where MSC is the composite indentation modulus, -y, = 2(1 + v,)/3(1 - 2v,) > 0, and

MSC is the composite shear-to-solid shear moduli ratio:

1 5 3
Msc = I _ - 3 - (3 -#)
2 4 16
1
1 \/144(1 - y) - 480(1 - #) + 400(1 -
16
#) 2 + 4087(1 - 4) - 120y,(l - #) 2 + 972(3 - 0)2

(B.18)

The indentation hardness on the other hand depends on three parameters: the

strength properties of the granular solid (cohesion c' and friction coefficient as),

porosity = 1 - y, and the geometry of the indenter probe.

H
we= iH (B. 19)
hs

where h. is the asymptotic hardness of the frictional-cohesive granular solid that

313
obeys Drucker-Prager mlodel: h' is given by Eq. (B.12) and H1 , is given by Eq. (B.3).

Finally. to extract the packing density distribution and the asyiptotic solid prop-

erties of a single granular porous solid with heterogeneous distribution of porosity.

we apply the statistical analysis of the grid indentation. Grid indentation test on

a porous solid provides a bivariate data set with N entries of indentation moduli

M- .N and indentation hardness Hjj=1 . N. Making use of this data set, we can back

calculate the indentation mnodulis 'm, Poisson's ratio ius. indentation hardness he,

friction coefficient o,. packing deinsity at percolation 1 /0. as well as a set local packing

densities i=J ...N by nfninizing the quadratic error between model predictions and

the experimental values:

2]
- T/1gS /) 2 t h.H1/(c+ Ili, '1/0) (B.20)
IIIin 1 -- + 1
A exS
l ofn1.
s io Hi -
Oli -=1...N
A)

An examiple of such inimiization is show~ln in Figure B-1.

40

35

30

25
-- M model (G pa)
20 M -(-Gpa)-(M gO-RD-09)
M (Gpa) (M gO-RD-08)
15 M (Gpa) (M gO-RD-07)
I M (Gpa) (M gO-Pis-03)
10
/

I
I
I M (Gpa) (M gO-Pis-02)
I
5- M (Gpa) (Ng gO-Pis-01)
I

0
0 200 400 600 80 0 1000
H (MPa)

Figure B-1: Indentation noduills NI and indentation laIrdness H comnpiled fron the
diffient M'\Igo specinens shown in Table 6.1.

314

You might also like