You are on page 1of 201

EFFECT OF A NEWLY DEVELOPED

LIGNOSULPHONATE SUPERPLASTICIZER ON
PROPERTIES OF CEMENT PASTES AND
MORTARS

SUN DAO JUN

NATIONAL UNIVERSITY OF SINGAPORE

2008
EFFECT OF A NEWLY DEVELOPED
LIGNOSULPHONATE SUPERPLASTICIZER ON
PROPERTIES OF CEMENT PASTES AND
MORTARS

SUN DAO JUN


(B. Eng. (Hons.), NUS)

A THESIS SUBMITTED
FOR THE DEGREE OF MASTER OF ENGINEERING
DEPARTMENT OF CIVIL ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE
2008
Acknowledgements

The author would like to take this opportunity to express his sincere appreciation and

deep gratitude to his supervisor, Associate Professor Zhang Min-Hong for her

invaluable guidance, patience, kind encouragement and full support throughout the

entire course of this research.

The author’s heartfelt appreciation goes to Dr. Sisomphon Kritsada, Dr. Kåre Reknes

(Borregaard, Norway) and Mr. Philip Chuah (Borregaard, Singapore) for their useful

comments and constructive discussion.

This project would not have been successful without the kind assistance of the lab

technicians at Structural and Concrete Laboratory; special thanks to Mr. Ang Beng

Oon, who assisted the author to conduct some of the laboratory experiments. The

author also would like to deliver his gratefulness to his friends for their moral support;

special thanks to Mr. Lee Wah Peng and Wang Zengrong.

Sincere gratitude is extended to Borregaard Ligno Tech (Sarpsborg, Norway) for

providing the research grant.

Finally, the author dedicates this study to his dear parents who have given him fullest

support and unconditional love all these years.

i
Table of Contents

Acknowledgements .............................................................................................................. i

Table of Contents ................................................................................................................ii

Summary ...........................................................................................................................vii

List of Notations.................................................................................................................. x

List of Acronyms...............................................................................................................xii

List of Tables ....................................................................................................................xiv

List of Figures .................................................................................................................xvii

Chapter 1 Introduction ...................................................................................................... 21

1.1 Background.........................................................................................................21

1.2 Objectives ...........................................................................................................24

1.3 Scope...................................................................................................................25

Chapter 2 Literature Review ............................................................................................. 28

2.1 Nature of Water Reducing Admixtures ..............................................................28

2.1.1 Regular Water Reducing Admixtures (WRAs) ............................................29

2.1.1.1 Lignosulphonate....................................................................................29

2.1.1.2 Hydroxyl carboxylic acids and their salts.............................................30

2.1.1.3 Carbohydrates.......................................................................................31

2.1.1.4 Other compounds ..................................................................................31

2.1.2 Superplasticizers (SPs).................................................................................31

2.1.2.1 Modified lignosulphonate(MLS) ...........................................................33

2.1.2.2 Sulphonated melamine / naphthalene formaldehyde condensates

(SMF/SNF) ........................................................................................................34

2.1.2.3 Polycarboxylate based (PCE) ...............................................................34

ii
2.2 Mechanisms of Water Reduction ........................................................................36

2.2.1 Electrostatic Repulsion.................................................................................37

2.2.2 Steric Hindrance ...........................................................................................38

2.2.3 Solid-Liquid Affinity....................................................................................38

2.2.4 Mechanisms of WRA and SP of Different Natures......................................38

2.3 Portland Cement Hydration ................................................................................39

2.3.1 Chemistry of Portland Cement Hydration....................................................39

2.3.2 Heat Evolution of Portland Cement Hydration ............................................42

2.3.2.1 Measurement of heat evolution of cement hydration ............................42

2.3.2.2 Effect of the admixtures on heat evolution of cement hydration ...........42

2.4 Effect of the Admixtures on Cement Hydration .................................................43

2.4.1 Effect of LS Admixtures ..............................................................................44

2.4.2 Effect of SNF Admixtures............................................................................45

2.4.3 Effect of PCE Admixtures............................................................................47

2.5 Effect of the Admixtures on Workability............................................................48

2.5.1 Workability and Rheological Parameters .....................................................48

2.5.2 Effect of Admixtures on Initial Workability.................................................50

2.5.3 Effect of Admixtures on Workability Retention...........................................52

2.6 Effect of the Admixtures on Setting ...................................................................54

2.7 Effect of the Admixtures on Pore Structure & Strength Development...............55

2.7.1 Principle of Mercury Intrusion Porosimetry and Characterization of Pore

Structure ................................................................................................................55

2.7.1.1 Total porosity ........................................................................................56

2.7.1.2 Critical pore diameter...........................................................................56

2.7.1.3 Threshold pore diameter .......................................................................57

iii
2.7.1.4 Pore Size Distribtuion ...........................................................................57

2.7.1.5 Evaluation of MIP .................................................................................58

2.7.2 Effect of Admixtures on Pore Structure of Cement Paste ............................59

2.7.2.1 Effect of PCE Admixtures......................................................................59

2.7.2.2 Effect of SNF Admixtures ......................................................................60

2.7.2.3 Effect of LS Admixtures .........................................................................61

2.7.2.4 Comparisons of Effect of PCE, SNF and LS Admixtures......................61

2.8 Drying Techniques of Cement Paste and Testing Methods ................................62

2.8.1 Drying Techniques for Cement Paste...........................................................62

2.8.1.1 Oven drying ...........................................................................................62

2.8.1.2 D-drying ................................................................................................63

2.8.1.3 Vacuum drying ......................................................................................63

2.8.1.4 Solvent exchange ...................................................................................63

2.8.1.5 Freeze drying.........................................................................................64

2.8.2 X-Ray Diffraction (XRD) ............................................................................65

2.8.3 Thermogravimetric Analysis (TG) ...............................................................68

Chapter 3 Experimental Details ........................................................................................ 76

3.1 Introduction.........................................................................................................76

3.2 Materials .............................................................................................................76

3.2.1 Cement and Water ........................................................................................76

3.2.2 Aggregates....................................................................................................77

3.2.3 Water Reducing Admixtures ........................................................................77

3.3 Mix Proportions of Cement Pastes and Mortars.................................................79

3.4 Preparations for Cement Pastes and Mortars......................................................80

3.4.1 Preparation for Cement Pastes .....................................................................80

iv
3.4.2 Preparation for Mortars ................................................................................82

3.5 Test Methods and Analyses.................................................................................83

3.5.1 Heat Evolution of Cement Hydration ..........................................................83

3.5.2 Degree of Cement Hydration .......................................................................86

3.5.2.1 X-ray Diffraction (XRD)........................................................................86

3.5.2.2 Thermogravimety Analysis (TG) ...........................................................88

3.5.2.3 Non-Evaporable Water (NEW) Content................................................89

3.5.3 Workability Retention of Mortars ................................................................91

3.5.4 Setting Time of Mortars ...............................................................................94

3.5.5 Pore Structures of Pastes ..............................................................................94

3.5.6 Compressive Strength of Mortars ................................................................95

Chapter 4 Results and Discussion ................................................................................... 107

4.1 Heat Evolution of Cement Hydration ...............................................................107

4.2 Degree of Cement Hydration............................................................................113

4.2.1 Reduction of C3S in Cement Pastes ...........................................................113

4.2.2 Hydration Progress in the Cement Pastes ..................................................114

4.2.2.1 Calcium hydroxide (CH) in cement pastes..........................................114

4.2.2.2 Non-evaporable water in cement pastes .............................................117

4.2.3 Degree of Hydration in Cement Pastes ......................................................119

4.3 Workability Retention of Mortars with Time....................................................121

4.3.1 Change in the Yield Stress of Mortars with Time ......................................121

4.3.2 Change in Plastic Viscosity of Mortars with Time.....................................125

4.3.3 Change in Flow Value of Mortars with Time .............................................126

4.3.4 Relationship between the Yield Stress and Flow Value .............................127

4.4 Setting Times of Mortars ..................................................................................128

v
4.5 Pore Structure of Cement Pastes.......................................................................130

4.5.1 Total Porosity of Cement Pastes with Admixtures.....................................131

4.5.2 Pore Size Distribution of Cement Pastes with Admixtures........................132

4.5.3 Threshold and Critical Pore Diameters ......................................................135

4.6 Compressive Strength of Mortars .....................................................................137

Chapter 5 Conclusions and Recommendations............................................................... 163

5.1 Conclusions.......................................................................................................163

5.2 Recommendations.............................................................................................166

References ....................................................................................................................... 168

Appendix ......................................................................................................................... 181

vi
Summary

Lignosulphonate (LS) has been widely used in concrete as regular water reducers for

many decades due to its relatively low price. Significant advances have been made in

process and production of LS based admixtures. There is a wide range of

lignosulphonates available and their performance in concrete varies from regular

water reduction and strong retardation to high range water reduction. With the

development of a new modified LS superplasticizer (PLS), it is possible to produce

self-compacting concrete with such an admixture. However, there is not much

information available on the effect of the newly developed modified LS

superplasticizer on cement hydration, workability retention, and pore structure of

pastes in comparison to those of polycarboxylate, naphthalene and the other modified

LS superplasticizers and to those of regular LS water reducing admixtures. The

present research was, therefore, carried out.

Six admixtures were used which included four superplasticizers (one polycarboxylate

(PCE), one naphthalene (SNF), two modified lignosulphonates (PLS and UNA)) and

two regular water reducing admixtures (lignosulphonates (BCS and BCA)). Mortars

and cement pastes were designed to have similar workability. The dosages of

admixtures were determined to achieve an initial target yield stress of 75 ± 15 Pa for

mortars with w/c ratios of 0.34 and 0.40. This yield stress level will produce concrete

with slump ≥ 100 mm. For w/c of 0.40, all six admixtures were investigated; whereas

for w/c of 0.34, only four superplasticizers were investigated due to the difficulty in

achieving required initial yield stress by using regular water reducing admixtures.

vii
The results indicate that the water reducing admixtures and superplasticizers delayed

cement hydration for both w/c ratios at early ages, but did not have significant effect

on cement hydration at later ages from 7 to 91 days. The retardation of the pastes was

in the order of SNF < PCE < PLS < UNA < BCA < BCS.

The workability loss of the mortars with the LS superplasticizers was similar within

the first hour, but less than those with the SNF and PCE superplasticizers. The

workability loss of the mortars with the two regular water reducing admixtures was

more significant than those with the superplasticizers.

The order of setting times of mortars with admixtures agreed with the length of

induction periods in the heat curves of the respective pastes, i.e. SNF < PCE < PLS <

UNA< BCA < BCS. The admixtures had strong influence on the initial setting times.

However, once the mortars reached the initial setting, the final setting was not

significantly affected by the admixtures.

For the four superplasticizers, better workability retention corresponded to longer

setting time. However, the two regular LS based water reducing admixtures had

longer setting time, but poor workability retention which was probably related to their

acceleration for the cement hydration in the first hour according to the rate of heat

curves.

At 28 and 91 days, the porosity of the pastes with the LS superplasticizers at w/c of

0.34 was similar to that with the SNF superplasticizer, but higher than that with the

PCE superplasticizer. At w/c of 0.40, the total porosity of the pastes with different

viii
admixtures was not significantly different at 28 days. Pore size distribution of the

pastes changed with time due to cement hydration and they differed with respect to

w/c ratios and admixtures. In general, the proportions of small capillary pores in the

pastes investigated were not significantly different, and the differences were mainly

on the large and medium capillary pores. The pastes with LS superplasticizers had

similar large pores at 91 days compared to the paste with PCE superplasticizer, but

less large pores compared to the paste with SNF superplasticizer at both w/c ratios.

However, the pastes with LS superplasticizers had more medium pores compared to

the pastes with PCE and SNF superplasticizers. The pastes with the regular LS

admixtures (BCS and BCA) appeared to have less large capillary pores at 91 days

compared to those with the superplasticizers. The threshold and critical pore

diameters of the pastes were not significantly affected by the admixtures.

The chemical admixtures investigated affected early compressive strength of the

mortars due to their different retarding effects. However, the strength of the mortars

was not significantly affected by the admixtures beyond 7 days.

Keywords: cement hydration; compressive strength; lignosulphonate; naphthalene;

plastic viscosity; polycarboxylate; pore structure; setting times; superplasticizer; water

reducing admixture; workability retention; yield stress.

ix
List of Notations

A the first peak in heat curves

AFm monosulfoaluminate

AFt ettringite

B the second peak in heat curves

C2 S di-calcium silicate 2CaO.SiO2

C3 A tri-calcium aluminate 3CaO.Al2O3

C3 S tri-calcium silicate 3CaO.SiO2

C4AF tetra-calcium aluminoferrite 4CaO.Al2O3.Fe2O3

CH calcium hydroxide Ca(OH)2

C-S-H or C3S2H8 calcium silicate hydrates, 3CaO.2SiO2.8H2O

d spacing of the crystal planes or diameter

g flow resistance related to yield stress

h relative viscosity related to plastic viscosity

H height

I peak intensity in XRD spectrum

N rotational speed / velocity

n an integer

P pressure

Ri radius of inner cylinder

Ro radius of outer cylinder

t time

T torque

V volume

x
W weight fraction

w/c water-to-cement ratio

α degree of hydration of cement paste

γ surface tension

γ& shear rate

ε total porosity

θ angle of the diffraction peak

λ wavelength of radiation

µ plastic viscosity

τ shear stress

τo yield stress of fresh concrete

ωo angular velocity

xi
List of Acronyms

ASTM the American Society for Testing and Materials

HCP hardened cement paste

HRWRA high-range water reducing admixture

LOI loss on ignition

MIP mercury intrusion porosimetry

(M)LS (modified) lignosulphonate

NEW non-evaporable water

OPC ordinary Portland cement

PCE polycarboxylate

RH relative humidity

rpm rounds per minute

rps rounds per second

sbwc solid by weight of cement

SEM scanning electron microscopy

SMF/PMS sulfonated melamine formaldehyde condensates /

polymelamine sulfonate

SNF/PNS sulfonated naphthalene formaldehyde condensates /

poly-naphthalene sulfonate

SP superplasticizer

SRM standard reference material

SSD saturated surface dry

TG thermogravimetry

WRA water reducing admixture

xii
XRD X-ray diffraction or X-ray diffractometer

XRF X-ray fluorescence

xiii
List of Tables

Chapter 1

Table 1-1 Typical Effects of Water-Reducing Admixtures (Mindess et al, 2003)............. 27

Table 1-2 Classification of WRAs According to ASTM C494 ......................................... 27

Chapter 2

Table 2-1 Classification of Pores (information summarized from Mindess et al, 2003) .. 70

Table 2-2 Solvents and subsequent drying conditions used in solvent exchange ............. 70

Table 2-3 XRD powder pattern of a typical Portland cement (Taylor, 1997) ................... 71

Table 2-4 Summary of a typical cement paste TG graph .................................................. 71

Chapter 3

Table 3-1 Chemical & Mineral Compositions and Physical Properties of Cement Used

...............................................................................................................................96

Table 3-2 Physical properties and sieve analysis of sand.............................................96

Table 3-3 Characteristics of admixtures used in the project ........................................97

Table 3-4 Mix proportion of mortars to achieve an initial yield stress of 75 ± 15 Pa..97

Table 3-5 Mix procedures of mortars and pastes .........................................................98

Table 3-6 Analytical techniques and equipment used in this study..............................98

Table 3-7 Seven samples used to produce C3S calibration chart .................................99

Table 3-8 Process parameters set on BML Viscometer 3 for determination of the yield

stress and plastic viscosity.....................................................................................99

xiv
Chapter 4

Table 4-1 Times of peak appearance in heat curves of pastes......................................... 139

Table 4-2 Amount of CO2 from decomposition of CaCO3 in Fig. 4-9 TG curves .......... 139

Table 4-3 Times for C3S reduction & CH appearance in the cement pastes with and

without admixtures detected by XRD & TG............................................................ 139

Table 4-4 Degree of hydration of pastes at various ages (w/c = 0.34) ............................ 140

Table 4-5 Degree of hydration of pastes at various ages (w/c = 0.40) ............................ 140

Table 4-6 Flow values of mortars with time (w/c = 0.34)............................................... 141

Table 4-7 Repeatability of MIP on mortars with and without admixtures ...................... 141

Table 4-8 Critical and threshold pore diameters for pastes with w/c = 0.34................... 142

Table 4-9 Critical and threshold pore diameters for pastes with w/c = 0.40................... 142

Appendix

Table A-1 Intensity ratios of CH, C3S to anatase in pastes from XRD (w/c = 0.34) ...... 181

Table A-2 Intensity ratios of CH, C3S to anatase in pastes from XRD (w/c = 0.40) ...... 182

Table A-3 Non-evaporable water content in pastes from furnace burning (w/c = 0.34). 182

Table A-4 Non-evaporable water content in pastes from furnace burning (w/c = 0.40). 183

Table A-5 Calcium hydroxide content in pastes from TG (w/c = 0.34).......................... 183

Table A-6 Non-evaporable water content in pastes from TG (w/c = 0.34) ..................... 183

Table A-7 Calcium hydroxide content in pastes from TG (w/c = 0.40).......................... 183

Table A-8 Non-evaporable water content in pastes from TG (w/c = 0.40) ..................... 184

Table A-9 Yield stress, plastic viscosity and flow values of mortars with time (w/c =

0.34).......................................................................................................................... 184

Table A-10 Yield stress, plastic viscosity and flow value of mortars with time (w/c =

0.40).......................................................................................................................... 185

xv
Table A-11 Capillary pore size distribution and total porosity of pastes (w/c = 0.34).... 186

Table A-12 Capillary pore size distribution and total porosity of pastes (w/c = 0.40).... 187

Table A-13 Average and standard deviation of mortar compressive strengths (w/c =

0.34).......................................................................................................................... 188

Table A-14 Average and standard deviation of mortar compressive strengths (w/c =

0.40).......................................................................................................................... 188

xvi
List of Figures

Chapter 2

Fig. 2-1 (a) SMF condensate, (b) SNF condensate, (c) Repeating unit of

lignosulphonate (LS) molecule (d) Molecular structure of polycarboxylate ........72

Fig. 2-2 (a) Flocculation of cement particles resulting trapped water (b)

Deflocculation of cement particles upon adsorption of water reducing admixtures

(Law, 2004) ...........................................................................................................72

Fig. 2-3 Repulsion of cement particles by (a) electrostatic repulsion..........................73

Fig. 2-4 Rate of heat evolution during hydration of Portland cement (Mindess et al,

2003)......................................................................................................................73

Fig. 2-5 Structure of cement pastes (Illston and Domone, 2001) ................................74

Fig. 2-6 Effect of water, water-reducing and air-entraining admixtures ......................74

Fig. 2-7 Critical pore and threshold pore diameters of MIP analysis...........................75

Fig. 2-8 Comparison of MIP and image analysis pore size distribution for the same .75

Chapter 3

Fig. 3-1 Grading curve of fine aggregate (sand) used................................................100

Fig. 3-2 (a) Isothermal calorimeter (b) Sample loading and unloading .....................100

Fig. 3-3 Schematic diagram of an X-ray diffractometer ............................................101

Fig. 3-4 Calibration chart of C3S in materials of interest...........................................101

Fig. 3-5 Schematic diagram of a thermogravimeter...................................................102

Fig. 3-6 Determination of mass loss from a thermogravimetry curve (Haines, 2002)

.............................................................................................................................102

Fig. 3-7 Schematic diagram of a furnace ...................................................................103

xvii
Fig. 3-8 Schematic diagram of the BML-Viscometer (Source: ConTec Ltd., 2003) .103

Fig. 3-9 The relation between (a) torque - rotation speed and ...................................104

Fig. 3-10 A typical ramp down T-N curve from test on mortar by BML Viscometer 104

Fig. 3-11 Schematic diagram of flow table set-up .....................................................104

Fig. 3-12 Schematic diagram of a penetrometer ........................................................105

Fig. 3-13 Schematic diagram of a mercury intrusion porosimeter.............................106

Fig. 3-14 Schematic diagram of a compressive strength tester..................................106

Chapter 4

Fig. 4-1 Effect of SO3 content on heat of cement hydration ........................................... 143

Fig. 4-2 Rate of heat evolution of cement pastes (w/c = 0.34) ....................................... 143

Fig. 4-3 Rate of heat evolution of cement pastes (w/c = 0.40) ....................................... 144

Fig. 4-4 Cumulative heat evolution of cement pastes (w/c = 0.34) ................................ 144

Fig. 4-5 Cumulative heat evolution of cement pastes (w/c = 0.40) ................................ 145

Fig. 4-6 Rate and cumulative heat evolution of two control mixes ................................ 145

Fig. 4-7 C3S reduction in paste with time (w/c = 0.34)................................................... 146

Fig. 4-8 C3S reduction in paste with time (w/c = 0.40)................................................... 146

Fig. 4-9 A typical TG curve showing mass loss over time (w/c=0.40 control paste) ..... 147

Fig. 4-10 CH content in pastes increases with time from TG curves (w/c=0.34)........... 147

Fig. 4-11 CH content in pastes increases with time from TG curves (w/c=0.40) ........... 148

Fig. 4-12 Non-evaporable water content with time from furnace burning (w/c = 0.34). 148

Fig. 4-13 Non-evaporable water content with time from furnace burning (w/c = 0.40). 149

Fig. 4-14 Comparisons of non-evaporable water content ............................................... 149

Fig. 4-15 Average and standard deviation of the initial yield stresses of all mortars...... 150

Fig. 4-16 Yield stress response on mortars with time at 30 ± 3 °C (w/c = 0.34) ............ 150

xviii
Fig. 4-17 Yield stress response on mortars with time at 30 ± 3 °C (w/c = 0.40) ............ 151

Fig. 4-18 Responses of yield stresses of mortars at normalized time ............................. 151

Fig. 4-19 Plastic viscosity response on mortars with time (w/c = 0.34) ......................... 152

Fig. 4-20 Plastic viscosity response on mortars with time (w/c = 0.40) ......................... 152

Fig. 4-21 Change in flow value of mortars with time (w/c = 0.40) ................................ 153

Fig. 4-22 Relationship between yield stress and flow value .......................................... 153

Fig. 4-23 Initial and final setting times of prepared mortars (w/c = 0.34) ...................... 154

Fig. 4-24 Initial and final setting times of prepared mortars (w/c = 0.40) ...................... 154

Fig. 4-25 Relationship between the initial and final setting times of prepared mortars

and time to start the acceleration period in heat curves ........................................... 155

Fig. 4-26 A typical MIP graph (w/c = 0.40, paste with PCE superplasticizer) ............... 155

Fig. 4-27 Total porosity of pastes with w/c = 0.34 at various ages ................................. 156

Fig. 4-28 Total porosity of pastes with w/c = 0.40 at various ages ................................. 156

Fig. 4-29 Total porosities and pore size distributions of the pastes at 1 day................... 157

Fig. 4-30 Total porosities and pore size distributions of the pastes at 3 days ................. 158

Fig. 4-31 Total porosities and pore size distributions of the pastes at 7 days ................. 159

Fig. 4-32 Total porosities and pore size distributions of the pastes at 28 days ............... 160

Fig. 4-33 Total porosities and pore size distributions of the pastes at 91 days ............... 161

Fig. 4-34 Compressive strength of 50mm mortar cubes (w/c = 0.34) ............................ 162

Fig. 4-35 Compressive strength of 50mm mortar cubes (w/c = 0.40) ............................ 162

Appendix

Fig. A-1 TG curves of the control paste (w/c = 0.34) ..................................................... 189

Fig. A-2 TG curves of PCE paste (w/c = 0.34) ............................................................... 189

Fig. A-3 TG curves of SNF paste (w/c = 0.34) ............................................................... 190

xix
Fig. A-4 TG curves of PLS paste (w/c = 0.34)................................................................ 190

Fig. A-5 TG curves of UNA paste (w/c = 0.34) .............................................................. 191

Fig. A-6 TG curves of the control paste (w/c = 0.40) ..................................................... 191

Fig. A-7 TG curves of PCE paste (w/c = 0.40) ............................................................... 192

Fig. A-8 TG curves of SNF paste (w/c = 0.40) ............................................................... 192

Fig. A-9 TG curves of PLS paste (w/c = 0.40)................................................................ 193

Fig. A-10 TG curves of UNA paste (w/c = 0.40) ............................................................ 193

Fig. A-11 TG curves of BCS paste (w/c = 0.40) ............................................................. 194

Fig. A-12 TG curves of BCA paste (w/c = 0.40)............................................................. 194

Fig. A-13 MIP curves for PCE paste (w/c = 0.34) .......................................................... 195

Fig. A-14 MIP curves for SNF paste (w/c = 0.34) .......................................................... 195

Fig. A-15 MIP curves for PLS paste (w/c = 0.34)........................................................... 196

Fig. A-16 MIP curves for UNA paste (w/c = 0.34) ......................................................... 196

Fig. A-17 MIP curves for PCE paste (w/c = 0.40) .......................................................... 197

Fig. A-18 MIP curves for SNF paste (w/c = 0.40) .......................................................... 197

Fig. A-19 MIP curves for PLS paste (w/c = 0.40)........................................................... 198

Fig. A-20 MIP curves for UNA paste (w/c = 0.40) ......................................................... 198

Fig. A-21 MIP curves for BCS paste (w/c = 0.40) .......................................................... 199

Fig. A-22 MIP curves for BCA paste (w/c = 0.40).......................................................... 199

xx
Chapter 1 Introduction

Chapter 1 Introduction

1.1 Background

Over half of the concrete used worldwide contains chemical admixtures. Water

reducing admixtures (WRAs) are most commonly used. Water reducing admixture, as

its name suggests, reduces the water required to attain a given slump. They can be

utilized in the following three ways. Firstly, achieving a desired slump by reducing

the water content while keeping cement content unchanged means an effective lower

w/c ratio, resulting in a general improvement in strength, impermeability and

durability. Secondly, WRAs may be used to increase workability without increasing

water content and cement content, to ease the difficulty in placement. Lastly, they can

be used to reduce cement content either for economic (cement is the most expensive

ingredient in concrete) or technical reason (reduce the heat of cement hydration,

particularly for mass concreting) since a desired slump may be achieved by lowering

the cement content while keeping the w/c ratio unchanged.

Water reducing admixtures can be classified into low-range or regular (WRA, water

reducing capacity of 5% and above) and high-range (HRWRA, water reducing

capacity of 12-30%) (Table 1-1), according to ASTM C494. The HRWRA is

21
Chapter 1 Introduction

commonly referred to as superplasticizers (SPs). The ASTM Standard C494

categorizes several types of such admixtures according to their functions. Types A, D

(retarding) and E (accelerating) are regular water reducing admixtures (Table 1-2);

Types F and G (retarding) are both superplasticizers.

From composition point of view, there are four major categories of superplasticizers,

namely, sulfonated melamine formaldehyde condensate (SMF), sulfonated

naphthalene formaldehyde condensate (SNF), modified lignosulphonates (MLS), and

polycarboxylate (PCE) based superplasticizers. For decades, lignosulphonates (LS)

are one of the most commonly used regular WRAs in concrete industry worldwide

owing to their competitive prices and comparable performances.

The basic mechanisms of water reduction are through dispersion of cement particles

by electrostatic repulsion and/or steric hindrance. Fine particles such as cement grains

have a tendency to flocculate when mixed with water. When they flocculate, a certain

amount of water is often trapped inside agglomerates. Water reducing admixtures are

used to deflocculate and to free the trapped water. However, the effect of WRAs on

concrete performance depends on many influencing factors, such as cement type, mix

proportion, nature and dosage of WRAs, temperature, and time.

Basic LS based WRAs typically have water reduction capacity of 8-12%; modified

LS 15-25%; naphthalene formaldehyde condensed based superplasticizers (SNF)

22
Chapter 1 Introduction

typically 15-25%; and polycarboxylate based admixtures (PCE) more than 30%. Table

1-1 summarizes the water reduction capacity of the different types of water reducing

admixtures and their respective molecular structures and modes of action.

Lignosulphonate has been widely used in concrete for many decades due to its

relatively low price and is mainly regarded as basic water reducing admixture – at a

dosage of 0.05-0.1% they reduce the water requirement by 6 to 10% (Collepardi,

1993). In the past LS based admixtures are only used as normal WRAs since

excessive retardation and entrainment of air occur at high dosages (Ramachandran,

1995). However, significant advances have been made in process, production, and

application of LS based admixtures. There is a wide range of lignosulphonates

available and the performance in concrete varies from basic water reduction and

strong retardation to high range water reduction (Reknes, 2004). With the

development of a new modified lignosulphonate superplasticizer (PLS), it is possible

to produce self-compacting concrete (SCC) with such an admixture (Reknes and

Peterson, 2003).

With the modified lignosulphonate superplasticizers entering the market, its basic

performance, including workability, retardation and strength, have been researched.

However, there is not much information available in the literature on the effect of

these newly developed modified lignosulphonate superplasticizers on cement

hydration, workability retention and pore structure, in comparison to those of other

types of superplasticizers such as naphthalene and polycarboxylate based admixtures

23
Chapter 1 Introduction

and to those of traditional lignosulphonate water reducing admixtures. Therefore, the

current research was carried out.

1.2 Objectives

With limited literature on the earlier mentioned issues, the objectives of this research

project are as follows:

1. To determine the effect of water reducing admixtures and superplasticizers on

cement hydration based on heat evolution, reduction of clinker phase in

cement paste, and increased amount of hydration products;

2. To determine the workability retention of mortars incorporating different

admixtures by means of rheological parameters (yield stress and plastic

viscosity) and flow values changes with time;

3. To determine the retardation of cement hydration in terms of setting times of

prepared mortars and establish possible relationship between the setting times

and heat evolution of cement pastes;

4. To determine the pore structure of plasticized or superplasticized pastes and

the link between cement hydration and pore structure;

5. To determine the compressive strength development of mortars and possible

relations to hydration and pore structure; and

24
Chapter 1 Introduction

6. To evaluate and compare the performances of regular LS based water reducing

admixtures and modified LS based superplasticizers, with respect to fresh and

hardened pastes and mortars.

The focus of this study is on

1. Comparison of the newly developed LS superplasticizer (PLS) with

polycarboxylate (PCE), naphthalene (SNF), and the other modified

lignosulphonate (UNA) superplasticizers; and

2. Comparisons of the newly developed LS superplasticizer (PLS) with regular

lignosulphonate water reducing admixtures (BCS and BCA) and the other

modified lignosulphonate superplasticizer (UNA).

1.3 Scope

In practice, there are numerous situations in which concretes are designed to satisfy

specified workability and w/c requirements. The amount of the admixture may be

adjusted to achieve the requirements of workability and its retention at the specified

w/c. This opens up possibilities of using many different admixtures.

In this research, six admixtures were used which include four superplasticizers (one

polycarboxylate, one naphthalene, two modified lignosulphonates) and two regular

water reducing admixtures (lignosulphonates).

The dosages of admixtures were determined to achieve an initial target yield stress of

25
Chapter 1 Introduction

75 ± 15 Pa for mortars of w/c ratios of 0.34 and 0.40. This yield stress level will

produce concrete with slump of ≥ 100 mm. With w/c of 0.40, all six admixtures were

investigated; whereas for w/c of 0.34, only four superplasticizers were investigated

due to the difficulty in achieving the required initial yield stress by using regular

WRAs. The dosages obtained from mortars were kept the same for the respective

cement pastes.

Following parameters were determined to achieve the objectives:

1. Heat evolution of cement pastes up to 72 hours;

2. Reduction of C3S and increase in calcium hydroxide and non-evaporable water

content in cement pastes at various ages up to 91 days;

3. Changes on the rheological parameters (yield stress and plastic viscosity) of

mortars and flow values with time up to 60 minutes;

4. Setting times of mortars;

5. Pore structure of cement pastes at various ages up to 91 days; and

6. Compressive strength development up to 91 days.

Control pastes of both w/c ratios without admixtures were included in the

investigation of Items 1 and 2 above, but not in Items 3 - 6. The reason was that

setting times, pore structures, and compressive strength are strongly dependent on the

workability and compaction. Without admixtures, specified workability could not be

achieved. Hence, control mixes were not included in the investigation in Items 3 – 6.

26
Chapter 1 Introduction

Table 1-1 Typical Effects of Water-Reducing Admixtures (Mindess et al, 2003)

Water
Typical Increase in ASTM
Classification Common Name Reduction
Dosage#, % Slump, mm Specification
% w/c
Low-Range Regular 0.1 50 ~ 85 5-10 -0.05 C494
Mid-Range Mid-Range 0.5 50 ~ 100 10-15 -0.10
High-Range Superplasticizer 1.0 >100 15-30 -0.15 C494, C1017
#
Active ingreidnet by weight of cement, i.e. solid weight by cement (swbc)

Table 1-2 Classification of WRAs According to ASTM C494

Type of WRA Function of WRA


A Water Reducing Admixtures
D Water Reducing and Retarding Admixtures
E Water Reducing and Accelerating Admixtures
F Water Reducing, High Range Admixtures
G Water Reducing, High Range, and Retarding Admixtures

27
Chapter 2 Literature Review

An admixture is defined in ASTM C125 as “a material other than water, aggregates,

hydraulic cement and fiber reinforcement that is used as an ingredient of concrete or

mortar and is added to the batch immediately before or during its mixing”.

Admixtures, including water reducing admixtures (WRAs), have to fulfill

requirements for their use in concrete. Requirements for slump, water reduction,

setting times, compressive strengths and so on are specified in ASTM and other

relevant standards. For an understanding of the role of WRAs, mechanisms of the

action of the admixtures, workability, microstructure, durability and compatibility

between cement and admixtures, it is necessary to apply various research techniques.

2.1 Nature of Water Reducing Admixtures

Many different types of water reducing admixtures are available on the market.

According to ASTM C494, they are classified into categories based on their functions

in concrete as shown in Table 1-2. Based on the water reducing capacity,

28
Chapter 2 Literature Review

these admixtures can be classified into three broad categories: regular WRAs, mid-

range WRAs and SPs. Regular WRAs can reduce water content by 5 to 10% whereas

SPs have water reducing capacity of 15 to 30% as shown in Table 1-1.

Polycarboxylate based SPs often have water reducing capacities of more than 30%.

2.1.1 Regular Water Reducing Admixtures (WRAs)

There are many different types of regular water reducing admixtures available on the

market. The main compounds used in the manufacture of water reducing admixtures

can be divided into four groups, namely, lignosulphonate, hydroxyl carboxylic acids

and their salts, carbohydrates and other compounds (Ramachandran, 1995).

2.1.1.1 Lignosulphonate

Lignosulphonates, first discovered in 1930s, are the most widely used raw material in

the production of water reducing admixtures (Ramachandran, 1993; Collepardi, 1993).

Lignosulphonate is a by-product from the production of paper-making from wood

whose composition includes about 20-30% lignin. It consists of non-uniform

polyelectrolyte with varying molecular weight distributions, approximately 20,000 to

30,000 with the molecular weights varying from a few hundreds to 100,000 (Rixom

and Mailvaganam, 1999).

In their crude form, lignosulphonates contain many impurities, such as pentose and

hexose sugars, depending on process of neutralization, precipitation and degree of

fermentation, as well as type and age of the wood used (Rixom and Mailvaganam,

1999). Sugars are known to be good retarders of cement hydration processes and the

29
Chapter 2 Literature Review

presence of sugars in lignosulphonate may be accountable for its retarding effect in

cement hydration (Ramachandran et al, 1998). The two common types are calcium

(Ca2+) lignosulphonate and sodium (Na+) lignosulphonate based admixtures. Calcium

lignosulphonates are generally cheaper but less effective whereas sodium

lignosulphonates are more soluble and less liable to precipitation at low temperatures

(Hewlett, 1998).

Regular lignosulphonate at a dosage of 0.05 to 0.1% (solid by weight of cement, sbwc)

can reduce the water requirement in concrete by 6 to 10% (Malhotra, 1997). At higher

dosages, excessive retardation or excessive air entrainment may occur. To reduce air

entrainment, defoamer (commonly used is tributylphophate, TBP) may be added in

the production of these water reducing admixtures. Accelerating admixtures (such as

calcium chloride, calcium formate or triethanolamine) may be added to counteract the

retarding effect.

Because of the relatively low cost of lignosulphonates, there has been continued

interest in utilizing these products in concrete, even in the field of superplasticizers.

By special treatments such as ultrafiltration, desugarization and sulphonation,

modified lignosulphonate superplasticizers have been developed in recent years,

which can compete with melamine sulphonate (SMF) and naphthalene sulphonate

(SNF) based superplasticizers (Ramachandran, 1995).

2.1.1.2 Hydroxyl carboxylic acids and their salts

Salts of hydroxyl carboxylic acids were developed in 1950s. Although there is a

significant increase in their use, they are not used to the same extent as

30
Chapter 2 Literature Review

lignosulphonates. As its name suggests, hydroxyl carboxylic acids have several

hydroxyl (-OH) groups and one or two terminal carboxylic acids (-COOH) groups

attached to a relatively short carbon chain.

They are normally used as an aqueous solution of sodium salts, or occasionally as

salts of ammonia (NH4+) or triethanolamine. Since they are usually synthesized

chemically, they have high purity.

2.1.1.3 Carbohydrates

Carbohydrates include natural compounds such as glucose and sucrose or

hydroxylated polymers obtained from hydrolysis of polysaccharides to form polymers

with a low molecular weight and containing different amounts of glycoside units.

These admixtures also have very strong retarding effects.

2.1.1.4 Other compounds

Quite a number of patents claim that other organic compounds could function as

WRAs. According to a summary on development of WRAs by Ramachandran (1993),

many formulations of WRAs are based on acrylate and methacrylate polymers to

improve workability and/or increase strength. Examples are polymers of alkoxylated

monomers and copolymerizable acid functional monomers.

2.1.2 Superplasticizers (SPs)

Superplasticizers, also known as high range water reducing admixtures, are high

molecular weight and water soluble polymers capable of achieving a given

31
Chapter 2 Literature Review

workability at a much lower w/c ratio compared to that of low-range water reducing

admixtures. The superplasticizers can reduce water content by about 15 – 30% or

even higher.

Superplasticizers are adsorbed on cement particles and hydration products like

calcium hydroxide (CH) and calcium silicate hydrates (C-S-H) adsorb more SP

molecules than cement clinkers (Taylor, 1997). The adsorption rate of the

superplasticizers is affected by several factors such as the amount of tri-calcium

aluminates (C3A) present, the content of soluble sulphates, and the fineness of the

cement used.

Compared to regular WRAs, superplasticizers have lower air entrainment and less

retardation. The low air entraining ability is due to the repeating pattern of polar

groups which provide the molecules with no suitable hydrophobic region. As for the

weak retarding power, it can be attributed either to the assimilation of the

superplasticizers into the cement particles or the weak individual bonds between the

sorbent and sorbate which allow the sorbate to be displaced by ions added to the

product. The weak retarding power of superplasticizers allows the hydration products

to grow despite the presence of the sorbed material (Taylor, 1997).

Based on the main ingredient, there are four main types of superplasticizers, whose

molecular structures are shown in Fig. 2-1:

1. Modified lignosulphonate (MLS), essentially modified and/or purified

lignosulphonate plasticizers with the higher molecular weight fractions selected to

give greater efficiency

32
Chapter 2 Literature Review

2. Sulphonated melamine formaldehyde condensates (SMF), also known as poly-

melamine sulphonate (PMS)

3. Sulphonated naphthalene formaldehyde condensates (SNF), also known as poly-

naphthalene sulphonates (PNS)

4. Polycarboxylate based superplasticizers (PCE), which include polyacrylates,

acrylic esters and sulphonated polystyrenes. These have been most recently developed,

and are sometimes referred to as ‘new generation’ superplasticizers.

2.1.2.1 Modified lignosulphonate(MLS)

Modified lignosulphonate is higher molecular weight lignosulphonate that is

considerably improved by the treatments of the crude by-product to remove

carbohydrate impurities. Though the refining process can enhance the performance of

lignosulphonate, it also causes the modified lignosulphonate to have greater tendency

of entraining air (Hewlett, 1998).

Ramachandran (1995) commented that tailored LS may qualify as a superplasticizer,

but problems associated with its use had to be resolved. He suggested that caution

should be exercised in the selection of deformers, which would be used to de-train air

when LS is at high dosage, as they may affect the aggregate-cement bond. He also

mentioned that concrete with highly dosed LS would show a lower early compressive

strength and would be offset by the use of compatible accelerators.

Typically the LS based admixtures have a more retarding effect than other types of

admixtures. Generally, larger sugar content in admixtures would result in longer

setting times (Ramachandran, 1995). The workability of concrete decreases with time,

33
Chapter 2 Literature Review

known as slump loss; but the rates of slump loss are different for concrete made from

LS, SNF and PCE based admixtures. Hence, concrete made from different WRAs

have different setting times. Water reducing admixtures may also affect cement

hydration and temperature rise in concrete. These effects indicate that the

microstructure of cement paste and concrete may be influenced by the use of different

WRAs, which in turn would affect mechanical properties, permeability and durability

of concrete.

2.1.2.2 Sulphonated melamine / naphthalene formaldehyde condensates (SMF/SNF)

Sulphonated melamine formaldehyde condensates was first developed in Germany

and made commercially available in 1960s. Around the same time, SNF was first

developed in Japan. Both SMF and SNF based superplasticizers are linear anionic

polymers with sulphonate groups at regular intervals. Both types of superplasticizers

tend to give 16 to 25% water reduction. In comparison to SNF, SMF has a higher

molecular weight. Melamine based superplasticizers tend to reduce cohesion in the

mix with little or no retardation, making them effective at low temperatures or where

early strength is critical (Newman and Choo, 2003). On the other hand, less highly

polymerized SNF tends to increase air entrainment to provide cohesion and it also

poses greater retardation on the cement hydration than that of SMF (Newman and

Choo, 2003).

2.1.2.3 Polycarboxylate based (PCE)

Another common type of superplasticizer is polycarboxylate based. There are

polycarboxylates without graft chains, but their dispersing effect is usually limited

34
Chapter 2 Literature Review

and this type of admixture is not widely used (Hanehara and Yamada, 2007). In this

study, only PCE with graft/side chains are discussed. Polycarboxylate based

superplasticizers are more effective complexants1 for divalent and trivalent metal ions

in comparison to the sulphonated polymers. They can reduce water by about 20 to

35% with little retardation and good workability retention. They are very powerful

water reducers and as such a lower dosage is normally used (Newman and Choo,

2003). They are essentially designed for high dispersing ability and their high

workability retention with a minimum setting retardation (Houst et al, 2005). The

PCE based superplasticizers could be further classified into homopolymer or

copolymer. This classification is related to the backbone of the polymer. For

homopolymer, the backbone is made up of only one type of monomer, whereas

copolymer consists of two types of monomers in the backbone. Both types of

polycarboxylate based superplasticizers have side chains made up of polyether. The

term ‘comb polymer’ has been used to describe this molecular structure. The

introduction of polyether side chains improves various performance of

superplasticizer which shows superior water-reducing at low dosage. When dosage

was 0.6%, the water reduction was as high as 36% in concrete and concrete slump

loss was very little in 90 min (He et al, 2005). Li et al (2005) also reported that the

high dispersing and flowing retention properties of PCE superplasticizers are mainly

affected by the length of side chains through steric repulsive force. Sugamata et al

(2003) found that the workability retention was a combination of the amount of PCE

superplasticizer adsorbed and the side chains of PCE molecules which extended out to

form a thick layer on cement particles, giving rise to greater repulsion.

1
Substances capable of forming a complex compound with another material in solution

35
Chapter 2 Literature Review

The PCE superplasticizers are commonly used at a dosage up to approximately 1%

solid by weight of cement (sbwc). Instead of air entrainment, these superplasticizers

may actually decrease the amount of air entrained as a result of greater fluidity of the

mix (Taylor, 1997). When an overdose of superplasticizers is used, undesirable effects

such as excessive retardation and excessive slump loss may occur.

The dispersing action of the PCE superplasticizers is not only limited to ordinary

Portland cement. It can also be used together with other mineral admixtures to

produce higher quality concrete. Several different types of chemical and mineral

admixtures have been found compatible with these superplasticizers. These mineral

and chemical admixtures include fly ash, blast furnace slag, retarders, accelerators and

air entraining agents (Ramachandran et al, 1998).

It has been reported that PCE based superplasticizers ensured high plasticity concrete

mixes at dosages about one third that of conventional SNF based superplasticizer

(Faliman et al, 2005).

2.2 Mechanisms of Water Reduction

Water reduction by the regular WRAs and superplasticizers is achieved by

deflocculating the cement particles, thereby releasing the water trapped between the

cement agglomerates and making it available for mixing and workability. Without the

incorporation of these admixtures, the positively and negatively charged cement

particles will be attracted to each others, leading to flocculation as shown in Fig. 2-

2(a). Flocculation of these cement particles will trap part of the mixing water,

resulting in less water to be available for workability and cement hydration. With the

incorporation of these admixtures, flocculation of the cement particles is prevented or

36
Chapter 2 Literature Review

minimized. These chemical admixtures are surface active agents which when

adsorbed by the cement particles will give them negative charges that cause repulsion

between particles. With the deflocculation of the cement particles, the trapped water

will be released and made available for workability and cement hydration. On top of

that, more surface areas of the cement particles will be exposed for the hydration

process. This is illustrated in Fig. 2-2(b). The water reducing capacity of these

admixtures allows a given workability to be achieved at a lower water requirement or

increase the workability for the same mix proportion, hence early strength gain.

There are basically three mechanisms to explain for the water reducing capability of

these admixtures. They are electrostatic repulsion, steric hindrance, and solid-liquid

affinity. Among the three mechanisms, electrostatic repulsion and/or steric hindrance

are dominating in deflocculation of cement particles. Besides all the three mechanism,

retardation of the cement hydration and air entrainment of these admixtures also aid

the deflocculation process, enhancing the water reducing capacity of these admixtures.

2.2.1 Electrostatic Repulsion

Electrostatic repulsion is generated as a result of increased magnitude of the zeta

potential. In the absence of these admixtures, the charges on the cement particles were

too small to determine the zeta potential. However, with the incorporation of the

admixtures, the zeta potential increases with the increase of the negative charges on

the cement particles. When all these cement particles carry sufficient magnitude and

same sign of surface charge, these particles will repel each others, resulting in

electrostatic repulsion as illustrated in Fig. 2-3(a).

37
Chapter 2 Literature Review

2.2.2 Steric Hindrance

The steric hindrance effect, shown in Fig. 2-3(b), is due to the oriented adsorption of

the admixtures’ molecules which weaken the attraction between the cement particles.

One side of the admixtures’ molecules will be attached to the cement particles

whereas the other side to water. As a result of such attachment, a watery and

lubricating film is formed around the cement particles, weakening the attraction forces

between the cement particles. The mechanism of steric hindrance is exhibited mainly

by water reducing admixtures having branched molecular structures (Ramachandran

et al, 1998).

2.2.3 Solid-Liquid Affinity

The water reducing capacity of these admixtures could also be explained by the

increase in the solid-liquid affinity. When the solid-liquid infinity is increased with the

adsorption of these admixtures, the cement particles are attracted more to the water

rather than to each others. This will result in deflocculation of the cement particles.

2.2.4 Mechanisms of WRA and SP of Different Natures

The dispersing effect of LS based regular WRAs is due mainly to the mechanism of

electrostatic repulsion and to the mechanism of steric hindrance to some extent

(Uchikawa et al, 1997). Although LS based water reducing admixtures have a linear

molecular structure, the steric effect is caused by the very cross-linked molecules

taking up a relatively large volume on the cement surface (Ramachandran et al, 1998).

The mode action of superplasticizers is that they cause a combination of mutual

38
Chapter 2 Literature Review

repulsion and steric hindrance between the cement particles. Opinions differ about the

relative magnitude and importance of these two effects with different superplasticizers.

For example, Yoshioka et al (2002) believed that the dispersing effect for PCE based

admixture is accountable by the steric hindrance effects, which was supported by

Uchikawa et al (1997) and Flatt et al (2000), while the dispersing effect is due solely

to electrostatic repulsion for SNF based admixtures. However, the general consensus

(Collepardi, 1998; Hewlett, 1998) is that

1. For SNF and modified LS superplasticzers, the dominant mechanism is

electrostatic repulsion;

2. For PCE superplasticizers, steric hindrance is equally if not more important than

electrostatic repulsion. This is due to a high density of polymer side chains on the

polymer backbone which protrude from the cement particle surface.

2.3 Portland Cement Hydration

2.3.1 Chemistry of Portland Cement Hydration

For an initial period after mixing, cement paste gradually loses the fluidity and starts

to stiffen at a much faster rate upon initial set. Strength gain does not start until after

the final set. The gain is rapid for the next few days, and continues for at least a few

months at a steadily decreasing rate.

The cement hydration reactions are exothermic, and the rate of heat evolution is a

direct indication of the rate of reactions. Figure 2-4 shows a typical curve of the rate

of heat evolution during hydration of Portland cement paste. Immediately after the

contact of the cement with water, there is a high peak (A), lasting only a few minutes.

39
Chapter 2 Literature Review

This quickly declines to a low value for the induction (or dormant) period, when the

cement is relatively inactive. This may last for a few hours and the rate then starts to

increase rapidly, at a time corresponding roughly to the initial set, and reaches a broad

peak (B), some time after the final set. The reactions then gradually slow down,

sometimes with a narrow peak (C).

The main contribution to the short, intense first peak (A) is rehydration of calcium

sulphate hemihydrate, which is from gypsum decomposition during grinding process

(Coole, 1984). Additional contributions to this peak come from the hydration of the

free lime, the heat of wetting, heat of solution and the initial reactions of the

aluminate phases (Bensted, 1987).

In the early stages of hydration, C3A reacts very violently with water, resulting in

“flash set” of the paste if no gypsum is available in the system. Gypsum is added in

Portland cement to prevent the flash set because it reacts with C3A to produce

ettringite (AFt) and the reaction is much slower than that of the C3A alone. When

gypsum in the cement is depleted, ettringite is gradually transformed to calcium

monosulfoaluminate (AFm). If all the gypsum is consumed before all the C3A, the

direct hydrate C3AH6 is formed. The combined effect of AFt conversion to AFm and

direct hydration of C3A causes the short third peak C, which can occur 2 or 3 days

after hydration starts. Whether this peak occurs depends on the relative amount of

gypsum and C3A in the cement. The C4AF phase reacts over similar time scales, and

the reactions and products are both similar to those of C3A. The products contribute

little to the overall cement behaviour.

40
Chapter 2 Literature Review

The two calcium silicates C3S and C2S form the bulk of anhydrated cement, and their

hydration products (calcium silicate hydrates and calcium hydroxide) give the

hardened cement paste most of its mechanical properties such as strength and stiffness.

Their reactions and reaction rates therefore dominate the properties of the hardened

cement paste and concrete. Their hydration reactions are shown in Eqs. 2-1 and 2-2.

2C 3 S + 11H → C 3 S 2 H 8 + 3CH (2-1)

2C 2 S + 9 H → C 3 S 2 H 8 + CH (2-2)

Figure 2-5 provides a visual illustration of cement hydration, where clinker phases

reduce in quantities and hydration products increase in volume with time. This leads

to the decrease in pore volume as the hydration products have lower specific gravity

than anhydrated cement particles, and thus occupy large volume in the paste.

The aluminate (C3A) phase is the most reactive, and therefore has the largest impact

on workability as it consumes large amounts of water upon hydration. The alite (C3S)

is the second most reactive phase, with strong impact on set and strength development.

Therefore the setting time is strongly affected by the free water content and by the

rate of reaction of both aluminate and alite (Sandberg, 2004).

In summary, cement hydration is a complicated process that includes a number of

exothermic chemical reactions. The characteristics of the heat of hydration are unique

for each concrete mix under each environmental condition. The hydration process also

directly influences paste/mortar/concrete workability, setting behavior, rate of

strength gain, and pore structure development.

41
Chapter 2 Literature Review

2.3.2 Heat Evolution of Portland Cement Hydration

2.3.2.1 Measurement of heat evolution of cement hydration

Different test methods are available for measuring heat of hydration, including the

calorimeter method. There are four major calorimeters available: adiabatic, semi-

adiabatic, isothermal, and solution calorimeter. Isothermal conduction calorimetry has

been extensively used for monitoring the cement hydration and cement compounds.

Depending on the temperature, w/c ratio, nature and dosages of admixtures, the

intensity of the heat liberated varies with time. Much of the heat during cement

hydration is given off in the first few days. In conduction calorimeter, the rate of heat

evolution is recorded as a function of time. The area under the rate of heat curve gives

the total heat generated for a recorded period of time, varying from hours to days.

2.3.2.2 Effect of the admixtures on heat evolution of cement hydration

The measurement of the rate of heat evolution provides information on the rate of

cement hydration in the presence of WRAs. The hydration is generally delayed in the

case of cement hydration incorporated with WRAs. Bensted (1987) observed that the

first peak (A) was increased in cement paste with calcium LS admixture compared to

the control. The second peak (B) was delayed for the paste with WRA but the total

heat evolution was higher for plasticized paste. He further reasons that although the

calcium silicate hydration is delayed, the hydration of other cement phases (aluminate

and ferrite) increase and furthermore, once the delayed calcium silicate hydration

starts, it is effectively a delayed acceleration as indicated by the calorimeter heat

curve. Also, he comments that C3S can be either accelerated or retarded depending on

the types of superplasticizers used. Uchikawa et al (1995) found that the peak due to

42
Chapter 2 Literature Review

C3S hydration (Peak B) was delayed by one hour when a SNF admixture was added

with the mixing water, and was shifted more to the right when addition of SNF was

delayed. Uchikawa et al (1995) observed that the second peaks (Peak B) in a heat

curve were similar for a sodium LS and a SNF based admixture, if other conditions

are the same. Pang et al (2005a) reported that a modified LS superplasticizer

remarkably delayed heat evolution of cement hydration. Peschard et al (2004)

suggested that the origin of retardation could be linked to an adsorption of admixtures

on the first hydrates forming a less permeable coating. On the other hand, Koizumi et

al (2007) found that SNF superplasticizer did not delay cement hydration at low

dosages based on the heat curves from calorimeter.

The use of superplasticizers reduces the heat generation during the setting period,

however, they do not affect the total heat of concrete (Wang et al, 2006). Aïtcin et al

(1987) also found that the total heat of hydration was not affected when SNF

superplasticizer was added in cement paste, even at a dosage as high as 1.45% sbwc.

2.4 Effect of the Admixtures on Cement Hydration

It is stated that water reducing admixtures do not change the kinetic laws, the Avrami-

Erofeev nucleation and growth law, and the three-dimensional diffusion involved in

the hydration processes (Ridi et al, 2003). However, hydrations of calcium silicates

and aluminates in cement may be affected by the presence of admixtures.

43
Chapter 2 Literature Review

2.4.1 Effect of LS Admixtures

It is generally agreed that LS admixtures retards the hydration of C3S (Odler and

Becker, 1980; Bishop and Barron, 2006) or silicates (Ciach, 1971; Myrvold, 2006) or

cement (Mollah et al, 1995; Mikanovic et al, 2000; Carazeanu et al, 2002; Pang et al,

2005) at early ages. Modified LS superplasticizers also delayed cement hydration

(Pang et al, 2005a). Sakai et al (2006) found that the degree of C3S hydration in the

presence of refined LS superplasticizer at 3 days was smaller than that in other pastes

with SNF and PCE superplasticizers, and the degree of the hydration of C3S at 28

days was almost the same as that in pastes with the other superplasticizers.

Researchers have been investigating how the silicates are delayed. Khalil and Ward

(1973) attributed the delay of silicate hydration to an adsorption of the admixture on

the surface of cement particles. Bishop and Barron (2006) proposed that the calcium

ions (Ca2+) from LS based admixtures were involved in the formation of a semi-

permeable layer onto the cement grains, which acted as a diffusion barrier to delay the

cement hydration. Myrvold (2006) suggests that the ions released from the rapid

aluminates (C3A/C4AF) hydration may modify lignosulphonates and that the modified

lignosulphonates covers the silicate phases, which in turn slow down silicate

hydration.

In addition to the effect on the hydration rate, Odler and Becker (1980) found that the

amount of CH in pastes with the LS admixtures was distinctly lower than that in

control paste without the admixtures, and they suggest that the stoichiometric

composition of the C-S-H phase formed in the former may have higher Ca/Si ratio

than the latter without the LS admixtures. Dodson (1967) also found that the addition

of calcium LS admixture to C3S system could alter the morphology of calcium

44
Chapter 2 Literature Review

hydroxide to form irregular crystals differing dramatically from the normal hexagonal

crystals. In addition, the LS admixtures seem to have an effect on the number of

calcium hydroxide crystals formed per unit volume of paste, although this can be

either increased or decreased depending on the type of material used (Rixom and

Mailvaganam, 1999).

The hydration of the silicates are delayed during early ages, however, it was reported

that both LS based (Bishop and Barron, 2006) and modified LS based (Pang et al,

2005a) admixtures accelerated ettringite formation at an early age. It was further

observed that the ettringite crystals in pastes with LS admixtures (Pang et al, 2005)

and modified LS admixtures (Pang et al, 2005a) were finer compared to those in

control paste without the admixtures. Odler and Samir (1987) reported otherwise.

They found that in the presence of a sodium LS, the hydration of C3A was also

retarded along with the C3S phase.

2.4.2 Effect of SNF Admixtures

Odler and Becker (1980) found that SNF admixtures retarded the hydration of pure

C3S and C3S component of Portland cement. The Ca/Si ratio of the C-S-H gel formed

in C3S hydration was distinctly increased by the admixtures. Koizumi et al (2007)

reported that a lower Ca/Si ratio at late age was observed. It was shown (Aïtcin et al,

1987) that SNF based superplasticizer, used at high dosage (1.45% sbwc) retarded the

hydration process of Portland cement. Based on the High Frequency Arc (HFA)

diameters obtained from AC impedance spectroscopy measurements and porosities

from mercury intrusion porosimetry (MIP), Gu et al (1994) concluded that SNF

superplasticizers had retardation effects at early ages (more evident after 7 hours), and

45
Chapter 2 Literature Review

the effect was still notable at 28 days. The slower cement hydration reactions in the

presence of superplasticizers were confirmed by SEM examination and Ca(OH)2

content determined by thermal gravimetric (TG) method. The heat studies by Simard

et al (1993) supported this argument. On the other hand, Sakai et al (2006) and

Uchikawa et al (1992) found that SNF did not noticeably delay cement hydration.

Collepardi et al (1980) found that there were no substantial changes in the rate of C3A

hydration when 0.6% SNF was added. These different opinions presented are likely

related to the different dosages of the SNF admixtures used by researchers.

It is well known that SNF superplasticizer molecules are not only adsorbed on

anhydrated cement particles but also on some of their hydrates (Yilmaz and Glasser,

1989; Sarkar and Xu, 1992). It was found that the adsorbed SNF superplasticizer

molecules slowed down (Ramachandran, 1995; Hekal and Kishar, 1999; Mikanovic et

al, 2000; Mollah et al, 2000; Pourchet et al, 2006) or even stopped (Prince et al, 2002)

the growth of ettringite. However, normal growth of ettringite resumed when the

superplasticizer is consumed (Prince et al, 2002). The reasons for the slow down are

not clear. Prince et al (2003) found that the adsorption of SNF superplasticizer

molecules onto cement particles and their hydrates decreases the dissolution rate of

the constituents. Pourchet et al (2006) suggests that the slow down in the average rate

of ettringite precipitation may be linked to a decrease in C3A dissolution rate.

Roncero et al (2002), however, had different findings. The SNF superplasticizer they

investigated caused a lower C–S–H gel formation but the superplasticizer accelerated

formation of ettringite compared to the control paste where no admixture was added.

46
Chapter 2 Literature Review

Not only the rate of ettringite precipitation may be altered, but the morphology of

ettringite may be modified as well. Hekal and Kishar (1999) studied a sodium salt of

SNF superplasticizer on the hydration of C3A with gypsum in a suspension. It was

found that the presence of the SNF superplasticizer caused a decrease in the size of

ettringite crystals, which agreed with the finding from Rößler et al (2007) that the

crystal size decreased as the dosage of superplasticizer increased. Prince et al (2002,

2003) also observed small massive ettringite clusters, rather than the usual needle-like

ettringite crystals, on an amorphous looking paste when an SNF superplasticizer was

added. However, when the SNF superplasticizer molecules were depleted, ettringite

crystals started to grow again in their usual shape (Prince et al, 2002).

2.4.3 Effect of PCE Admixtures

It is found that at early ages PCE admixtures retard cement hydration and the effect is

more pronounced at higher dosages (Kreppelt et al, 2002; Puertas et al, 2005). Based

on scanning electron microscope (SEM) observations, Kreppelt et al (2002) found that

very few hexagonal ettringite crystals were formed even after 24 hours of reaction

whereas Lothenbach et al (2007) did not observe distinct retardation of ettringite

formation. This may be due to the preferable adsorption of PCE onto C3A, observed

by Heikal et al (2006). Xu and Beaudoin (2000) also reported that PCE

superplasticizer reduced the rate of hydration during the first day. However, the

amount of C-S-H gel was similar after 28 days of hydration (Jolicoeur and Simard,

1998; Xu and Beaudoin, 2000; Puertas et al, 2005). In fact, Lothenbach et al (2007)

did not observe significant difference in the amount of hydrates formed after 6 days.

Although mineralogical analyses showed that the same hydration products were

47
Chapter 2 Literature Review

formed in all pastes - mainly C–S–H gel (Puertas et al, 2005), it was revealed that a

few alterations in the structure and composition of C–S–H gel existed in the PCE

pastes. They found that pastes without admixtures had a somewhat higher Ca/Si ratio

than those containing 1% PCE superplasticizer at 2 days and that the Ca/Si ratio was

slightly lower in the control paste at 28 days. The lower Ca/Si ratio at a late age was

also reported by Koizumi et al (2007). Zhang et al (2006) found that PCE

superplasticizer molecule increased the crystallinity of C-S-H by intercalating into the

inner interlayer surface of the C-S-H.

2.5 Effect of the Admixtures on Workability

2.5.1 Workability and Rheological Parameters

Before discussing the effect of admixtures on workability and rheological parameters,

it is helpful to examine the basic principles of rheology. Rheology is a science dealing

with the deformation and flow of matter under stress. The simplest fluid is so-called

Newtonian fluid, which obeys Newton’s law (Eq. 2-3) of viscous flow.

τ = µγ& (2-3)

where τ = shear stress;

µ = coefficient of viscosity; and

γ& = rate of shear, or the viscosity gradient.

The relation of the shear stress vs shear rate of Newtonian fluids, which include very-

diluted suspensions of solid particles in liquid, is a straight line through the origin in a

48
Chapter 2 Literature Review

shear stress τ - shear rate γ& graph. However, concentrated suspensions such as mortar

and concrete do not behave as Newtonian fluids. There is considerable evidence that

the flow behaviour of fresh mortar and concrete can be reasonably approximated by

Bingham Model (Eq. 2-4) (Mindess et al, 2003; Ferraris and de Larrard, 1998).

τ = τ o + µγ& (2-4)

where τ = shear stress;

τ o = yield stress or yield value;

µ = plastic viscosity; and

γ& = rate of shear, or the viscosity gradient.

They have to overcome a definite shear stress before flow can occur. This shear stress

is referred to as yield stress ( τ o ). Once the flow starts, the rate of flow is controlled by

the plastic viscosity (µ). For such materials, a single-point test such as slump or flow

table value may not be sufficient to describe their flow behavior.

Rheological parameters of fresh concrete are of great importance in understanding

workability of fresh concrete. The yield stress and plastic viscosity are two

rheological parameters, which are dependent on shear history, shear rate, and time of

measurement (Banfill, 2003). They are influenced by the use of chemical admixtures

such as air-entraining admixtures and water-reducing admixtures (Gjørv, 1994). The

general trends of the influences of regular water-reducing admixtures,

superplasticizers, and air entraining admixtures are shown in Fig. 2-6.

49
Chapter 2 Literature Review

Most rotational rheometers are based on the principle that the material is stirred at a

controlled speed and the resulting torque is measured (Ferraris and Martys, 2003). In

the case of a Newtonian fluid, the viscosity is defined as the ratio between the shear

stress and shear rate (Mindess et al, 2003). Concrete and mortar are generally

accepted to be Bingham fluids (Ferraris and de Larrard, 1998). In such materials, the

plastic viscosity is defined as the slope of the shear stress versus shear rate once the

yield stress is overcome. Most rotational rheometers measure torque versus rotational

speed. Therefore, to obtain the true or absolute plastic viscosity, the slope of the curve

should be corrected by a function, f, which depends on the rheometer geometry and

experimental conditions.

2.5.2 Effect of Admixtures on Initial Workability

As shown in Fig. 2-6, regular WRAs and SPs generally reduce the initial yield stress

substantially compared to that of control mix. Gołaszewski and Szwabowski (2004)

found that PCE superplasticizers with the same dosage as SNF superplasticizer

considerably reduced g value (related to yield stress), but increased h values (related

to plastic viscosity) than those with SNF superplasticizer. It was reported that a

modified LS with an average molecular weight around 10,000 showed comparable

fluidity to SNF superplasticizer (Ouyang et al, 2006). For plain concrete with normal

w/c ratio of 0.40 or high w/c ratio of 0.5, the effectiveness of PCE and SNF

superplasticizers were similar (Gołaszewski and Szwabowski, 2004). This is quite

expected as superplasticizers are created for low w/c ratios.

At the same dosage, PCE superplasticized mixes are generally the most workable,

50
Chapter 2 Literature Review

followed by those with SNF and LS admixtures. The performance of the modified LS

superplasticizers is dependent on the modification process and the sugar content,

counter-ions, and average molecular weight of the products. They may be comparable

to SNF superplasticizers in terms of their effect on the initial workability of mortar

and concrete.

There are few studies on plastic viscosity of mortars incorporating water reducing

admixtures. Some researchers studied apparent viscosity2 of cement pastes instead of

plastic viscosity due to their shear thinning effect. Odler and Becker (1980) found that

SNF and LS based admixtures lowered the apparent viscosity of both systems

compared with that of the control Portland cement paste.

Many factors can influence workability of cement paste, mortar and concrete, for

example, cement chemistry, nature and dosage of admixtures, temperature and age.

Lombois-Burger et al (2006) found that the yield stress level of cement pastes was

governed by the adsorption of superplasticizers, which was in competition with SO42-

to be adsorbed onto cement particles. They found that when cement containing mainly

hemihydrate was used, a rapid decrease of SO42- concentration with time resulted in

an increases in SNF adsorbed, which in turn led to less flow loss of cement paste.

Similar findings were reported by Hanehara and Yamada (1999) and Nakajima and

Yamada (2004). Zhor (2006) studied the effects of functional group of admixtures on

cement pastes and found that sulphonate had a very low correlation with dispersing

effect while carboxyl was the group most correlated to dispersing effect. The

2
It is defined as the ratio of shear stress to shear rate (at a given shear rate), as if the liquid
were Newtonian. If the liquid is actually non-Newtonian, the apparent viscosity depends on
the type and dimensions of the apparatus used and the shear rate.

51
Chapter 2 Literature Review

workability of cement paste was also found to be dependent on counter-ions of WRAs,

with Na+ better than Ca2+ for SNF admixtures (Simard et al, 1993).

2.5.3 Effect of Admixtures on Workability Retention

Although water reducing admixtures are able to lower the initial yield stress, the

pastes with the admixtures also have workability loss with time just like the control

cement pastes. Björnström and Chandra (2003) reported that both yield stress and

plastic viscosity increased with time, from immediately after mixing to 45 min, for

cement pastes at w/c of 0.3 made with LS, SNF and PCE superplasticizers for two

cements with low and high C3A contents. However, neither the initial yield stresses

nor the admixture dosages were controlled at the same level as workability retention

is dependent on the initial workability.

It is generally agreed that concrete with SNF superplasticizer have very high slump

loss (Houst et al, 1999; Ouyang et al, 2006). Lim et al (1999) reported that SNF paste

lost over 50% of its initial slump at 120 min. Nawa et al (2000) found that SNF

superplasticizer had significant loss of flow but the loss decreased with an increase in

the dosage of SNF superplasticizer.

The rates of such loss may differ for admixtures of different natures. Chan et al (1996)

studied PCE and SNF superplasticizers at their respective saturation dosages, and

found that the workability retention of PCE was better than that of SNF. The

saturation dosage is defined as the dosage of water reducing admixtures beyond which

no significant increase in the workability is obtained. At the same dosage,

Gołaszewski and Szwabowski (2004) found the same workability retention of mortars

52
Chapter 2 Literature Review

with PCE and SNF superplasticizers. The flow variations with time of pastes

(Uchikawa et al, 1995) and mortar (Houst et al, 1999) with PCE superplasticizer are

small.

Houst et al (2005) reported that a newly developed LS superplasticizer was much

more effective in terms of workability of concrete as a function of the dosage

compared to SNF superplasticizer. Chandra and Björnström (2002) observed that the

slump loss measured by flow value was higher with SNF compared with that with LS

admixtures when the initial flows were similar.

It has been recognized that flow and slump of concrete are related more to yield stress

than plastic viscosity. Thus, a lot of researches have been done on yield stress in the

last two decades. Plastic viscosity is not well studied. Gołaszewski and Szwabowski

(2004) found that PCE superplasticizer significantly reduced h values (related to

plastic viscosities) with time from 10 to 50 minutes while SNF one just slightly

decreased.

Reasons for workability loss are not yet quite understood. Bonen and Sarkar (1995)

concluded that the higher the ionic concentration in the pore solution of an SNF

superplasticized paste, the faster the slump loss. Hanehara and Yamada (1999)

suggested that slump loss and stiffness were caused by the production of large

amounts of ettringite based on their results of PCE, SNF, and LS admixtures. It was

reported (Uchikawa et al, 1983; Chandra and Björnström, 2002) that LS based

admixtures took up Ca2+ from the pore solution and thus the slump loss of mortars

with LS admixtures was lower than for those with SNF admixtures. Myrvold (2007)

53
Chapter 2 Literature Review

suggests that workability retention of cement paste with LS admixtures may be

attributed to the adsorption of the LS on the silicate phases (C3S and C2S) and on the

hydration products (particularly CH, AFt) which causes retardation.

2.6 Effect of the Admixtures on Setting

Uchikawa et al (1984) found that when ion concentration (particularly Ca2+, SO42- and

OH-) was low, a large amount of needle-like ettringite crystals were produced, which

caused the stiffness and pseudo-setting of cement pastes. According to them, the PCE

superplasticizer seems to affect initial setting which was thought to correlate to C3S

hydration. Further, Uchikawa et al (1992) found that LS admixtures form complex

salts with Ca2+ in pore solutions more easily than SNF admixtures. Thus the LS

admixtures delayed the saturation of Ca2+ in the pore solutions which delayed the

setting of cement paste more than the SNF admixtures.

When SNF superplasticizer is added at small dosages, there is no or very little

retardation. However, Simard et al (1993) showed that the increase in the retardation

was roughly proportional to the SNF concentration (0.4-0.8% sbwc) in four cement

pastes with different C3S, C3A, and SO3 contents. Agarwal et al (2000) reported very

strong retardation of cement paste at SNF dosages of 1% and beyond.

Houst et al (2005) reported that the set retardation of a newly developed LS

superplasticizer was the same as traditional LS admixtures which retard the setting

more than SNF superplasticizer.

Vikan and Justnes (2007) suggest that the duration of the induction period is less

54
Chapter 2 Literature Review

dependent on the silicate phases (C3S and C2S) than the aluminates (C3A and C4AF)

in cement pastes incorporating SNF and LS admixtures.

2.7 Effect of the Admixtures on Pore Structure & Strength

Development

2.7.1 Principle of Mercury Intrusion Porosimetry and Characterization of

Pore Structure

Mercury intrusion porosimetry (MIP) is a method often used to characterize the pore

structure of cement paste, mortar, and concrete in spite of its limitations. The MIP test

is governed by the Washburn (1921) equation, Eq. 2-5.

4γ cos θ
d =− (2-5)
P

where P = pressure, MPa;

γ = surface tension of the liquid, N/m;

θ = contact angle of the liquid with the solid, degree; and

d = diameter of the capillary, nm.

The following parameters are commonly obtained from the MIP test: (1) total porosity,

(2) critical pore diameter, (3) threshold pore diameter and (4) pore size distribution.

55
Chapter 2 Literature Review

2.7.1.1 Total porosity

Porosity can be defined into two classes: total (or absolute) porosity and the effective

porosity. Total porosity, ε , including both open and closed pores, is the volume of

pores with respect to bulk volume of the material (often expressed in percentage), as

in Eq. 2-6:

V pore
ε= × 100% (2-6)
Vbulk

where V pore = the total pore volume in the bulk material; and

Vbulk = the bulk volume of the material.

Effective porosity is the fraction of open and interconnected pores with respect to bulk

volume of the material. The test with mercury intrusion porosimeter, therefore, can

only determine the open pores volume; hence effective porosity. The porosity

determined by MIP is defined as the ratio between the total injected mercury volume

and the total volume of the sample.

2.7.1.2 Critical pore diameter

The critical pore diameter (dc) is the pore size corresponding to the highest rate of

mercury intrusion. This is the point with the steepest slope of the cumulative intrusion

volume against pore diameter as shown in Fig. 2-7, therefore the point where

[dV/d(lnD)] is the maximum.

56
Chapter 2 Literature Review

2.7.1.3 Threshold pore diameter

Winslow and Diamond (1970) defined threshold diameter (dt) of pores as that

corresponding approximately to the minimum diameter of channels that are

essentially continuous through the paste at a given age. The threshold diameter is the

diameter on the cumulative pore volume curve below which the pore volume rises

sharply. Above this value, there is comparatively little intrusion but immediately

below, the greatest portion of intrusion occurs.

2.7.1.4 Pore Size Distribtuion

At any stage of cement hydration, hardened cement paste (HCP) consists of solid

products such as calcium silicate hydrates, crystals of calcium hydroxide, calcium

sulphoaluminate hydrates, anhydrated cement, and pore space originally occupied by

water. The pores in the HCP can be classified into three categories, namely gel pores,

capillary pores, and air voids (see Table 2-1).

Gel pores do not play any significant role in the flow of water through concrete.

Capillary pores represent the space not filled by solid components in HCP and

therefore its volume and size depends on the distance between anhydrated cement

particles and degree of hydration (Parrott and Killoh, 1984). In well hydrated and low

w/c ratio pastes, the capillary pores may range from 10 nm to 50 nm (Mindess et al,

2003). Air voids in concrete are either entrapped during casting or intentionally

entrained using an air-entraining agent. The entrapped air may be as large as 3 mm

and the entrained air may be in the range of 50 – 200 µm. Both are larger than

capillary pores and have an effect on concrete permeation.

57
Chapter 2 Literature Review

2.7.1.5 Evaluation of MIP

The microstructure of cement paste influences bulk properties such as compressive

strength, permeability and ion migration (Zhang and Glasser, 2000). The MIP

measurements also showed that the total volume as well as the distribution and

connectivity of pores significantly controlled various properties of concretes

(Diamond, 2000).

The MIP test has been used in the determination of microstructure for a long time as it

is one of the analytical techniques that permits an analyst to acquire data over such a

broad dynamic range using a single theoretical model (Webb, 2001).

However, the MIP method has limitations. Besides the required spherical pore shape

assumption, “ink-bottle” or “neck-bottle” (Diamond and Leeman, 1995) effect has

been found to have significant influences on the pore size distributions. As seen in

Fig. 2-8, large discrepancies were observed between pore size distribution determined

by the MIP method and that from image analysis, which are less affected by the

morphological features of the pore structure. Diamond (2000) criticized that the

features of pore structure characterized by the MIP method are not representative of

the real pore structure because of improper assumptions made on the shape of the

pores and their connectivity in concrete in the MIP method. Based on his study, he

further limited MIP credit only to threshold diameters and intruded pore space

measurements.

According to Cook (1991), the MIP method will indicate smaller than actual porosity

values where pores are too small or too isolated to be intruded by mercury. On the

other hand, MIP porosities may be closer to actual values than those indicated by

58
Chapter 2 Literature Review

other techniques where mercury pressures can collapse small pores or break through

to isolated pores (Cook and Hover, 1999).

2.7.2 Effect of Admixtures on Pore Structure of Cement Paste

It is generally understood that the porosity of cement paste with or without admixtures

decreases with hydration time and that the greater the rate of hydration, the more

rapidly the porosity decreases. However, pastes with water reducing and retarding

admixtures may have different effects on the development in pore structures at a given

w/c ratio (Ramachandran et al, 1998).

2.7.2.1 Effect of PCE Admixtures

Puertas et al (2005) found that PCE admixtures used in the study modified

microstructure in the pastes which in turn reduced porosity and refined pore size. Xu

and Beaudoin (2000) found that PCE superplasticizer lowered overall porosity of

mortar at 28 days compared to the control mortar; however, the threshold pore

diameters were similar (around 70 nm) with or without the admixture. Xu and

Beaudoin (2000) also observed that the volume of pores ≥ 0.1 µm decreased while the

volume of pores < 0.1µm increased when PCE was added in the mortar.

Puertas et al (2005) observed that the presence of the PCE admixtures did not affect

paste strength at either 2 or 28 days. Nkinamubanzi and Aïtcin (2004) found that the

early age compressive strength of PCE concrete was less affected compared to

concrete with other superplasticizers due to the lower dosage required of PCE

admixtures to achieve similar slumps. Farrington (2007) found that PCE

59
Chapter 2 Literature Review

superplasticizer retarded early cement hydration according to isothermal calorimeter

measurement, but it did not retard setting and early strength development when used

at the same dosage in concrete. It is important to realize that the effects of admixtures

on properties of cement paste, mortar and concrete may not be identical.

2.7.2.2 Effect of SNF Admixtures

Gu et al (1994) investigated effect of SNF superplasticizers (both calcium and sodium

SNF) on pore structure development of cement pastes at various ages from 1 to 28

days using MIP. The difference in the porosity between the pastes with the SNF

superplasticizer and the control paste were significant at early ages but relatively

small at late ages. They found that the superplasticized pastes had larger mean pore

sizes than that of the control paste at various ages up to 28 days. The porosity results

were supported by TG analysis, from which they found less CH in the pastes with the

SNF admixtures than the control one.

However, it was found that the addition of SNF superplasticizer to the cement pastes

led to pore narrowing (Khalil, 1999; Hwang and Lee, 1989). The total pore system of

cement pastes with SNF superplasticizer contained mainly small and/or medium

capillary pores at 28 days. Gu et al (1982) also found that SNF superplasticized pastes

reduced pore volume and capillary pore size when SNF was used to reduce the water

requirement, which effectively resulted in a lower w/c ratio. Apart from the

performance of SNF, the discrepancy may be related to different mix proportions and

to the purpose of superplasticizer addition.

60
Chapter 2 Literature Review

2.7.2.3 Effect of LS Admixtures

Ramachandran (1995) reported that the total porosity of cement paste was increased

slightly in the presence of LS admixtures compared to that of the control paste. Pang

et al (2005a) found that the total pore volume of hydrated cement paste increased with

the increase in dosage of calcium LS. However, the average pore diameter of hydrated

cement paste with the calcium LS decreased, and the portion of pore with a diameter

over 30 nm decreased and the gel-pore with a diameter less than 10 nm increased

sharply. Pang et al (2005) investigated cement paste with a modified LS

superplasticizer and found that the results showed the same trend as that with the

calcium LS admixture - the portion of the gel pore increased and the average pore

diameter was refined. However, the incomplete growth of hydrate crystals and the

increase in pore volume in the cement led to reduction in the strength of cement paste

with the calcium LS within 28 days compared with the control cement paste.

2.7.2.4 Comparisons of Effect of PCE, SNF and LS Admixtures

Sakai et al (2006) observed that when SNF (without retardation effect) or LS (with

retardation effect) was added, almost similar pore structures were formed in the HCP.

They suggest that the difference in the pore structure of HCP with various types of

SPs is not related to the texture changes of the hydrates due to the retardation of

cement hydration, but related to the dispersion of the hydrates. Further, they found

that the volume of large pores of ≥ 0.1 µm in HCP with LS or SNF admixtures was

higher than that with PCE admixture cured for 28, 56 and 91 days. They suggest that

the size of the cluster of aggregated particles in the paste with PCE admixture may be

smaller than that with LS or SNF admixtures due to higher dispersing capacity of the

PCE admixture.

61
Chapter 2 Literature Review

2.8 Drying Techniques of Cement Paste and Testing Methods

2.8.1 Drying Techniques for Cement Paste

In determination of the constitutional water associated with the C-S-H gel, it is

important to differentiate the free water from that bound by the gel. Since the interest

is in the chemically bound water, it is necessary to remove the free water in the

cement paste by some drying techniques. In all the drying methods presented below,

the samples must be crushed so that the drying can be achieved within a reasonable

period of time. (Hewlett, 1998)

There are five commonly used drying techniques for cement paste research. They are

conventional oven drying, D-drying, vacuum drying, solvent exchange and freeze

drying. Each has its advantages and disadvantages.

2.8.1.1 Oven drying

Oven drying is the most common drying technique. The sample is dried at an elevated

temperature, most commonly at 105oC. It was observed that at this temperature, it had

a destructive effect on microstructure and induced microcracking, consequently

leading to an overestimation of total porosity (Gallé, 2001). Balasubramanian et al

(1997) suggests a combination of 50oC followed by 105oC oven drying, and they

believe that this two-step drying would allow reasonably fast drying but cause less

alteration to the microstructure of the cement pastes.

62
Chapter 2 Literature Review

2.8.1.2 D-drying

In D-drying procedure, samples are kept in a desiccator which is connected to a trap

cooled by a mixture of dry ice and ethanol at a temperature of -79 oC under vacuum

until the samples reach constant weights. The partial pressure of water vapor over the

ice precipitated in the trap is 5 x 10-4 torr (Hewlett, 1998). To put it simply, it is an

evacuation over dry ice (Ramachandran and Beaudoin, 1999).

2.8.1.3 Vacuum drying

Vacuum drying also has been used to dry the cement paste samples. However, vacuum

drying at low temperatures is a relatively slow process and therefore most suitable for

specimens older than 28 days, when not much free water is present in the paste

samples.

2.8.1.4 Solvent exchange

Solvent exchange is also known as solvent substitution or replacement. In this

technique, samples are to be immersed in a large volume of solvent, around 100:1,

due to the low solubility of water in most organic solvents (Ramachandran and

Beaudoin, 1999). The solvent immediately penetrates into the sample and replaces the

pore solution. The solvent should be renewed regularly. Normally, solvent exchange

is followed by oven drying (Ramachandran and Beaudoin, 1999). Researchers may

use different solvents and apply different temperatures. Some commonly used

solvents and references are summarized in Table 2-2.

This technique has been used by many researchers since it is simple and

63
Chapter 2 Literature Review

straightforward. It is worth noting that it may partially dehydrate phases such as C-S-

H and ettringite; however, such dehydration probably has little effect on the outward

morphology of the hydration products (Scrivener, 1997).

Solvent replacement method can significantly affect the amount of calcium hydroxide

depending on the type of solvent used. Day (1981) reported that methanol altered

sample composition by reacting with CH to form a carbonate-like product. Taylor and

Turner (1987) reported that some organic liquids, including methanol and acetone,

may react with the hydration products and thus may affect the test results. Chemical

interactions between solvents such as isopropanol, methanol, and acetone, and the CH

surface were observed (Beaudoin et al, 1998). Marchand (1993) used isopropanol and

methanol as solvents and found that the amount of water replaced by methanol was

greater than the volume of the evaporable water in the pore solution of the sample. He

suggests that methanol molecules may be able to penetrate into C-S-H structure and

replace some of the structural water.

2.8.1.5 Freeze drying

The underlining principle of freeze-drying is to rapidly freeze cement paste samples in

the drying process to minimize the growth of ice crystals in the samples. The rapid

freeze can be achieved by immersing small samples in Freon cooled by liquid

nitrogen. Subsequently, the ice is to be sublimated into gas directly under vacuum.

This technique is fast and induces minimal damage to the microstructure of samples.

However, thermal shock of immersing samples in liquid nitrogen may shatter

specimens (Scrivener, 1997).

64
Chapter 2 Literature Review

Gallé (2001) found that this technique did generate limited damage related to thermo-

mechanical stress; but it did not introduce capillary pores by the drying process.

Kjellsen and Diamond (2007) reported that freeze-dried and conventionally oven-

dried specimens at 60°C for more than 6 hours showed identical features on

microstructure.

2.8.2 X-Ray Diffraction (XRD)

X-ray diffraction is a powerful technique for identification of crystalline materials

because of the unique XRD patterns of individual crystalline phases. However, the

method is not useful for amorphous materials such as C-S-H in cement pastes.

In XRD spectra, the position of peaks is determined by the spacing of the

crystallographic planes according to Bragg’s law (Eq. 2-7) (1914):

nλ = 2d sin θ (2-7)

where n = an integer;

λ = wavelength of the radiation used;

d = spacing of the crystal planes; and

θ = angle of the diffraction peak.

The intensity of the peaks is affected by the types and positions of the atoms in the

crystal lattice according to the structure factor, which is beyond the scope of this

project. The peak intensity is also influenced by the quantity of a phase in a sample

and packing of the sample, and the latter is related to operators.

65
Chapter 2 Literature Review

Quantitative analysis of samples from XRD spectra can be performed by measuring

the intensity of the peaks acquired in the scan in comparison to that of an internal

standard. In samples containing a mixture of phases, the intensity of a peak is

proportional to the mass fraction and the atomic number of the phase responsible for

that reflection. The derivation of the equations relating intensity and mass are

described by Nuffield (1966). To consider a simple mixture with two components (“a”

and “b”), the ratio of the intensities of the two components is given by Eq. 2-8.

Ia W K a ρb
=K a and K = = const. (2-8)
Ib Wb Kb ρa

where I a , I b = intensity of peak associated with components ‘a’ & ‘b’, respectively;

Wa , Wb = mass fraction of components ‘a’ & ‘b’, respectively;

K a , K b = constant dependent on the instrumental arrangements and the nature

of the components ‘a’ & ‘b’, respectively; and

ρ a , ρ b = density of components ‘a’ & ‘b’, respectively.

This equation shows that the intensity ratio of two peaks from the components in a

sample is proportional to the mass ratio of the two components. This equation is valid

for any two components in a mixture with two or more components.

If a standard of known mass fraction is added to a sample, and K is known, the mass

fraction of the second component can be determined.

If a constant amount of internal standard, for example anatase, was added in

calibration samples, Eq. 2-8 can further be simplified as in Eq. 2-9.

66
Chapter 2 Literature Review

Wanatase I x
Wx = (2-9)
K I anatase

where Wx ,Wanatase = mass fraction of component x & anatase, respectively;

K = a constant, as in Eq. 2-8; and

I x , I anatase = peak intensity associated with component x & anatase, respectively.

The breadth of the peak is a function of instrumental parameters and the size of the

crystals in powder samples, but is constant for each sample (William et al, 2003).

Smaller crystals produce broader peaks, which in turn decrease the risk of preferred

orientation (William et al, 2003).

This is the simplest yet useful approach for semi-quantitative phase analysis of

cement and its paste (Mansoutre and Lequeux, 1996). It is possible to identify and

estimate anhydrated cement phases (e.g. C3S and C2S) and hydration products (e.g.

AFt and CH) in Portland cement pastes.

This method, however, has some limitations: (1) use of non-overlapping internal

standard for calibration makes the sample preparation tedious, (2) presence of

preferred orientation of crystals may affect results, and (3) overlapping peaks of

phases, e.g. alite and belite, makes the determination of the mass fraction of

individual phase difficult (Parrott et al, 1990; Mansoutre and Lequeux, 1996). The

C3S phase exhibits strong preferred orientation about the (001) 3 crystallographic

3
It is a representation of a plane in a structure, called Miller Index.

67
Chapter 2 Literature Review

direction. This affects not only its own relative peak intensities, but also the derived

peak intensities of overlapping phases, especially C2S (Scarlett et al, 2001).

Other methods include the partially or fully computerized Rietveld method, which has

been used on XRD data analysis (Wiles and Young, 1981). The method consists of

fitting the complete experimental diffraction pattern with a calculated profile and

background. It requires the knowledge of the crystal structure of all phases in the

samples to be analyzed. By using Rietveld method, common problems associated with

quantitative phase analysis by XRD such as peak overlapping and preferred

orientation can be minimized (Scrivenera et al, 2004).

Table 2-3 presents XRD powder patterns of a typical Portland cement (Taylor, 1997).

It can be seen that alite (A) has some reasonably strong peaks at 29.4° 2θ) (I = 60) and

51.7°/51.8° (2θ) (I = 33-35) which do not overlap with those from other phases. For

belite (B), there is no strong distinguishable peak that does not overlap with others.

The peak at 29.4° was suitable to identify Alite in cements and cement pastes. The

51.7°/51.8° peak can be used to identify alite phase when there is a relative high

quantity in a sample. According to Taylor (1997), When a cement paste sample is 28-

day and older, care is needed to avoid errors by using the 51.7°/51.8° peak.

2.8.3 Thermogravimetric Analysis (TG)

In thermogravimetric analysis, the mass changes due to dehydration or decomposition

of compounds are determined as the sample is heated at a uniform rate to high

temperatures. By determining the mass loss at a given temperature range where a

material dehydrates or decomposes, the quantity of a component in the material

68
Chapter 2 Literature Review

corresponding to the temperatures can be determined.

For a typical TG curve of a cement paste, the first mass loss, around 100-200 °C, is

mainly due to the dehydration of C-S-H and ettringite. The second major mass loss is

often observed at 450-550 °C which corresponds to the dehydration of CH. The third

mass loss is sometimes observed at 700-900 °C, corresponding to the decomposition

of calcium carbonate. The calcium carbonate can originate from Portland cements due

to the addition of limestone powder or from the carbonation of cement paste or a

combination of the two. Table 2-4 summarizes the temperature ranges in which

various reactions take place in cement paste samples.

69
Chapter 2 Literature Review

Table 2-1 Classification of Pores (information summarized from Mindess et al, 2003)

Table 2-2 Solvents and subsequent drying conditions used in solvent exchange

Subsequent drying
Authors (Year of Publication) Organic liquid used
conditions
Taylor and Turner (1987) /
acetone and diethyl ether(#) -
Rickert and Thielen (2004)
Vacuum dried at room
Ftikos and Philippou (1990) / temperature for 24h /
acetone and diethyl ether
Lilkov et al (1997) Vacuum Oven dried at
105°C for 3h
Parrott et al (1990) AR Grade methanol -
Marchand (1993) 1)isopropanol 2)methanol -
Gu et al (1994) / Oven dried at 105 °C for
acetone
Gruskovnjak et al (2006) 24h / Oven dried at 40 °C
ethanol and diethyl ether
Chotard et al (2001) -
(1:1 by volume)
Peschard et al (2004) &
ethanol -
Govin et al (2006)
Note: - means unknown or not mentioned by the authors
#
Authors may have used the synonyms of diethyl ether (ether or ethyl ether).

70
Chapter 2 Literature Review

Table 2-3 XRD powder pattern of a typical Portland cement (Taylor, 1997)

2θ° d (nm) Ipk Phases 2θ° d (nm) Ipk Phases


11.7 0.756 5 G 36.7 0.2449 6 A
12.1 0.731 6 F 37.4 0.2404 2 B
14.9 0.595 6 A 38.8 0.2321 12 A
20.7 0.429 7 G 39.5 0.2281 5 B
21.9 0.406 2 Al 41.3 0.2186 41 A, B
23 0.3867 7 A 41.6 0.2171 16 A, B
23.4 0.3802 3 B 44.1 0.2053 6 F
24.4 0.3648 3 F 44.5 0.2036 3 B
25.3 0.3520 4 A 44.7 0.2027 2 B
26.4 0.3376 2 B 45.8 0.1981 10 A, B
27.6 0.3232 2 B, A 47 0.1933 11 A
28.1 0.3175 4 A, B 47.4 0.1918 8 F
29.1 0.3069 5 G 47.8 0.1903 7 Al
29.4 0.3038 60 A 49.9 0.1828 5 A, F
30.1 0.2969 19 A 51.7 0.1768 33 A
31.1 0.2876 4 B, G 51.8 0.1765 35 A
32.2 0.2780 100 A, B, F 56 0.1642 2 A
2.6 0.2747 85 A, B 56.6 0.1626 18 A, B
33.2 0.2698 40 Al, A 58.7 0.1573 3 B, F
33.9 0.2644 23 F 59.4 0.1555 3 Al
34.4 0.2607 83 A, B 59.9 0.1544 6 A
Note: CuKα radiation; 2θ = diffraction angle; d = lattice parameter;
Ipk=relative peak intensity; A = Alite, C3S; B = Belite, C2S; Al = Aluminate,
C3A; F = Ferrite, C4AF; G = Gypsum; phases are given in decending order of
their major peak contributions.

Table 2-4 Summary of a typical cement paste TG graph

Authors, Year of Publication Temperature Observations


the evaporable water and part of the
Alarcon-Ruiz et al, 2005 30–105 °C
bound water escape
decomposition of gypsum and
Zhou and Glasser, 2001 110–170 °C
ettringite
Ramachandran and removal of loosely bound water and
< 200 °C
Beaudoin, 1999 firmly held water from C-S-H gel #
Khoury, 1992 180–300 °C the loss of bound water from C-S-H
General understanding 450–550 °C decomposition of calcium hydroxide
Grattan-Bellew, 1996 700–900°C decomposition of calcium carbonate
#
Note: DTA (differentitial thermal analysis) technique was used.

71
Chapter 2 Literature Review

Fig. 2-1 (a) SMF condensate, (b) SNF condensate, (c) Repeating unit of
lignosulphonate (LS) molecule (d) Molecular structure of polycarboxylate
(References: (c), Hewlett, 1998; (a, b, d), Borregaard admixture handbook, 2006)

Fig. 2-2 (a) Flocculation of cement particles resulting trapped water (b)
Deflocculation of cement particles upon adsorption of water reducing admixtures
(Law, 2004)

72
Chapter 2 Literature Review

Fig. 2-3 Repulsion of cement particles by (a) electrostatic repulsion


(b) steric hindrance (Ramachandran et al, 1998)

Fig. 2-4 Rate of heat evolution during hydration of Portland cement (Mindess et al,
2003)

73
Chapter 2 Literature Review

Fig. 2-5 Structure of cement pastes (Illston and Domone, 2001)

Fig. 2-6 Effect of water, water-reducing and air-entraining admixtures


on rheological behaviors of concrete (adapted from Gjørv, 1994)

74
Chapter 2 Literature Review

0.20
0.18

Cumulative Intrusion (ml/g)


0.16
Intrusion
0.14 dv/d(lnd)
0.12
0.10 dc
0.08
0.06
0.04
0.02
0.00 dt
0.001 0.01 0.1 1 10 100 1000
µm)
Pore Diameter (µ

Fig. 2-7 Critical pore and threshold pore diameters of MIP analysis

Fig. 2-8 Comparison of MIP and image analysis pore size distribution for the same
mix at 28 days with w/c = 0.40 cement paste (Diamond and Leeman, 1995)

75
Chapter 3 Experimental Details

Chapter 3 Experimental Details

3.1 Introduction

In the first part of this chapter, details of materials used in this project are described,

including but not limited to chemical, mineral compositions and physical properties of

cement, and the natures of various water reducing admixtures. Secondly, mix

proportions of cement paste and mortar are tabulated. Procedures of mix preparation

are also outlined. Finally, various test methods to achieve different objectives are

presented.

3.2 Materials

The materials used for all the experiments in this research project consist of cement,

water and fine aggregates, and six different water reducing admixtures (two regular

WRAs and four superplasticizers).

3.2.1 Cement and Water

Ordinary Portland cement (OPC) of ASTM Type I conforming to ASTM C150-02

requirements was used in the project. Table 3-1 summarizes its physical properties

and chemical and mineral composition. The chemical composition was determined by

76
Chapter 3 Experimental Details

ARL 9800 XP Sequential X-ray fluorescence spectroscopy (XRF), and oxide content

was calculated.

Deionized water, prepared by Barnstead Deionizer, was used for mixing cement

pastes. Normal tap water was used for mixing mortars.

3.2.2 Aggregates

Fine aggregate (sand) used in this project met the requirements of ASTM C136 -01,

and its grading curve is shown in Fig. 3-1. The fine aggregate sieve analysis was

conducted in accordance with ASTM C136-01 and the result is presented in Table 3-2,

together with the specific gravity and water absorption of the sand. The specific

gravity and absorption capacity of fine aggregate was determined according to ASTM

C128-97.

3.2.3 Water Reducing Admixtures

Six different admixtures including two regular WRAs and four superplasticizers were

investigated in this research project. Among these six admixtures, one was

polycarboxylate based (PCE), one was naphthalene based (SNF), and the other four

were lignosulphonate based (LS); and their characteristics are summarized in Table 3-

3. Polycarboxylate based admixture was obtained as solution with a concentration of

39.3%. The rest of the admixtures were initially in powder form and solutions were

made in laboratory.

The PCE and SNF based superplasticizers conform to requirements of ASTM C494

77
Chapter 3 Experimental Details

Type F high- range water-reducing admixtures (see Table 1-2). Lignosulphonate

admixture PLS is a highly purified sodium lignosulphonate product made from

modified softwood lignosulphonate. It has less than 0.5% of reducing substances.

Admixture UNA is a high molecular weight modified sodium lignosulphonate. It

contains approximately 1% of reducing substances, and has moderate set retardation.

Both the PLS and UNA admixtures can be used as ASTM Type F/G admixtures,

according to manufacturer.

Admixture BCS is a calcium lignosulphonate produced from hardwood. It contains

approximately 3% reducing substances. Admixture BCA is a calcium lignosulphonate

produced by fermentation of calcium lignosulphonate from softwood and contains

approximately 7% reducing substances. The raw material used for the production of

BCA is the same as that for the production of admixture UNA. Both BCS and BCA

admixtures are Type D water-reducing and retarding admixture.

For convenience and consistency, the powder admixtures were made into solution

before being added into mortar or cement paste. All powder admixtures were

dissolved in water to produce 30% solutions as recommended by the manufacturers

except for the PLS admixture. For the ease of handling, a solution of PLS admixture

with a concentration of 28% was made instead, due to the higher viscosity of PLS

solution at higher concentration.

The admixture powders were first dried in a 105 °C oven and cooled down to room

temperature in a desiccator. The admixture powder was then slowly added into

deionised water pre-heated to 50 °C, and stirred constantly for about 1 hour by a

78
Chapter 3 Experimental Details

magnetic stirrer to aid the dissolution of powder admixture until a uniform solution

was obtained. After admixture powder had been fully dissolved, tributylphosphate

(TBP, commonly known as defoamer) was added while stirring, at 0.5% by mass of

admixture powder, to the solution to control the air that may be entrained in pastes

and mortars. The admixture solutions were bottled with a tight cap and kept in a

refrigerator. Not to complicate the admixture solutions, no biocide was used to

lengthen the shelf life of the prepared admixture solutions. Instead, admixture

solutions were freshly made and kept in the refrigerator for no more than one month.

3.3 Mix Proportions of Cement Pastes and Mortars

Mix proportion of the cement pastes and mortars used for various tests was designed

to have similar workability. All the mixtures have the same proportions of cement,

water, and sand where used. Dosage of the admixtures was varied (Table 3-4) to

achieve yield stress of mortar at 75 ± 15 Pa. Based on tests on concrete, this yield

stress level of mortar will produce concrete with a slump of approximately 100 mm.

Two w/c ratios of 0.34 and 0.40 were used. For the w/c of 0.40, the effect of the six

admixtures were evaluated, whereas for the w/c of 0.34, only the effect of the four

superplasticizers were evaluated as the target yield stress was not achievable with the

regular water reducing admixtures at that low w/c. The air content of the concrete

measured right after mixing according to ASTM C138 ranged from 1.5 to 3%.

It is noted that the sand used in the project was oven dried to better control the

consistency of the mixtures. It is unlikely that full absorption will be achieved within

the time of mixing, but as the sand content is kept constant, the effective w/c ratio

remains similar for all mixes. The batching weight presented in Table 3-4 was in

79
Chapter 3 Experimental Details

saturated surface dry (SSD) condition. The water was compensated for the dry sand

based on its water absorption capacity of 1.3%.

Two mortar/cement paste mixtures with the same w/c ratios as those with the

admixtures were included as controls for comparison in experiments to determine the

effect of the admixtures on cement hydration. Without water reducing admixtures or

superplasticizers, however, these two control mixtures would have lower workability

compared with the corresponding cement pastes or mortars with the admixtures.

Therefore, no control cement pastes and mortars were used for evaluation of

workability, setting time, pore structure, and compressive strength of the materials as

these properties are affected by their initial workability.

3.4 Preparations for Cement Pastes and Mortars

Mortars and cement pastes were cast in the procedures described in Table 3-5.

3.4.1 Preparation for Cement Pastes

Various test methods to determine degree of cement hydration and pore structure

demand very careful sample preparations. The methods used for monitoring the

cement hydration include X-ray diffraction, thermogravimetry, and non-evaporable

water content. The pore structure was determined by MIP test. Care should be

particularly taken on XRD and TG samples since the studies are related to curing ages

and samples may be carbonated if exposed to atmosphere for an extended period of

time.

80
Chapter 3 Experimental Details

The above mentioned tests to monitor the cement hydration were conducted at various

curing ages of 2, 4, 8, 12 hours and 1, 3, 7, 28, 91 days. The MIP tests were conducted

on cement pastes of 1 day and older. The procedure of cement paste sample

preparation comprises the following steps:

1. Pastes were cast in a 5-quart (4.7-liter) Hobart mixer, according to the

procedure described in Table 3-5. The amount of cement used for each mix

was 1000 g.

2. After mixing, small plastic bottles were filled with the cement pastes. Bottled

pastes were sealed with preheated wax and were capped to minimize possible

carbonation.

3. Sealed sample pastes were rotated for the first 12 hours to minimise

segregation/bleeding and to ensure homogeneity of the cement paste samples.

After that, the sample bottles were stored in the sealed condition at about 30

°C for curing.

4. For samples cured up to 12 hours, cement hydration was stopped by acetone

(solvent exchange) thoroughly for 3 – 5 times, and the samples were ground at

the same time. With great rate of evaporation and the help of a cold-wind hair

drier, acetone took water away from sample fairly quickly and most samples

were surface dry – the color changed from dark to grayish white - within about

one hour. Samples were further ground till the particles were not coarser than

150 µm, and oven-dried at 105 °C for 2 to 4 hours before being stored in

capped glass bottles in a desiccator.

5. For samples that were 1 day and older, the cement pastes were removed from

the bottles at specified ages. Top and bottom part of the paste together with

that in contact with inner wall of the bottle were removed by a chisel. The

81
Chapter 3 Experimental Details

remaining sample was crushed into small pieces, and the samples were divided

into three parts for the tests of

a. Thermogravimetry/X-ray diffraction: samples were crushed and ground

into powder so that particle sizes were not larger than 150 µm;

b. Non-evaporable water: samples were not larger than 2-3 mm;

c. Mercury intrusion porosimetry: samples were carefully chiseled into

pieces of approximately 10 mm x 5 mm x 2 mm.

6. Samples in 5) were then stored in a vacuum oven at 60 °C till constant weight

(usually about 1 week). The acceleration of cement hydration is believed to be

minimal when the relative humidity is way below 80%, at which hydration

will slow down significantly (Mindess et al, 2003). Finally they were bottled

and stored in the desiccators.

It should be noted that grinding small samples into fine powder was not that simple.

Ball mill grinding is not recommended due to the consideration of contamination and

partial decomposition of hydrates by the heat generated from high-speed grinding. In

this study, the samples were first crushed into small pieces, and then manually ground

in an agate mortar.

3.4.2 Preparation for Mortars

Mortars were mixed in a 30-quart (28.4-liter) Hobart mixer. Each casting had the

same batch weight and thus roughly the same batch volume was produced. The

mortars were used for the tests on yield stress, plastic viscosity, flow table value, and

setting times of fresh mortars, and compressive strength of hardened mortars.

82
Chapter 3 Experimental Details

3.5 Test Methods and Analyses

Cement hydration is a chemical reaction and it is exothermal. Therefore, it can be

evaluated from several different angles. First, the difference in original and remaining

reactants gives the amount of cement compounds (C3S, C2S) reacted. The cement

compounds can be monitored by XRD techniques. Second, the amount of hydration

products (CH, CSH) qualitatively indicates that the reactions, although clinker

reactions are extremely complex. The analytical techniques and equipments used in

this study have been summarized in Table 3-6.

3.5.1 Heat Evolution of Cement Hydration

Heat of cement hydration was directly measured using a TAM Air isothermal

calorimeter (Fig. 3-2) by monitoring the heat generated from cement hydration for

cement pastes with and without the chemical admixtures. Heat generated from cement

pastes with w/c of 0.34 and 0.40 were determined.

There were eight channels in the TAM Air calorimeter, and each channel was

constructed in twin configuration with one side for the sample and the other side for

an inert reference. In the current project, water was used as reference material. During

experiment, both the sample and reference materials were held in 20 ml sealed

ampoules. Each side of the calorimetric channel was then covered with a removable

cylindrical metal heat sink plug to prevent thermal disturbance from the circulating air.

The twin configuration of the sample and reference within a channel allowed the heat

flow from the active sample to be compared directly with the heat flow from the inert

reference. The voltage difference was a quantitative expression of the overall rate of

83
Chapter 3 Experimental Details

heat production in the sample. The rate of heat production, or heat flux, is defined as

the rate by which heat evolved by the sample. From the power output data and with

some appropriate conversions (Eqs. 3-1 and 3-2), the amount of heat produced in the

sample (energy output) could then be derived.

The data extracted from the calorimeter was power output (rate of heat evolution or

heat flux) generated from the cement hydration process. The differences in the sample

mass between the samples can be accounted for by normalizing the power output

generated from each channel with its respective sample mass. For every data point,

the normalized power output was calculated according to Eq. 3-1.

Pn = Po m (3-1)

where Pn = normalized power output (rate of heat evolution) obtained from cement

hydration, mW/g;

Po = power output from cement hydration obtained from calorimeter, mW; and

m = sample mass of the respective channel measured, g.

The energy curve represents the amount of heat liberated from the cement hydration

per gram of the sample. For every normalized data point, the energy conversion at that

point was calculated as Eq. 3-2.

E = Pn ⋅ t 1000 (3-2)

where E = normalized energy output (heat evolution) from cement hydration, J;

84
Chapter 3 Experimental Details

Pn = normalized power output from cement hydration, mW/g; and

t = time interval, seconds, in this case, t = 5 min = 300s.

The isothermal calorimeter was calibrated and conditioned at 30 °C for a day before

experiments and the amplifier range was set up to 600 mW. Before the experiments,

all the materials, mixing utensils, and sample ampoules were pre-conditioned in a 30

°C oven for at least 12 hours. The preconditioned cement was added into deionised

water, and hand mixed for about 1 minute. The cement paste sample of 10 ± 2 g

(sample masses were recorded) was then transferred into the sample ampoule. After

capping, the sample and reference ampoules were inserted into the calorimeter. The

calorimeter started to record heat 10 minutes after the cement was in contact with

water. Because of this procedure, the heat generated initially during mixing of the

cement and water was not captured.

The setting file defined the sampling interval and maximum number of samples

recorded at a specific temperature. In this project, the sampling interval was set at 5

minutes and the heat generated from the samples was monitored continuously for 3

days. It should be noted that machine sensitivity often limits the measurement to

about 7 days duration, beyond that the signal becomes virtually indistinguishable from

the background noise (NIST, 1996, online access 20074). Due to the background noise

and time required for the inserted sample ampoules to reach equilibrium with the

reference, the data extracted for the first 10 minutes was not used. In other words,

only the data collected 20 minutes after the cement in contact with water was used for

analysis.

4
http://ciks.cbt.nist.gov/bentz/phpct/database/thermal.html

85
Chapter 3 Experimental Details

3.5.2 Degree of Cement Hydration

There are many ways to determine the degree of cement hydration, either directly or

indirectly. The tests employed in this study were three commonly used ones, namely,

XRD analysis, TG analysis and non-evaporable water content. The XRD method was

used to monitor the reduction of cement clinkers such as C3S with the progress of

cement hydration. The TG method was used to monitor the increase in the calcium

hydroxide with the cement hydration. The non-evaporate water was used to monitor

the increase in the water associated with the formation of hydration products such as

calcium silicate hydrates and calcium hydroxide.

3.5.2.1 X-ray Diffraction (XRD)

The X-ray diffraction analyses were carried out using Shimadzu XRD-6000

o
diffractometer (Fig. 3-3) with Cu Kα ( λ = 1.54056 A ) radiation at 40 kV and 30 mA.

The XRD scan was between 5° to 60° (2θ) with a scan speed of 0.5°/min (0.02° step

and 2.4 seconds preset time or counting time). The range of 2θ was selected because it

contained major peaks of interested phases, including possible ettringite AFt and

monosulfoaluminate AFm (Yousuf et al, 1995; Williams et al, 2003). The X-ray

diffractometer was calibrated using 99.99% pure silicon. It was operated under the

following slit specifications: divergence slit 1°, scatter slit 1° and receiving slit 0.3

mm. Each sample was tested for three times consecutively and the average peak

intensities were taken for further analysis.

The results from XRD analyses were presented in terms of the intensity ratios

between the phases of interest and the reference material anatase (TiO2, 2θ = 25.3 o),

86
Chapter 3 Experimental Details

which was 10% by mass blended in the cement paste samples. The ratios indicate the

relative but not the actual amount of the phases of interest.

To link the X-ray diffraction peak intensity ratios to the actual C3S content in the

sample materials, a calibration chart of C3S content in materials of interest (Fig. 3-4)

was produced using the cement (Table 3-1) and mixtures of the cement with sodium

carbonate (Na2CO3, which has no overlap peaks with the cement). The C3S peak at 2θ

= 29.4 o was used for this calibration curve. The mixtures of the cement with sodium

carbonate were used so that the calibration chart would cover a wider range of C3S

content (Table 3-7) in the sample materials.

Seven data points were collected from the cement and its mixtures with sodium

carbonates. They were well spread across the range concerned. A linear function,

which exhibits the relationship between the XRD intensity ratio of C3S and the

reference anatase and the actual amount of C3S presented in the samples, is shown in

Eq. 3-3.

(I )
C3 S 2θ = 29.4°
C 3 S % = 68.575 ⋅ (3-3)
(I anatase )2θ = 25.3°

With the calibration chart, the approximate C3S content in the sample can be

calculated according to the XRD peak intensity. It is recognized that Bogue

calculation of C3S content is approximate, and possible errors may be introduced by

using this calibration chart. Also, XRD results have high variability depending on

particle size and sample packing. Nevertheless, this may serve as semi-quantitative

analyses for relative comparison.

87
Chapter 3 Experimental Details

3.5.2.2 Thermogravimety Analysis (TG)

Thermogravimetric analysis is used to monitor the increase in calcium hydroxide with

the progress of cement hydration. In a TGA (Fig. 3-5), the weight loss of a sample is

recorded while it is being heated at a uniform rate in a nitrogen environment. The

weight loss over specific temperature ranges provides information on the dehydration,

decomposition or phase changes of samples. Although the sample may be

decomposed at a temperature which is the characteristic of the compound, the shape

of the decomposition curve may be affected by many factors. Haines (2002)

recommended carrying out experiments with high thermal capacity furnaces, with

small and lightweight crucibles and using a low rate of heating, and small sample

sizes.

Calcium hydroxide decomposes at around 450 – 550 oC according to the chemical

reaction Eq. 3-4.

−
o
Ca(OH) 2 450 → CaO + H 2 O(g)
550 C
(3-4)

By determining the mass loss due to the loss of water in the decomposition of CH, the

amount of the CH and thus the degree of cement hydration can be determined

(Ramachandran, 2003).

Carbonation is a chemical reaction shown in Eq. 3-5.

Ca(OH) 2 + CO 2 → CaCO 3 + H 2 O (3-5)

The effect of carbonation on the amount of CH can be determined from the amount of

88
Chapter 3 Experimental Details

carbon dioxide given off at higher temperature according to Eq. 3-6.

o
−
CaCO 3 700 → CaO + CO 2
900 C
(3-6)

Samples used in the TG test were prepared as described earlier. The TG was carried

out using a Linseis L81-II thermogravimetric analyzer. For each test, approximately

100 mg of powdered sample was heated from room temperature to 950 oC at a rate of

10 oC/min in a nitrogen environment. A ‘‘blank test’’ (calibration without specimen)

showed that a fictitious mass gain (0.001 mg/°C) was recorded by the apparatus

during heating. It was recorded as a “zero” file. The final results presented in the

following chapter take this into account.

Figure 3-4 shows the calculation of mass loss from a TG curve (Haines, 2002). First,

two tangent lines corresponding to the initial and final baselines were drawn. Second,

after finding the largest slope between the temperatures before and after the mass loss,

a tangent line was drawn corresponding to this slope. This intersected with the earlier

two tangent lines. The difference between the two intersections was used as the mass

loss from the decomposition of a particular crystalline matter. It should be noted that

the above process is subjective and depends on the experience of the experimentalist

(Williams et al, 2003).

3.5.2.3 Non-Evaporable Water (NEW) Content

Non-evaporable water in cement paste samples includes chemically combined water

which is associated with hydration products such as CH, C-S-H etc. The NEW was

89
Chapter 3 Experimental Details

determined by mass loss determined by (1) igniting samples in a furnace at 950oC and

(2) thermalgravemetric analysis.

In the first method, clean crucibles weighing around 30g were burned in a muffle

furnace (Fig. 3-7) at 950 oC for 2 hours and left in the furnace till the temperature

went down to around 120 oC. The crucibles were then cooled to room temperature in a

desiccator. Burning the empty crucibles and handling them with tongs were to

minimize errors - some organisms that might be stuck on the crucible surfaces.

Approximately 50 g representative dried sample was weighed in the burned crucible.

The crucible containing the sample was ignited at 950 °C in the furnace for 3 hours.

After it was cooled down to around 120 oC, the crucible containing the sample was

placed in the dessicator and cooled to room temperature. The mass of the samples was

determined again. The mass loss was calculated from the difference before and after

the ignition. The non-evaporable water content was calculated according to Eq. 3-7.

W1 − W2
Wn = − LOI (3-7)
W1

where Wn = non-evaporable water content per gram of dried cement paste sample;

W1 = mass of the sample before the ignition;

W2 = mass of the sample after the ignition; and

LOI = loss on ignition of anhydrated cement in one gram of dried paste sample.

The TG curve provides mass loss during the heating, which is also associated with the

loss of water from the dehydration and decomposition of the hydration products. In

90
Chapter 3 Experimental Details

this thesis, the NEW determined from TG is referred to as NEW from TG, whereas the

NEW determined from the furnace ignition is referred to as NEW from furnace.

3.5.3 Workability Retention of Mortars

The workability retention of fresh mortars with and without WRAs was monitored by

the yield stress, plastic viscosity, and flow value. The initial test started at 10 minutes

from the addition of water into the mix, followed by tests at 30 and 60 minutes. The

flow table value was determined 2-3 minutes after the determination of the yield

stress and plastic viscosity.

The rheometer used to determine the yield stress and plastic viscosity of the mortars

in this study was a coaxial cylinder rheometer, ConTec BML Viscometer 3 (Fig. 3-8).

Sample container is placed on the plate during experiment and acts as the outer

cylinder. The inner cylinder has three components. The upper unit measures the torque.

The lower unit is used to eliminate the shear influence from the bottom of the outer

cylinder. The top ring is to keep a constant height where the torque is measured. The

ribs of both inner and outer cylinders are to reduce the tendency of slippage.

During the test, the outer cylinder rotates, at different rotation velocities N (angular

velocities ω 0 = 2π ⋅ N ). Torque (T) required to keep the inner cylinder stationary is

measured and registered.

Once torque-rotation speed relation (Eq. 3-8) is obtained from T-N curve, Reiner-

Riwlin Equation (Eq. 3-9) (Reiner, 1949) for coaxial cylinder viscometer can be used

91
Chapter 3 Experimental Details

to determine the yield stress τo (Eq. 3-10) and plastic viscosity µ (Eq. 3-11) for

Bingham Model. The effective shear rate γ& , which varies with the position r

measured from the center of the cylinders in the annulus, may be calculated from τ 0

and µ (Eq. 3-12). Assuming no plug flow occurs, Eqs. 3-9 and 3-12 may be combined

to yield Eq. 3-13. Figure 3-9 illustrates the transformation from torque-rotation speed

to shear stress-shear rate.

Equations 3-8 to 3-12 are summarized as follows (note that the symbols used in Eqs.

3-10 to 3-12 are the same as the ones defined in Eqs. 3-8 and 3-9):

T = g + hN (3-8)

where T = torque, Nm;

N = rotation speed, rps;

g = flow resistance, Nm; and

h = relative viscosity, Nm.s.

 Ro 
τ 0 ln  + ω 0 µ
R
 i
T = 4πH (3-9)
1 1
2
− 2
Ri Ro

where H = constant height between inner cylinder bottom and top ring bottom, m;

R0 = radius of the outer cylinder, m;

Ri = radius of the inner cylinder, m; and

ω0 = angular velocity = 2πN, rps.

92
Chapter 3 Experimental Details

 1 1 
 − 2
 Ri2
R o 
τ0 =  g (3-10)
 Ro 
4πH ln 
 Ri 

 1 1 
 − 2
 Ri2
R o 
µ= h (3-11)
8π 2 H

1 T 
γ& =  −τ 0  (3-12)
µ  2π ⋅ r H
2

−1
2  1 1   τ 0  Ro   τ
γ& = 2  2 − 2  ln  + ω 0  − 0 (3-13)
r R Ro  µ R   µ
 i   i 

The parameters used in the test (Table 3-8) were selected to produce smoothly the T-N

curve (Wallevik, 2003). Assume the position “r” in Eq. 3-13 equals the radius of the

inner cylinder; the shear rates experienced by the inner cylinder were in the range of 5

- 50 s-1. The rotation speed was from 0.5 rps to 0.1 rps, which means that yield stress

and plastic viscosity were determined from a ramp down T-N curve (see Fig. 3-10).

The rheological properties of cement paste, mortar, and concrete are strongly

dependent on the shear history (Wallevik, 2003). The mixing procedures described in

Table 3-5 were closely followed. Between any two tests using the BML viscometer,

the mortar mixture was left in the Hobart mixer bowl and was covered with a plastic

sheet to minimize water loss from evaporation. Before testing, the mortar was pre-

sheared at Speed 1 (139 rpm) for one minute to homogenize the samples.

Flow value of the same mortar was determined (see Fig. 3-11) according to ASTM

93
Chapter 3 Experimental Details

C230M-98 except that the value was measured after 10 drops instead of the standard

25 drops of the plate with the sample mortar. It was done to avoid overflow of the

mortars from the plate. The plate had a diameter of 250 mm. The flow cone used had

top diameter of 70 mm and bottom 100mm.

3.5.4 Setting Time of Mortars

The most widely used method for determining setting time of concrete (Fig. 3-12) is

ASTM C403 – Standard test method for time of setting of concrete by penetration

resistance (Lamond and Pielert, 2006). As the title suggests, the setting times are

determined from the changes in the penetration resistance of a sample as a function of

time. A sample of mortar is obtained either by sieving the fresh concrete on a 4.75 mm

sieve to remove the coarse aggregates or by preparing the mortar directly (Lamond

and Pielert, 2006). In this project, mortars were prepared directly to determine the

setting times. The sample mortar was placed into a 150-mm cubic steel mould and

stored in a room with a constant temperature at 30 °C. The mortars were covered with

a plastic sheet to prevent moisture loss, and excessive bleeding water on the top

surface was drawn carefully into a disposable syringe.

According to the ASTM standard, the initial and final setting times are defined when

the penetration resistance reached 3.5 MPa (500 psi) and 27.6 MPa (4000 psi),

respectively.

3.5.5 Pore Structures of Pastes

Mercury intrusion porosimetry test was used to determine total porosity and pore size

94
Chapter 3 Experimental Details

distribution of the cement pastes at 1, 3, 7, 28 and 91 days. The test was performed on

a Micromeritics Autopore WIN9400 Series mercury porosimeter (Fig. 3-13) with a

maximum pressure of 412.5 MPa. The minimum pore access diameter reached under

the maximum pressure was about 3.8 nm assuming a contact angle of 141.3° and a

mercury surface tension of 0.485 N/m (Ramachandran and Beaudoin, 1999)

Approximately 1.5 grams of hardened cement paste samples (2-3 pieces) were used

for each test. One MIP test was conducted on each sample. For each series of the paste

samples, one or two testing ages were randomly selected and samples were repeated

to ensure the repeatability of the test.

3.5.6 Compressive Strength of Mortars

The compressive strength of the mortars was determined at the ages of 1, 3, 7, 28 and

91 days in accordance with ASTM C109M. For each mortar mixture, fifteen 50 mm

cubes were cast, i.e. three cubes for each testing age. The cubes were covered with

plastic sheet and left in laboratory for the first 24 hours. They were demoulded after

that and cured in a fog room with a temperature of about 28 °C till their testing ages.

The loading rate employed for the test (Fig. 3-14) was 1670 N/s (100 kN/min), which

is within the range of 900 to 1800 N/s specified by the ASTM standard.

95
Chapter 3 Experimental Details

Table 3-1 Chemical & Mineral Compositions and Physical Properties of Cement Used

Properties ASTM C150


Initial Setting Time, min 180 ≥45
Physical Properties
Final Setting Time, min 210 ≤375
Blaine Fineness, m2/kg 363 ≥280
Calcium Oxide, CaO 63.19
Silica, SiO2 20.26
Aluminium Oxide, Al2O3 4.31
Iron Oxide, Fe2O3 3.61
Magnesia, MgO 3.25 ≤6.0
Chemical
Sodium Oxide, Na2O 0.25
Composition*, %
Potassium Oxide, K2O 0.30
Sulphuric Anhydride as SO3 2.06
Loss on Ignition (LOI) 2.53 ≤3.0
Insoluble Residue 0.26 ≤0.75
Total Alkalinity as Na2O+0.658K2O 0.35 ≤0.60
Mineral Tricalcium Silicate, C3S 63.2
Composition Dicalcium Silicate, C2S 10.4
According to Tricalcium Aluminate, C3A 5.3
Bogue Calculation,
% Tetracalcium Alumninoferrite, C4AF 11.0
* The chemical composition was determined by ARL 9800 XP Sequential X-ray
fluorescence spectroscopy (XRF), and oxide content was calculated.

Table 3-2 Physical properties and sieve analysis of sand

Sieve Size % retained by mass


4.75 mm 2.6
2.36 mm 11.2
1.18 mm 34.9
600 µm 72.4
300 µm 92.2
150 µm 98.5
Fineness Modulus 3.12
Absorption Capacity 1.30%
Specific gravity, SSD* 2.65
SSD* = saturated surface dry

96
Chapter 3 Experimental Details

Table 3-3 Characteristics of admixtures used in the project


Soluble Molecular Molecular
Reducing
Water SO4, % of Weight Weight
Type Notation substances,
Reduction dry Distribution, Distribution,
%
admixture Mw Mn
&
polycarboxylate PCE > 30% 0 N. A. N. A. N. A.
naphthalene SNF 15 ~ 25% 0 N. A. N. A. N. A.
purified
PLS 25% 0.5 0.2 41800 10650
lignosulphonate
modified
UNA 15 ~ 20% 1 0.7 47800 5800
lignosulphonate
Ca-
BCS 8 ~ 10% 3 0.2 5700 1600
lignosulphonate
Ca-
BCA 8 ~ 10% 7 0.5 21100 3050
lignosulphonate
&
N. A. = not available

Table 3-4 Mix proportion of mortars to achieve an initial yield stress of 75 ± 15 Pa


Sand
Concentr- Dosage, % Cement, Total
w/c Mix Admixture (SSD),
ation, % (sbwc^) kg water*, kg
kg
R# - - -
A PCE 39.3 0.13
B SNF 30 0.30
0.40 C PLS 28 0.28 8.50 3.400 14.97
D UNA 30 0.35
E BCS 30 0.60$
F BCA 30 0.45
&
V - - -
I PCE 39.3 0.17
0.34 II SNF 30 0.39 9.60 3.264 14.91
III PLS 28 0.35
IV UNA 30 0.45
R# = Control mix for w/c = 0.40 as reference
V& = Control mix for w/c = 0.34 as reference
sbwc^ = solid by weight of cement, %
0.60$ = maximum recommended dosage used, since even at 1% dosage it failed to
achieve the target yield stress
Total water* includes the water presented in admixture solutions

97
Chapter 3 Experimental Details

Table 3-5 Mix procedures of mortars and pastes

Time, Action
min Mortar (Speed 1 at 139rpm) Paste (Speed 1 at 139rpm)
dry mix of cement,
-1 dry mix of cement and sand
pour admixture solution into water
0 addition of water addition of water and admixture
1 addition of admixture -
3 idle for one minute to scrape deposit on the mixer walls
4 mix for another 1 min
5 start test seal in plastic bottles until specific ages

Table 3-6 Analytical techniques and equipment used in this study

Properties of paste or mortar investigated Techniques and equipment used


TAM Air Isothermal Calorimeter
Heat evolution
(multi-channel)
Calcium hydroxide, Shimadzu X-ray diffractometer (XRD-
Tricalcium / dicalcium silicate content 6000)
Calcium hydroxide and
Linseis L81-II Thermogravimetry (TG)
Non-evaporable water content
Mass loss at a high temperature of 950 oC
Non-evaporable water content
(Lenton Furnace)
Yield stress and plastic viscosity
ConTec BML Viscometer 3 (coaxial)
change with time
Flow value change with time Motorized flow table (H-3624)
Setting times Humboldt Penetrometer (H-4133)
Mercury intrusion porosimetry (MIP),
Porosity, pore structure
Micromeritics Autopore WIN9400 Series
Compressive strength Automax5 automatic compression tester

98
Chapter 3 Experimental Details

Table 3-7 Seven samples used to produce C3S calibration chart

% of the cement 100 95 80 65 50 35 20

% of Na2CO3 0 5 20 35 50 65 80

IC3S/ITiO2 (2θ=29.4°)* 0.96 0.84 0.75 0.60 0.44 0.31 0.18

C3S in sample 63.2** 60.0 50.6 41.1 31.6 22.1 12.6


* determined by XRD
** based on Bogue calculation

Table 3-8 Process parameters set on BML Viscometer 3 for determination of the yield
stress and plastic viscosity

Cylinder dimensions Run time Parameters


Height of inner cylinder, m 0.115 Max. rotation velocity, rps# 0.5
Radius of inner cylinder, Ri, m 0.085 Min. rotation velocity, rps 0.1
Radius of outer cylinder, Ro, m 0.1 No. of T/N points* 10
Beater control Transient interval, sec 2
Beater penetration time, sec 5 Sampling interval, sec 1
Penetration speed, 0.1-1 0.5 No. of sampling points 50
Note: rps# stands for rounds or revolutions per second
T/N point* was the average of the lowest 10 out of the 50 sampling points

99
Chapter 3 Experimental Details

100
ASTM Lower
90 ASTM Upper
Sand Used
80

70
% Passing by Mass

60

50

40

30

20

10

0
4.75 2.36 1.18 0.6 0.3 0.15
Standard Sieve Size, mm

Fig. 3-1 Grading curve of fine aggregate (sand) used

(a) (b)

Fig. 3-2 (a) Isothermal calorimeter (b) Sample loading and unloading

100
Chapter 3 Experimental Details

Fig. 3-3 Schematic diagram of an X-ray diffractometer

70

60
C3S % in material of interest

50

40
C3S = 68.575IC3S/Ianatase
2
R = 0.99
30

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
o
XRD intensity ratio, IC3S @29.4 /Ianatase

Fig. 3-4 Calibration chart of C3S in materials of interest

101
Chapter 3 Experimental Details

Fig. 3-5 Schematic diagram of a thermogravimeter

Fig. 3-6 Determination of mass loss from a thermogravimetry curve (Haines, 2002)

102
Chapter 3 Experimental Details

Fig. 3-7 Schematic diagram of a furnace

Fig. 3-8 Schematic diagram of the BML-Viscometer (Source: ConTec Ltd., 2003)

103
Chapter 3 Experimental Details

(a) (b)

Fig. 3-9 The relation between (a) torque - rotation speed and
(b) shear stress - shear rate (Bingham model)

Fig. 3-10 A typical ramp down T-N curve from test on mortar by BML Viscometer

Fig. 3-11 Schematic diagram of flow table set-up


(http://www.durhamgeo.com/testing/concrete/cement-flowtable.htm)

104
Chapter 3 Experimental Details

Fig. 3-12 Schematic diagram of a penetrometer

105
Chapter 3 Experimental Details

Fig. 3-13 Schematic diagram of a mercury intrusion porosimeter

Fig. 3-14 Schematic diagram of a compressive strength tester

106
Chapter 4 Results and Discussion

Chapter 4 Results and Discussion

This chapter presents results and discusses effects of the admixtures on the cement

hydration, workability, setting time, pore structure, and strength development of

cement pastes and mortars. The dosages of the water reducing admixtures and

superplasticizers were determined so that the initial (at 10 min after the cement came

in contact with water) yield stress of the mortars was 75 ± 15 Pa for both w/c ratios of

0.34 and 0.40 with the exception of that with the BCS admixture as it failed to achieve

the target yield stress even though a dosage of 60% more than the maximum

recommended dosage was used. The same dosages of the admixtures based on solid

by weight of cement were correspondingly used for the cement paste mixtures.

Since the setting time (determined by penetration resistance), pore structure, and

compressive strength of the cement pastes and mortars are related to the compaction,

which is influenced by the workability of the cement pastes and mortars, the control

pastes and mortars were not included for comparison in Sections 4.3 – 4.6 as the

target yield stress of mortars could not be achieved without the admixtures.

4.1 Heat Evolution of Cement Hydration

As illustrated in Fig. 2-4, the heat curve of OPC generally exhibits one peak “A”

107
Chapter 4 Results and Discussion

before induction and two peaks “B” and “C” after that. Peak “B” normally

corresponds to C3S hydration and peak “C” to C3A hydration, but they may be

switched depending on the SO3 content in OPC and w/c ratios (Sandberg, 2004).

The SO3 content of the cement used was 2.06%, and the heat curve of the cement

paste with w/c of 0.40 without any admixture was shown in Fig. 4-1. As mentioned

earlier in Chapter 3, Peak “A” was not obtained due to the limitation of the equipment

used. From Fig. 4-1, it can be observed that peak “C” was higher than peak “B”. In

order to confirm that peak “C” was indeed related to C3A reaction, SO3 in terms of

gypsum was increased in the cement paste, to 2.20%, 2.35%, 2.50%, 3.00% and

3.50% by the total weight of cement and additional gypsum. As seen in Fig. 4-1, the

position and magnitude of the peak “B” was relatively unchanged with the increase in

the SO3 content, whereas the peak “C” changed with the increase in the SO3 content,

and eventually flattened. This confirms that the peak “C” was related to C3A reaction.

As shown in Figs. 4-2 and 4-3, all the admixtures, regardless of regular water

reducing admixtures or superplasticizers, retarded cement hydration. However, the

degree of retardation varied with different admixtures. From the peak intensity and

time of peak occurrence in the heat curves, the cement paste mixtures may be

categorized into four groups, (1) control and SNF mixtures (little or no retardation); (2)

Mixture with PCE (weak retardation); (3) Mixtures with lignosulphonate

superplasticizers PLS and UNA (medium retardation) and (4) Mixtures with

lignosulphonate Water reducing admixtures BCS and BCA (strong retardation).

The heat curve of the mixtures with SNF did not differ from the control mixtures

108
Chapter 4 Results and Discussion

significantly at both w/c ratios of 0.34 and 0.40. The hydration was delayed slightly,

shown by a small shift in the heat curves to the right.

For the mixtures with lignosulphonate-based admixtures (PLS, UNA, BCS and BCA),

the peak “C” was not visible in the heat curves shown in Figs 4-2 and 4-3. Although

the soluble SO3 contents in admixtures (Table 3-3) are not the same, the differences

are too small to cause a significant change in the total SO3 content available in the

system. These mixtures also showed reduction in the intensity of the silicate heat

peaks and delayed time of the peaks compared with the control mixtures.

It was observed that the peak intensities of the LS based admixtures were lower than

that of the control and the PCE/SNF superplasticized pastes. Various mechanisms of

set retardation have been proposed, which was summarized in a review paper by Zhor

and Bremner (1999).

1. Hansen (1959) proposed that the adsorption of LS molecules onto the surface of

anhydrated cement created a barrier to cement hydration.

2. Daughterty and Kowalewski (1968) proposed that chelation of functional groups of

admixtures to metal ions could be an important factor in the mechanism of retardation.

3. Watanabe et al (1969) proposed that the precipitation of calcium salts on

anhydrated cement caused the retardation.

4. In a review paper, Young (1972) presented a nucleation model which was proposed

by Greening. He reported that the nucleation of crystalline CH was inhibited by

soluble silica and thus proposed that this inhibition was a self-retarding feature of C3S

hydration. This principle was later supported by Berger and McGregor (1972). In their

109
Chapter 4 Results and Discussion

paper, it was found that the morphology and number of CH crystals was clearly

influenced by LS admixtures.

5. Young (1972) further proposed a possible mechanism consolidating the earlier

proposed mechanisms.

From the results, it is unlikely to be the adsorption or precipitation mechanisms alone

as neither could explain the initial acceleration activity that occurred during the first

hour for these two admixtures. The observation probably can be explained by the

consolidated mechanism with a combination of (1) to (4). After the initial activity, the

admixtures were incorporated into the structures of the hydrates and removed from

the solution and the retardation of C3S predominates, which was governed by the

effect of admixture on CH nucleation. Based on the observation, it is quite reasonable

to attribute the peak suppression to the LS admixtures. However, the exact interaction

between LS molecules and CH is not clear.

The heat curves of the mixtures with PCE were between those with SNF and LS-

based admixtures.

Comparing the heat curves of the cement pastes with w/c ratios of 0.34 and 0.40, the

control mixture and mixtures with the PCE and SNF admixtures showed both Peaks

“B” and “C”. The two peaks may look quite the same in shape, but differed in the

time of appearance. Summarized in Table 4-1, it is found that for control and SNF,

PCE superplasticized pastes, the durations between the C3S and C3A peaks were

shorter for w/c of 0.34 than that for w/c of 0.40.

110
Chapter 4 Results and Discussion

The length of induction period in the heat curves indicates the degree of retardation.

For both w/c ratios, the induction periods are in the ascending order of CTR < SNF <

PCE < PLS < UNA. For both w/c ratios, it shows that all admixtures delayed cement

hydration and that the four superplasticizers delayed cement hydration in the same

sequence, i.e. SNF < PCE < PLS < UNA. The SNF based SP had the least retardation

while the modified LS based SPs had the most significant retardation on cement

hydration among all the superplasticizers. The two regular WRAs delayed the onset of

acceleration period to as long as about 12 hours. Regardless of w/c ratios, all the LS

based admixtures induced a major reduction in the heat of hydration at an early age.

For w/c ratio of 0.40, their heat evolved had not reached that of the control mixture at

3 days. However, for w/c ratio of 0.34, they caught up with the control mixture after

48 hours and overtook it after that. The retardation effects of the admixtures were also

reflected in the cumulative heat curves shown in Figs. 4-4 and 4-5.

Figures 4-4 and 4-5 show that with the increase in the cement hydration and time, the

difference in the total heat released from the cement pastes with different admixtures

was reduced. Figure 4-4 shows that the differences of total heat evolved from

superplasticized cement pastes decreased with time, being 15.9% and 6.8% at 1 day

and 3 days. A similar trend is seen in Fig. 4-5, where the differences decreased from

54.2% at 1 day down to 12.3% at 3 days. This indicates that the total heat evolutions

for different mixes are getting closer with time, which would be the same eventually,

given the w/c ratio is the same for mixes. According to Ridi et al (2003), water

reducing admixtures only accelerate or delay cement hydration, but they did not

increase or decrease the total cement hydration.

111
Chapter 4 Results and Discussion

It is interesting to note that the heat evolved from the cement pastes with the two

regular LS based admixtures BCA and BCS in the first hour was significantly higher

than that of the other mixtures. Tagnit and Sarkar (1990) reported that lignosulphonate

based water reducing admixtures may interfere with the dissolution of sulfates in

cements. The greater heat evolved at early ages for the mixtures with the BCA and

BCS might be related to the reduced SO42- in solution and thus increased the reaction

of C3A.

The 3-day total heat evolved (Fig. 4-6) for the two control mixtures were similar

although lower w/c ratio had a higher rate of hydration and heat evolution initially.

However, comparing the heat curves in Figs. 4-4 and 4-5, it seems that all the pastes

with the admixtures had higher heat evolution at w/c ratio of 0.34 before 4 hours and

after 48 hours. The pastes with the w/c ratio of 0.34 took a shorter time to accumulate

a given amount of heat at early hours. From 48 to 72 hours, it is very clear that

cumulative heat curves of lower w/c ratio pastes were above the control while those of

higher w/c ratio below the control. This is likely due to the fact that the amount of

cement used in w/c 0.34 was 12% more than that used in w/c 0.40. It is seen in Table

3-4 that more admixtures were dosed for w/c of 0.34 and all admixtures showed some

degree of retardation at the dosages used. One may think that the pastes having more

admixtures would produce less heat as cement hydration would be retarded, which

was contradictory to what was observed. On the other hand, higher cement content in

the pastes with w/c of 0.34 may affect the adsorption of the admixtures and thus the

amount of admixtures retained in pore solutions may also affect the heat evolution.

Further research is required to understand such behaviours better.

112
Chapter 4 Results and Discussion

4.2 Degree of Cement Hydration

4.2.1 Reduction of C3S in Cement Pastes

Since C3S takes up more than 50% of the cement and hydrates from early age, its

reduction in quantity will provide information on the degree of cement hydration.

The C3S content will be reduced with cement hydration. In Figs. 4-7 and 4-8, it can be

seen that the C3S content in all the cement pastes was reduced with time, from the

earliest at 2 hours to 91 days. The chemical admixtures had significant effect on the

cement hydration at early age up to about 3 days. At 3 days and thereafter, the

differences in the C3S content in the paste samples with and without admixtures were

not significant at both w/c ratios. The results are consistent with the small differences

of the cumulative heat observed at 3 days (Figs. 4-4 and 4-5) for the various paste

samples. From Figs. 4-7 and 4-8, it is also observed that the influences of the

admixtures on the C3S hydration were dependent on the w/c ratio.

At a w/c of 0.34, the C3S in the control cement paste and the paste with the SNF

superplasticizer started to drop substantially as early as 2 to 4 hours. The C3S content

in the paste with PCE admixture began to decrease significantly at 4 to 8 hours. For

the pastes with PLS and UNA admixtures, the C3S began to reduce noticeably at 12

hours. The rate of reduction remained more or less the same till 3 days.

With a w/c of 0.40, the reduction in the C3S content was not noticeable in the control

paste and the paste with the SNF superplasticizer until 8 hours. This was more than 4

hours behind that with a w/c of 0.34. The C3S content in the cement pastes with PCE,

113
Chapter 4 Results and Discussion

PLS and UNA superplasticizers dropped noticeably at around 12 hours, whereas that

with the BCS and BCA admixtures had a noticeable drop 12 hours later.

The data of the C3S reduction and the cumulative heat both suggest that the cement

hydration of the pastes with the various admixtures was not delayed at 3 days and

beyond.

4.2.2 Hydration Progress in the Cement Pastes

The main hydration products from cement hydration are calcium silicates, calcium

hydroxide, and calcium sulphoaluminates (ettringite and monosulphoaluminate).

Since C-S-H is amorphous, XRD will not provide information for its formation. No

AFt and AFm phases were detected in the XRD analyses. This might be related to the

sample preparation in which the grinding of the dried cement pastes into powder

might have destroyed the calcium sulphoaluminate crystals. In this section, the

increases in the CH and the NEW content in the cement pastes are used as indicators

of the progress of cement hydration.

4.2.2.1 Calcium hydroxide (CH) in cement pastes

Both the C3S and C2S hydration produce calcium hydroxide. The C3S starts to hydrate

immediately after it comes in contact with water, whereas the C2S does not have

significant hydration at early age. The CH content in the cement pastes were

determined by TG analysis and the results are presented and discussed in this section.

Figure 4-9 shows TG curves of the control cement paste samples (w/c = 0.40) with the

114
Chapter 4 Results and Discussion

hydration ages from 2 hours to 91 days. The TG curves of the other dried cement

paste samples can be found in the appendix. From the curve, it is observed that the

drop around 500 °C, which corresponds to the decomposition of CH, occurred in the

paste sample after 4 hours hydration. This indicates that the CH appeared in the

cement paste after 4 hours of hydration. The drop became more substantial with the

increase in the curing age. It is seen in Fig. 4-9 that samples lost mass around 750 °C,

which corresponds to the decomposition of calcium carbonate (CaCO3). There are two

possible sources of calcium carbonate; either from the original cement in the form of

limestone powder or from the carbonation of CH. The average amount of CO2

released from decomposition of calcium carbonate of the same cement was 2.09%,

with sample size of 3 and standard deviation of 0.05%. The results presented in Table

4-2 show that the amount of CO2 released from the control paste at ages of 2 hours to

91 days were all not more than 2.16%. This confirms that carbonation of paste

samples was minimal and the calcium carbonate presented in the samples was mainly

from the cement. The current ASTM C150 allows the use of up to 5% limestone

(main composition is calcium carbonate) in Portland cements.

The increase in the CH content (determined from the TG analysis) with time of the

cement pastes with or without the admixtures is shown in Figs. 4-10 and 4-11 for the

pastes with w/c ratios of 0.34 and 0.40, respectively (data are presented in the

Appendix5). For both w/c ratios, it seems that the CH precipitation occurred at 4 to 8

hours in the paste with the SNF superplasticizer; at 8 to 12 hours in the paste with the

PCE admixture, and after more than 12 hours in the pastes with all LS based

admixtures except for paste with PLS admixture with a w/c ratio of 0.34 in which CH

5
The raw data can be found in the appendix wherever there is a figure plotted.

115
Chapter 4 Results and Discussion

precipitated at 8 to 12 hours. However, the overall CH produced in these pastes with

various admixtures at 91 days was similar to that of the corresponding control paste.

This indicates that the admixtures had significant influence on the cement hydration at

early age, but not at late age. At an early age from 2 hours to 1 day, the effect of the

admixtures on the time of occurrence and the amount of the CH in the cement pastes

were in an increasing order from the SNF < PCE < PLS < UNA < BCA < BCS.

The CH content was also monitored by X-ray diffractometer with 2θ = 18o, and the

time when the CH crystals precipitated in the pastes was determined and compared

with that from TG analysis in Table 4-3. However, the CH content was not determined

from the XRD spectrum due to the difficulty in quantitative XRD analysis.

Ramachandran (1979) also reported that XRD method underestimates CH by not

registering the presence of microcrystalline or near amorphous CH.

From Table 4-3, it is observed that the time of the CH appearance detected in the TG

analysis corresponded well to the time when significant C3S reduction was detected

for all the pastes at both w/c ratios. However, the XRD seemed to detect the CH

appearance earlier than the TG analysis.

Compared to the heat curves in Fig. 4-4, it is natural for CH to appear in the order of

SNF < PCE < PLS < UNA. But it is quite surprising that the control paste had the

lowest cumulative heat in the first 2 hours, yet it was the first to start CH precipitation.

This may be due to the flocculation effect when w/c was low and no admixture was

added, thus slower rate of ion dissolution and lower heat. In Fig. 4-5, the zoomed-in

graph clearly indicates the hydration rates were in the order of CTR > SNF > PCE >

116
Chapter 4 Results and Discussion

PLS > UNA > BCS > BCA within 12 hours.

It was reported by Chan et al (1996) and Chandra and Björnström (2002) that ion

concentrations in pore solutions at lower w/c were higher than at higher w/c so that

crystals could precipitate faster to stiffen the pastes of lower w/c. This indicates that

the cement hydration at the lower w/c ratio of 0.34 would be delayed less than that at

the higher w/c ratio of 0.40, which has been confirmed by the cumulative heat up to 3

days (Fig. 4-4) and C3S reduction and CH increase (Table 4-3).

4.2.2.2 Non-evaporable water in cement pastes

Figures 4-12 and 4-13 show the increase in the NEW content in the cement pastes

determined by ignition in a furnace with the hydration time. From the figures, it is

apparent that all the pastes with the admixtures had less NEW content than their

corresponding control pastes at earlier age up to 12 hours except for the paste with the

SNF admixture at the w/c of 0.34. This particular paste had NEW content similar to

that of the corresponding control paste from 2 hours to 28 days. It is clear that all the

admixtures used delayed cement hydration to certain extents.

It was also noted that the two cement pastes with the regular BCS and BCA always

had lower NEW contents than those with the superplasticizers at an early age up to 1

day. However, they had higher NEW contents than the control pastes and the pastes

with the superplasticizers at 28 and 91 days. Similar trend was also observed for the

pastes with the LS based superplasticizers at the w/c of 0.34. These two pastes had

lower NEW content at an early age up to about 8 to 12 hours, but had higher NEW

content than the pastes with the SNF and PCE superplasticizers at 91 days.

117
Chapter 4 Results and Discussion

The non-evaporable water in the cement pastes were also determined from the TG

curves. Figure 4-14 shows that the trend of the NEW determined from the ignition in

the Lenton furnace agreed well with that determined from the TG curves (R2 = 0.97).

However, the CH contents in the pastes determined from the former were lower than

those determined from the latter by about 3%. As mentioned in Chapter 3, the samples

used for both tests were vacuum-dried to constant weight and subsequently stored in

desiccators till the time of testing. However, there were some differences in the testing

conditions. The heating in both tests started from room temperature (around 30 °C)

with a constant heating rate of 10°C/min till 950 °C. However, in the TG analysis the

measurement of the mass was stopped immediately after the temperature had reached

950 °C, whereas the samples in the furnace stayed at 950 °C for another 3 hours, and

then were weighed after being cooled down. The other difference was that in the TG

test, the samples were heated in an environment with a flow of nitrogen gas (N2),

whereas the samples in the furnace were heated without N2 gas. It is very unlikely to

be caused by carbonation, which was confirmed in the earlier subsection. Moreover, if

it were due to carbonation, both the furnace and TG would have registered the

carbonation effect during the tests. No explanation has been found and further

research on this is required.

It is interesting to compare the data analyzed for CH and NEW contents of the pastes.

In Fig. 4-10, the pastes with PLS and UNA at w/c of 0.34 had lower CH contents at 7,

28, and 91 days compared to those with PCE and SNF superplasticizers. In Fig. 4-12,

all four superplasticized pastes had similar NEW contents at 7 and 28 days; but the

pastes with PLS and UNA showed higher contents at 91 days. Similar variations were

118
Chapter 4 Results and Discussion

also observed in the case of w/c of 0.40. The small variations at discrete ages for

independent test methods are believed to be normal experimental variability. This is

further supported by the C3S remaining in the pastes at 7, 28 and 91 days, shown in

Figs. 4-7 and 4-8. The independent testing methods suggest the water reducing

admixtures under investigation had insignificant effect on the hydration at later ages

from 7 till 91 days.

4.2.3 Degree of Hydration in Cement Pastes

The degree of hydration, α, is a measure of the extent of the reaction between the

cement and water. It is a function of time and it varies between 0% at the beginning of

hydration and 100% when complete hydration is reached. In reality, full hydration

may never be reached for concrete at jobsites, particularly for those with low w/c

ratios.

Generally, hydrated cement paste is assumed to be comprised of three components:

anhydrated cement, hydration product (gel) and capillary pores. According to Taylor

(1997), the water presented in the cement paste is categorized as evaporable and non-

evaporable water (wn). The degree of hydration (α) at a certain age (t) can be related

to the ratio of the amount of non-evaporable water at time t to that of complete

hydration denoted by the infinite sign ∞, as shown in Eq. 4-1.

(t )
w
α (t ) = n( ∞ ) (4- 1)
wn

The typical value for wn for complete hydration is 0.23 - 0.25 (Taylor, 1997).

119
Chapter 4 Results and Discussion

As mentioned earlier in Chapter 3, the non-evaporable water contents (Figs. 4-12 and

4-13) were determined as the relative mass loss between 105 and 1000 °C, corrected

for the loss on ignition (LOI) of the dry cement itself. Assuming that the ultimate

amount of non-evaporable water wn (∞) is 0.24, the degrees of hydration at various

times are shown in Tables 4-4 and 4-5.

Similar to Eq. 4-1, the degree of hydration also can be estimated from the chemical

reaction equations of calcium silicates, namely C3S and C2S. Shown in Eqs. 2-1 and

2-2, both silicates6 produce calcium hydroxide. Since we know the original amount of

C3S and C2S in cement from the Bogue calculations, the total calcium hydroxide may

be estimated at the time of complete hydration. The theoretical total CH content will

be about 24% for the cement paste samples. Therefore, the degree of hydration at

various time, t, can be estimated using Eq. 4-2.

CH (t )
α (t ) = (4- 2)
CH ( ∞ )

Based on this estimation, the degrees of hydration at various times are shown in

Tables 4-4 and 4-5.

From the estimated degrees of hydration of pastes with and without admixtures for

both w/c ratios, the behaviours are the same as the trends observed earlier in the NEW

and CH contents over the hydration time from 1 day to 91 days. However, the

differences between the two methods of degree of hydration are noticeably large for

both w/c ratios.

6
The relative molecular masses of C3S, C2S and CH are 228.32, 172.24 and 74.09
respectively.

120
Chapter 4 Results and Discussion

The ultimate degree of hydration, αu, is strongly affected by w/c ratio (Hansen, 1986).

According to Mindess et al (2003), the minimum w/c ratio should not fall below 0.42

for complete hydration if cement paste is cured under sealed condition, as shown in

Eq. 4-3.

(w / c )min = 0.42α (4-3)

It is noted that this is not affected by the curing temperature. Therefore the ultimate

degree of hydration is about 81% for w/c of 0.34 and around 95% for w/c of 0.40.

From Tables 4-4 and 4-5, the 91-day degrees of hydration for NEW method are quite

close to the ultimate degrees of hydration estimated from Eq. 4-3. It seems that the

NEW and CH contents were obtained through the laboratory tests to certain accuracy

and both can be used for estimation in terms of the degrees of hydration of pastes with

and without admixtures. Compared to the estimation of the CH method, NEW

estimation seems to be better. As discussed earlier in Section 4.2.2.2, it is most likely

related to the test methods.

4.3 Workability Retention of Mortars with Time

4.3.1 Change in the Yield Stress of Mortars with Time

With initial yield stress controlled at 75 ± 15 Pa for the mortars with the admixtures; it

makes comparison of the yield stress change with time meaningful. The mortar with

the BCS admixture failed to achieve the target yield stress even though a dosage of

60% more than the maximum dosage recommended by manufacturer was used. The

121
Chapter 4 Results and Discussion

initial yield stress of control mortar with the w/c ratio of 0.40 was 490 Pa. No control

mortar mixture was included as it was impossible to achieve the same initial target

workability without the admixtures. Figure 4-15 shows the average and standard

deviation of the initial yield stress of all the mortar mixtures with the admixtures.

Although the BCS admixture had the initial yield stress of 247 Pa which was not

comparable to that of the mortars with the other admixtures, it was still included in the

study to see its effect.

Regardless of the w/c ratios, the yield stresses increased (R2 ≥ 0.95) with time up to

60 minutes and the mortars stiffened with time (Figs. 4-16 and 4-17). Petit et al (2006)

reported linear relationship between the yield stresses and elapsed time up to the end

of the induction period (> 2 hours) of mortars with a SNF superplasticizer.

For both w/c ratios, the slopes of the yield stress are in the same order of SNF > PCE

> PLS > UNA for the four superplasticzers. This indicates that SNF has the least

workability retention and will experience greater loss of workability with time among

the four superplasticizers. The workability retention capability of the two LS based

superplasticizers (PLS and UNA) was better than that of the PCE and SNF

superplasticizers. The performance of the UNA and PLS superplasticizers was similar,

but the UNA admixture was slightly better than the PLS admixture on workability

retention.

The two regular LS based WRAs showed similar rate of yield stress increase and

workability loss. However, their rates of the yield stress increase were much higher

than those of the LS based superplasticizers, and even higher than the SNF

122
Chapter 4 Results and Discussion

superplasticizer (Fig. 4-17). This suggests that the cement in the mortars with these

two admixtures might have more hydration than that in the mortars with the other

admixtures during the first hour. For the paste with the BCS admixture, dispersing

capability of the admixture may also be an issue since the target yield stress was not

achieved. It is indeed shown in Fig. 4-4 that the cement pastes with the BCS and BCA

admixtures released heat much faster than the pastes with the other admixtures in the

first hour, and in Fig. 4-5 that the former ones had higher cumulative heat up to 4

hours than the latter ones.

It is well known that admixtures may be adsorbed not only onto the surface of the

cement particles but also onto the hydration products. Chen (2007) suggests that the

admixture molecules may even be covered by the hydration products. For the pastes

with the BCS and BCA admixtures, heat was released much faster than other pastes in

the first hour. It could probably be due to the accelerated C3A hydration. Because of

this rapid reaction, the admixtures may be removed from solution due to the

adsorption onto ettringite. This may explain the significant workability loss of the

mortars with the BCS and BCA admixtures based on the fact that the initial yield

stress was the same for mortars with admixtures. However, the results obtained from

the current studies only suggest but are unable to lead to definite conclusions. Further

research is needed.

Comparing Figs. 4-16 and 4-17, it is observed that the increase in the yield stresses of

the higher w/c ratio of 0.40 was less significant than those of the lower w/c ratio of

0.34, although the initial yield stresses were on the same level. The results are

consistent with the cumulative heat discussed in Section 4.1 and cement hydration

123
Chapter 4 Results and Discussion

discussed in Section 4.2.

The yield stresses were evaluated during the first hour which was in the induction

period of cement hydration from the heat curves obtained (Figs. 4-2 and 4-3). A

change in slope in the rate of heat curve can be easily detected for all the cement paste

mixes. This elapsed time corresponds to the end of the induction period (tf) and

reflects the beginning of the acceleration period of cement hydration (Lei and Struble,

1997). It is important to note that the end of the induction period is usually considered

as the point where the tangent lines of heat curves during the induction period and

acceleration period of cement hydration intersect (Lei and Struble, 1997). However,

instead of this conventional approach, the end of the induction period tf was

determined by the point corresponding to the lowest rate of heat before the

acceleration occurred in the rate of heat curve.

The normalized time t′ is defined as the ratio of any given elapsed time (e.g. BML and

flow table tests at 10 min, 30 min and 60 min) within the induction period to the final

time tf (t′ = t / tf). It is therefore a non-dimensional parameter and enables the

comparisons of variations in rheological parameters of the various mixtures at the

same relative scale of time, regardless of the length of the induction period.

As shown in Fig. 4-18, yield stresses increase linearly (R2 = 0.95) with the normalized

time, for superplasticized mortar mixes, being at w/c of 0.34 and 0.40. However, a

poor relationship exists for the two regular water reducing admixtures, due to their

extremely strong retardation, hence longer induction period.

124
Chapter 4 Results and Discussion

4.3.2 Change in Plastic Viscosity of Mortars with Time

Figures 4-19 and 4-20 show plastic viscosity changes with time for the mortars with

and without admixtures at the w/c of 0.34 and 0.40, respectively. At the lower w/c of

0.34, the initial plastic viscosities of the mortars (measured 10 minutes after the

cement came in contact with water) were 5.2 – 9.7 Pa.s, higher than those (3.3 – 4.4

Pa.s) at higher w/c of 0.40. At 60 minutes, the plastic viscosity of the mortars with the

w/c of 0.34 ranged from 6.3 to 9.8 Pa.s, and that with the w/c of 0.40 ranged from 2.6

to 4.3 Pa.s. It seems that the change in the plastic viscosity of the mortars with time

was not as significant as that in the case of yield stress.

It was noticed that for the mortars with the two modified LS based superplasticizers

(PLS and UNA), the initial plastic viscosities were higher than those with the PCE

and SNF superplasticizers, and they remain largely unchanged with time. In other

words, the modified LS superplasticizers had more effects on the yield stress τ 0 than

the plastic viscosity µ . This observation agreed with the results reported by Wallevik

(2003).

However, the other two superplasticizers behave differently for the two w/c ratios.

The plastic viscosity of the mortars with the PCE superplasticizer increased with time

at the w/c ratio of 0.34, but kept almost constant at the w/c ratio of 0.40. The plastic

viscosity of the mortars with the SNF admixture showed the opposite. It decreased at

the w/c of 0.34 but increased at the w/c of 0.40 with time.

For the two regular WRAs (BCS and BCA), the plastic viscosity of the mortar with

the BCS admixture increased substantially from 10 to 30 minutes, whereas that with

125
Chapter 4 Results and Discussion

the BCA decreased from 10 to 60 minutes. Due to the high initial yield stress and

plastic viscosity of the mortar with the BCS admixture and significant workability

loss, it was not possible to measure the yield stress and plastic viscosity after 60

minutes. .

4.3.3 Change in Flow Value of Mortars with Time

The flow value was reported in terms of the increase in average base diameter of the

mortar mass, expressed as a percentage of the original base diameter (100 mm) of the

flow cone after 10 drops of the flow table. For example, a mortar mass with a spread

of 150mm in diameter after 10 drops had a net flow of 50mm, and the flow value is

therefore 50%. Since the flow table has a diameter of 250mm, the term “overflow”

used in the Appendix (Table A-9) means a flow value of 150% or more.

Figure 4-21 show the change in the flow values of the various mortars at w/c ratios of

0.40 with time. The data for a w/c ratio of 0.40 were collected in the manner described

in Chapter 3 and presented in Fig. 4-21. The data for a w/c ratio of 0.34 were not

collected in the specified manner at 10 drops but at 25 drops because they even

overflowed at 30 minutes. The data of the latter were presented in Table 4-6.

Shown in Fig. 4-21, the flow values for the mortars were different at 10 minutes even

though their initial yield stress was controlled at 75 ± 15Pa except for the mortar with

the BCS admixture. For the two LS based superplasticizers, their mortars showed

similar flow values of about 140 % at both w/c ratios. Whereas for the PCE, SNF and

BCA admixtures, their mortars had roughly the same initial flow about 120 %. The

126
Chapter 4 Results and Discussion

mortar with the BCS admixture had the lowest flow of about 98% as expected.

It is seen in Table 4-6 that mortars with superplasticizers overflowed the flow table at

10 and 30 minutes when 25 drops were applied. It is shown that mortars with PCE

and SNF superplasticizers had similar flow values at 60 minutes and that mortars with

two modified LS superplasticizers also showed close flow values. These flow values

at 60 minutes supported the corresponding results of yield stress.

From Fig. 4-21, it seems that for a w/c of 0.40, the mortars had similar rate of

reduction in the flow values except for that with the PCE admixture. The mortar with

the PCE admixture had less than 5% reduction in the flow from 10 to 60 minutes,

which indicates that the PCE based superplasticizer was able to retain workability

well. This seems to contradict the results of the yield stress and plastic viscosity. This

suggests that the one-point flow value test may not sufficiently describe the

workability of mortars under the circumstance that their flow values fall within a

rather narrow range.

4.3.4 Relationship between the Yield Stress and Flow Value

As shown in Fig. 4-22, the yield stress is inversely proportional to the flow value with

an R2 value of 0.95, if the high flow values of 140 % (in the circle) are excluded.

These high flow values were not erratic. It merely suggests that the mortar was too

flowable to be properly measured by flow table, and the result was not proportional to

the yield stress. In fact, it is understood that when yield stress is sufficiently low, flow

should be infinite. The power fit in Fig. 4-22 thus is also reasonable, which has an R2

value of 0.91.

127
Chapter 4 Results and Discussion

This relationship is independent of w/c ratios, admixtures and time. Similar

relationships between yield stress and slump have been reported for concrete by

Ferraris and Brower (2001) and Emoto and Bier (2007).

Since the plastic viscosity of most mortars did not change significantly with time, no

definite relationship between the plastic viscosity and flow value was observed.

4.4 Setting Times of Mortars

Setting times of the mortars with and without the admixtures at w/c ratios of 0.34 and

0.40 are plotted in Figs. 4-23 and 4-24, respectively. The order of the setting times of

the superplasticized mortars agreed with the length of induction periods of the

corresponding pastes at w/c ratio of 0.34 (Fig. 4-2), i.e. SNF < PCE < PLS < UNA.

For w/c ratio of 0.40, it was noticed that paste with UNA superplasticizer had similar

length of induction period as that of BCA admixture (Fig. 4-3); however, the

corresponding mortar with UNA superplasticizer had 100 minutes shorter setting

times than those with BCA admixture. The magnitude of heat peak is another factor,

which indicates the admixture effect on set retardation. In the case of UNA and BCA

admixtures, paste with BCA admixture showed lower heat peak (Fig. 4-3). Hence, the

heat evolved from the pastes with the admixtures supported the order of the setting

times of the plasticized mortars, i.e. SNF < PCE < PLS < UNA < BCA < BCS.

Figure 4-25 shows the relationship between the initial and final setting time of the

mortars and the time when acceleration started after the induction period in the heat

curve. Generally, a later start of the acceleration period means that both the initial and

final setting times are longer, as indicated by the linear relationships with R2 ≥ 0.88

128
Chapter 4 Results and Discussion

for both w/c ratios. The result of the setting times is consistent with the time that CH

appeared in the pastes with different admixtures incorporated. The data shown in the

figure confirm that the setting times are affected by w/c ratio, besides the cement

hydration and dosages of admixtures. In other words, the heat curve itself is not

sufficient to determine the setting times of mortar or concrete.

Since the setting time was determined by the penetration resistance, the setting times

of the mortars with the lower w/c of 0.34 were generally shorter than those with the

higher w/c of 0.40 if no chemical admixture is used. However, the setting times are

also affected by the type and dosage of the admixtures, in addition to the w/c ratio.

For example, the setting times of the mortars with the PLS and UNA at w/c of 0.34

were higher than those at w/c of 0.40 because the formers had higher dosages of the

admixtures than those of the latter and the admixtures had retarding effects.

Comparing the setting times of superplasticized mortars with different w/c ratios, it is

interesting to note that, the difference between the initial and final setting times for the

mortars with the superplasticizers ranged from 65 to 100 minutes, despite the large

differences in their initial setting times. The differences in the initial and final setting

times for the mortars with BCA and BCS admixture were 160 and 300 minutes,

respectively, which were longer than those with the superplasticizers but the initial

setting times of the mortars with the two regular WRAs were also longer. This

indicates that admixtures may have a strong influence on the initial setting time but

less influence on the final setting time.

Sugar content in the LS admixtures is partly responsible for the retardation. However,

129
Chapter 4 Results and Discussion

the sugar content may not be the only factor (Vikan, 2005). As seen in Fig. 4-24,

mortar with the BCS admixture had an initial setting time almost 4 hours longer and a

final setting time 6 hours longer than those with the BCA admixture. Yet the BCS

admixture from hardwood had a reducing matter of 3%, whereas the BCA admixture

from softwood had 7% reducing matter.

The PCE and SNF based superplasticizers were sugar free, but they also showed

retardation effect to certain extent.

It is important to note that the effect of the admixtures might not be the same for the

set retardation and for the workability retention. In other words, longer set retardation

may not always correspond to better workability retention. For the four

superplasticizers used in the mortars at both w/c ratios, it was found that the

workability retention determined by the yield stress was better when the setting times

were longer. However, for the two regular water reducing admixtures, it seems to be

otherwise.

For the two LS based superplasticizers, the mortar with the PLS had shorter setting

times than that with the UNA at both w/c ratios. This may be related partly to the

lower dosage of the admixture used in the former than the latter.

4.5 Pore Structure of Cement Pastes

Figure 4-26 shows typical curves7 of the intruded cumulative pore volume versus the

7
The curves for the cement pastes with other admixtures at w/c ratios of 0.34 and 0.40 are
included in the Appendix (Tables A-13 to A-24).

130
Chapter 4 Results and Discussion

mean pore diameter of the cement paste with the PCE admixture at w/c of 0.40 at

various ages based on the mercury intrusion porosimetry. The total porosity of the

cement pastes with the w/c ratios of 0.34 and 0.40 cured for different periods are

shown in Figs. 4-27 and 4-28, respectively. In general, the cumulative pore volume in

the samples decreased with time due to cement hydration. The cumulative pore

volume of the cement pastes with the w/c ratio of 0.34 was smaller than that with the

w/c ratio of 0.40.

For the repeated MIP tests that were randomly selected for each series of pastes, it

was found that the variations (in terms of relative difference) in the total porosities

were within 2% for the pastes with w/c of 0.34, and 5% for those with w/c of 0.40

(Table 4-7). For example, for the paste with the PLS superplasticizer and w/c of 0.34

at 91 days, the total porosity of the repeat sample was 22.4%. Compared to the first

result of 22.7%, it has an absolute difference of 0.3%, and a variation of 1.3%. The

relatively small variations of the repeat samples indicate that the test results are

repeatable. Therefore, only one test was used for the remaining samples. For those

samples with duplicate tests, the first set of results was used in the analysis.

4.5.1 Total Porosity of Cement Pastes with Admixtures

For both w/c ratios, the differences in the porosity of the pastes with different

admixtures at early age are mainly associated with the retardation effect of admixtures.

The porosities of pastes with all admixtures were similar at 7 days with the given w/c

ratio.

For the w/c ratio of 0.34 (Fig. 4-27), the cement pastes with the PCE, SNF and PLS

131
Chapter 4 Results and Discussion

superplasticizers had similar porosities up to 3 days. The paste with modified LS

superplasticizer UNA had noticeably higher porosities at 1 day and 3 days, but its

porosity at 7 days were comparable to that of the other pastes. At 28 and 91 days, the

paste with the PCE superplasticizer had lower total porosity than those with the SNF

and LS superplasticizers, which may be due to the dispersion ability of the

superplasticizers. The pastes with the SNF and LS superplasticizers had similar total

porosity.

For the w/c of 0.40 (Fig. 4-28), the two LS based regular admixtures (BCS and BCA)

had slightly lower porosity at 1 day. This may be explained by the high rate of cement

hydration in the first 5 hours (related to C3A reaction) when these two admixtures

were incorporated in the cement pastes (Fig. 4-5). Having said that, it is noted that

mortar with BCS admixture had final setting time of nearly 23 hours and it had little

strength at 1 day. This suggests that its respective paste sample may not be strong

enough to produce reliable result. At 3 days, the pastes with PCE and SNF

superplasticzers appear to have lower porosity than those with the LS admixtures. The

total porosity of the pastes with different admixtures at 28 days was not significantly

different. At 91 days, however, the total porosities of the pastes with the BCS, BCA,

and UNA admixtures were lower than those of the paste with PCE, SNF, and PLS

superplasticizers. This might be related to the significant retardation effect of the

former on C3S hydration at early ages which might have resulted in more

homogenous distribution of hydration products in the cement pastes.

4.5.2 Pore Size Distribution of Cement Pastes with Admixtures

As discussed earlier, pores in cement paste can be either gel or capillary pores.

132
Chapter 4 Results and Discussion

Capillary pores can be further divided into small, medium and large capillaries

according to their sizes. The MIP under study was able to detect capillary pores with

diameters as small as 3.8 nm. According to Table 2-3 (Mindess et al, 2003), the pores

were divided into small capillary pores with diameters from 3.8 to 10 nm (denoted as

G), medium capillary pores with diameters from 10 to 50 nm (denoted as M), and

large capillary pores with diameters from 50 nm to 10 µm (denoted as L). The small

capillary pores are part of the gel pores (Mindess et al, 2003).

Pore size distributions of the cement pastes with admixtures at various ages from 1 to

91 days are shown in Figs 4-29 to 4-33. The pore size distribution changed with time

due to cement hydration, and it also differed due to the differences in w/c ratios and

the admixtures used in the cement pastes. For the cement pastes with w/c ratios of

0.34 and 0.40 with various admixtures, their proportions of small capillary pores were

not significantly different, and the differences were mainly in the large and medium

capillary pores. At early ages, particularly at 1 day, the proportions of the large and

medium pores were mainly affected by the effect of admixtures on the retardation of

cement hydration. It is the pore size distribution at ages of 28 and 91 days that is of

importance since it affects permeability and resistance to the penetration of harmful

substances.

For the pastes with w/c of 0.34, it appears that the paste with SNF superplasticizer had

the lowest sum of small and medium capillary pores while pastes with other

superplasticizers had similar amount at 1 day (Fig. 4-29(a)). It also shows that paste

with UNA superplasticizer had the largest amount of large capillary pores. At 3 days

shown in Fig. 4-30(a), the paste with UNA superplasticizer still had the largest

133
Chapter 4 Results and Discussion

amount of large capillary pores while it had slightly lower combination of small and

medium capillary pores than the other pastes which had similar pore size distributions.

At 7, 28, and 91 days (Figs. 4-31(a) to 4-33(a)) the two LS superplasticizers (PLS and

UNA) had less large pores but more medium pores compared with the paste with SNF

admixture. The pastes with the two LS and the PCE superplasticizers had similar large

pores at 28 and 91 days. However, the latter had less medium pores than the former.

Based on Figs. 4-31(a) to 4-33(a), it seems that at lower w/c ratio of 0.34, the two

modified LS, PLS and UNA, and PCE superplasticizers led to a significant reduction

in large and medium capillary pore volumes from 7 to 91 days and that SNF

superplasticizer was able to reduce medium capillary pore volume but not large ones.

The pastes with PLS and UNA superplasticizers at w/c ratio of 0.34 had lower CH

content (Fig. 4-10) compared to those with PCE and SNF ones at 7 and 28 days. This

confirms the general understanding that pore structure of paste with admixtures is

related to C-S-H produced from cement hydration, rather than the production of CH.

At w/c of 0.40, the pastes cured for 28 days (Fig. 4-32(b)) had similar medium and

large capillary pores except that the paste with PCE admixture had slightly less and

the paste with UNA admixture had slightly more large capillary pores. At 91 days

(Fig. 4-33(b)), the pastes with the LS admixtures had similar medium pores compared

with the pastes with PCE and SNF admixtures. However, the pastes with the LS

superplasticizers had less large capillary pores than that with the SNF admixture, but

similar large pores compared with the PCE paste. The pastes with the regular water

reducers (BCS and BCA) appear to have lower large capillary pores at 91 days

compared with those with the superplasticizers. This is probably also related to the

134
Chapter 4 Results and Discussion

strong retardation of the regular water reducers that allowed a more even distribution

of hydration products and resulted in a more homogeneous microstructure. Overall,

the performances of the all the admixtures appear similar, and this suggests that the

effect of the admixtures on pore structures is not pronounced at high w/c ratios

compared with one another.

4.5.3 Threshold and Critical Pore Diameters

The threshold diameter is the first inflection point as shown in Fig. 4-26, and it

indicates the onset of percolation (Mindess et al, 2003). It corresponds approximately

to the minimum diameter of channels that are continuous through the paste at a given

age (Winslow and Diamond, 1970).

The critical pore diameter is the pore size corresponding to the highest rate of mercury

intrusion. This is the point where the slope of the curve of cumulative mercury

intrusion volume against pore diameter is the steepest. The critical pore diameter

represents the mean size of pore entryways that allows maximum percolation

throughout the pore system, and is also called continuous pore diameter (Mindess et al,

2003).

Both the threshold and critical diameters derived from pore size distribution curves

from the MIP tests are closely related to permeability and penetration resistance of the

cement pastes (Ramachandran and Beaudoin, 1999). The coefficient of permeability

increases with an increase in the threshold and critical diameters (Halamickova et al.,

1995; Ramachandran and Beaudoin, 1999). Similarly, the coefficient of chloride ion

135
Chapter 4 Results and Discussion

diffusion varied linearly with the critical pore diameter as determined by the MIP

(Halamickova et al., 1995).

The threshold and critical pore diameters for the cement pastes with w/c ratios of 0.34

and 0.40 are presented in Tables 4-8 and 4-9, respectively. The results show that up to

28 days the cement pastes with a w/c of 0.34 generally had smaller threshold and

critical diameters than the corresponding pastes with a w/c of 0.40. At 91 days, no

definite trend was observed.

For the pastes with a w/c of 0.34, the threshold and critical diameters did not change

significantly with time. For the pastes with a w/c of 0.40, however, the threshold and

critical diameter reduced significantly with time at early age up to 7 days, with no

significant change after that.

For the cement pastes with a w/c of 0.34, the threshold and critical diameters of the

two modified LS superplasticizers were slightly smaller than those of the PCE and

SNF superplasticizers at 28 days. However, both the threshold and critical diameter of

the superplasticized pastes were of the same order.

For the cement pastes with a w/c of 0.40 and the four superplasticizers, their threshold

and critical diameters were similar at 28 days. The pastes with the BCS water reducer

had larger critical and threshold diameters, but that with the BCA admixture had

smaller critical and threshold diameters, compared to those of superplasticized pastes.

136
Chapter 4 Results and Discussion

4.6 Compressive Strength of Mortars

The compressive strength development of the mortars with w/c ratios of 0.34 and 0.40

is shown in Figs. 4-34 and 4-35, respectively. The variations (in terms of standard

deviation calculated from three 50mm mortar cubes and indicated by the “I” bars) are

all below 10% of their respective average strength.

The compressive strengths of mortars with admixtures at 1 day were in a descending

order: SNF > PCE > PLS > UNA > BCA > BCS. This was exactly the opposite of the

setting times as the longer setting time means stronger retardation and less hydration,

hence lower strength at early ages.

For the w/c ratio of 0.34 (Fig. 4-34), the mortars with the PCE, SNF, and PLS

admixtures had similar compressive strength at 7 and 28 days. However, the lower

strength of the mortar with the UNA persisted. At 91 days, all the pastes had similar

strength.

For the w/c ratio of 0.40 (Fig. 4-35) the compressive strength at early ages,

particularly at 1 day, was also affected significantly by the retardation of the

admixtures. For the mortar with the BCS admixture, almost no strength was

developed at 1 day. At 7, 28, and 91 days, the compressive strength of the mortars

with various admixtures were not significantly different.

Compressive strengths of mortars are related to the mix proportions, w/c, degree of

cement hydration, particularly the hydration of calcium silicates C3S and C2S,

distribution of hydration products, porosity, pore structure, and bonding between the

137
Chapter 4 Results and Discussion

sand and cement paste. The admixture will affect rate of cement reaction at early ages,

and consequently may affect the distribution of hydration products and later strength.

Since compressive strength is such a complex mechanical property of mortars, effect

of admixtures on individual factor may not be exactly in line with the strength. For

example, the paste with PCE superplasticizer with w/c of 0.34 had smaller total

porosity and similar small and medium capillary pores compared to the paste with

SNF superplasticizer at 28 and 91 days, however, this does not seem to be reflected in

the corresponding strengths as the corresponding mortar with SNF superplasticizer

showed higher strength.

138
Chapter 4 Results and Discussion

Table 4-1 Times of peak appearance in heat curves of pastes


Mix Time to appear, min
w/c=0.34 Peak “B” Peak “C” Difference, min
CTR 375 430 55
PCE 650 700 50
SNF 440 515 75
w/c=0.40 Peak “B” Peak “C” Difference, min
CTR 450 555 105
PCE 750 805 55
SNF 540 650 110
Note: Peaks “B” and “C” overlap for pastes with other admixtures

Table 4-2 Amount of CO2 from decomposition of CaCO3 in Fig. 4-9 TG curves
Age, hours 0h* 2h 4h 8h 12h
CO2, % 2.09** 1.96 1.95 1.92 2.14
Age, days 1d 3d 7d 28d 91d
CO2, % 2.16 2.08 2.04 1.98 1.95
* This refers to the original cement which has not yet contacted with mixing
water.
** This was the average value of TG results (n = 3, σ = 0.05) for the same
cement.

Table 4-3 Times for C3S reduction & CH appearance in the cement pastes with and
without admixtures detected by XRD & TG
Significant C3S reduction (XRD) CTR PCE SNF PLS UNA BCS BCA
w/c = 0.34 2h 8h 4h 12h 12h - -
w/c = 0.40 8h 12h 8h 12h 12h >12h >12h
Ca(OH)2 appearance (TG) CTR PCE SNF PLS UNA BCS BCA
w/c = 0.34 4h 12h 8h >12h >12h - -
w/c = 0.40 4h 12h 8h >12h >12h >12h >12h
Ca(OH)2 appearance (XRD) CTR PCE SNF PLS UNA BCS BCA
w/c = 0.34 2h 4h 2h 8h 12h - -
w/c = 0.40 4h 8h 4h >12h >12h >12h >12h

139
Chapter 4 Results and Discussion

Table 4-4 Degree of hydration of pastes at various ages (w/c = 0.34)

w/c = 0.34 1d 3d 7d 28d 91d


CTR 58.4 69.6 74.1 78.4 81.0
Estimation PCE 55.0 70.8 71.8 76.0 77.2
from NEW SNF 60.8 68.4 70.5 78.1 84.0
Content PLS 54.5 70.2 74.6 77.4 82.1
UNA 49.5 70.7 74.3 78.2 82.0
CTR 47.7 50.8 60.4 63.4 66.5
Estimation PCE 41.4 51.3 61.2 62.4 67.9
from CH SNF 45.7 56.6 58.7 61.4 69.6
Content PLS 44.6 52.8 55.9 59.0 64.6
UNA 43.5 53.8 55.0 58.4 61.0

Table 4-5 Degree of hydration of pastes at various ages (w/c = 0.40)

w/c = 0.40 1d 3d 7d 28d 91d


CTR 55.5 76.0 76.8 84.3 93.3
PCE 58.0 70.3 76.6 80.7 90.7
Estimation SNF 54.3 74.3 75.9 82.3 87.4
from NEW PLS 43.9 71.8 78.2 83.1 88.3
Content UNA 53.2 65.5 76.2 89.4 90.7
BCS 37.6 71.3 77.1 88.9 93.7
BCA 43.3 68.3 77.3 85.3 92.2
CTR 44.6 51.0 56.2 69.5 72.5
PCE 43.8 54.0 64.8 67.4 69.0
Estimation SNF 45.7 56.9 60.2 61.1 68.7
from CH PLS 36.8 58.8 64.8 66.0 70.2
Content UNA 28.3 50.5 63.6 67.4 71.7
BCS 21.8 57.3 60.7 71.7 72.2
BCA 24.0 52.0 56.3 64.5 71.5

140
Chapter 4 Results and Discussion

Table 4-6 Flow values of mortars with time (w/c = 0.34)


Mortars with Flow Values, %
Age, minutes
Admixtures 10 drops 25 drops
10 n.m.* >150**
PCE 30 n.m. 120
60 n.m. 106
10 n.m. >150
SNF 30 n.m. 118
60 n.m. 100
10 n.m. >150
PLS 30 140 >150
60 102 132
10 n.m. >150
UNA 30 140 >150
60 112 144
* n.m. = not measured
** > 150% is used when an overflow of mortar occurs

Table 4-7 Repeatability of MIP on mortars with and without admixtures


Cement w/c = 0.34 w/c = 0.40
Age
pastes with Porosity, % Variation** Porosity, % Variation
PCE 28d 29.1(29.1) 0.0%
SNF 7d 33.3(34.3) 3.0%
PLS 91d 22.7(22.4) 1.3%
UNA 1d 40.3(40.3) 0.0%
28d 31.6(32.6) 3.2%
BCS 7d 32.9(32.4) 1.5%
91d 24.1(23.1) 4.3%
BCA 1d 42.7(41.4) 3.1%
* The repeated results are shown in parentheses.
** Variation is in percentage difference, not in absolute difference.

141
Chapter 4 Results and Discussion

Table 4-8 Critical and threshold pore diameters for pastes with w/c = 0.34
Pore Diameter Paste with PCE admixture Paste with SNF admixture
(nm) 1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
Critical, dc 67.2 54.4 67.7 54.2 67.1 67.7 54.2 54.4 67.7 54.2
Threshold, dt 102.6 82.3 125.7 125.8 126.5 101.4 101.9 82.3 101.3 81.9
Pore Diameter Paste with PLS admixture Paste with UNA admixture
(nm) 1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
Critical, dc 54.4 54.1 54.6 43.8 67.5 67.4 54.4 43.6 43.8 43.7
Threshold, dt 126.3 82.6 67.5 82.3 124.7 127.2 101.8 82.3 82.1 82.8

Table 4-9 Critical and threshold pore diameters for pastes with w/c = 0.40
Pore Paste with PCE admixture Paste with SNF admixture
Diameter
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
(nm)
Critical,
197.0 82.8 67.9 68.0 54.3 299.9 128.3 68.1 67.0 54.3
dc
Threshold,
288.6 125.7 101.5 103.5 82.3 397.3 160.0 101.2 100.9 82.6
dt
Pore Paste with PLS admixture Paste with UNA admixture
Diameter
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
(nm)
Critical,
742.5 126.4 68.1 67.6 67.1 202.4 102.2 67.4 83.0 67.7
dc
Threshold,
1415.9 191.0 83.1 100.9 100.7 296.2 127.1 101.1 101.3 82.6
dt
Pore Paste with BCS admixture Paste with BCA admixture
Diameter
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
(nm)
Critical,
1421.2 158.3 67.7 83.0 53.8 473.0 82.9 54.2 43.5 53.8
dc
Threshold,
1744.0 196.2 100.4 125.9 82.9 907.0 157.2 82.3 82.3 82.9
dt

142
Chapter 4 Results and Discussion

3.5

3
Rate of heat evolution, mW/g

W/C = 0.40

2.06% 2.20%
2.5
2.35% 2.50%

2 3.00% 3.50%

1.5

0.5

0
0 500 1000 1500 2000 2500
Time, minutes

Fig. 4-1 Effect of SO3 content on heat of cement hydration

4.0
PCE SNF
W/C = 0.34
PLS UNA
3.5 4.2 CTR
Zoom in
3.0
Rate of Heat Evolution, mW/g

2.5
3.7

2.0

1.5
3.2
250 450 650 850 1050
1.0

0.5

0.0
0 500 1000 1500 2000 2500 3000 3500 4000
Time, minutes

Fig. 4-2 Rate of heat evolution of cement pastes (w/c = 0.34)

143
Chapter 4 Results and Discussion

4.0

3.5

3.0
Rate of Heat Evolution, mW/g

W/C = 0.40

2.5 PCE SNF


PLS UNA

2.0 BCS BCA


CTR
1.5

1.0

0.5

0.0
0 500 1000 1500 2000 2500 3000 3500 4000
Time, minutes

Fig. 4-3 Rate of heat evolution of cement pastes (w/c = 0.40)

250

W/C = 0.34
PCE SNF
200 PLS UNA
CTR
Cumulative Heat Evolution, J/g

150
50

100
25
Zoom in

50

0
0 120 240 360 480
0
0 500 1000 1500 2000 2500 3000 3500 4000
Time, minutes

Fig. 4-4 Cumulative heat evolution of cement pastes (w/c = 0.34)

144
Chapter 4 Results and Discussion

250
W/C = 0.40
PCE SNF

PLS UNA
200
BCS BCA
Cumulative Heat Evolution, J/g

CTR

150
50

100
25

50 Zoom in

0
0 120 240 360 480 600 720
0
0 500 1000 1500 2000 2500 3000 3500 4000
Time, minutes

Fig. 4-5 Cumulative heat evolution of cement pastes (w/c = 0.40)

4.0 250

3.5

200

Cumulative Heat Evolution, J/g


3.0
Rate of Heat Evolution, mW/g

2.5
150
W/C = 0.40
2.0
W/C = 0.34
100
1.5

1.0
50

0.5

0.0 0
0 500 1000 1500 2000 2500 3000 3500 4000
Time, minutes

Fig. 4-6 Rate and cumulative heat evolution of two control mixes

145
Chapter 4 Results and Discussion

70
63.2% C3S in the cement

60 W/C = 0.34

CTR PCE
C3S, % by mass of paste sample

50
SNF PLS

UNA
40

30

20

10

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-7 C3S reduction in paste with time (w/c = 0.34)

70
63.2% C3S in the cement

60
W/C = 0.40
C3S, % by mass of paste sample

CTR PCE
50
SNF PLS

40 UNA BCS

BCA
30

20

10

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-8 C3S reduction in paste with time (w/c = 0.40)

146
Chapter 4 Results and Discussion

Fig. 4-9 A typical TG curve showing mass loss over time (w/c=0.40 control paste)

18

16
CH (TG) , % by mass of paste sample

14

12

10
W/C = 0.34
8
CTR PCE
6
SNF PLS
4

UNA
2

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-10 CH content in pastes increases with time from TG curves (w/c=0.34)

147
Chapter 4 Results and Discussion

18

16
CH(TG), % by mass of paste sample
14

12

10
W/C = 0.40
8
CTR PCE

6 SNF PLS

4 UNA BCS

BCA
2

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-11 CH content in pastes increases with time from TG curves (w/c=0.40)

18

16

14
Non-evaporable water (furnace),
% by mass of paste sample

12

10

W/C = 0.34
8
CTR PCE
6
SNF PLS

4 UNA

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-12 Non-evaporable water content with time from furnace burning (w/c = 0.34)

148
Chapter 4 Results and Discussion

18

16

14
Non-evaporable water (furnace),
% by mass of paste sample

12

10

8 W/C = 0.40
CTR PCE
6
SNF PLS
4
UNA BCS
2
BCA

0
2h 4h 8h 12h 1d 3d 7d 28d 91d
Age

Fig. 4-13 Non-evaporable water content with time from furnace burning (w/c = 0.40)

25
Non-evaporeable water (TG), % by mass of paste sample

20

15

y = 1.06x + 2.86
2
R = 0.97
10
W/C 0.34
W/C 0.40

0
0 2 4 6 8 10 12 14 16 18 20
Non-evaporeable water (Furnace), % by mass of paste sample

Fig. 4-14 Comparisons of non-evaporable water content

149
Chapter 4 Results and Discussion

125

100
Average Initial Yield Stress, Pa

W/C = 0.40 W/C = 0.34

75

50

25

0
PCE SNF PLS UNA BCS BCA PCE SNF PLS UNA

Fig. 4-15 Average and standard deviation of the initial yield stresses of all mortars

480

420

W/C = 0.34
360
PCE SNF
Yield Stress, Pa

300 PLS UNA

240

180

120

60
0 10 20 30 40 50 60
Time, min

Fig. 4-16 Yield stress response on mortars with time at 30 ± 3 °C (w/c = 0.34)

150
Chapter 4 Results and Discussion

480

W/C = 0.40
420
PCE SNF

360 PLS UNA


Yield Stress, Pa

BCS BCA
300

240

180

120

60
0 10 20 30 40 50 60
Time, min

Fig. 4-17 Yield stress response on mortars with time at 30 ± 3 °C (w/c = 0.40)
500

450
PCE SNF
SP
400
PLS UNA
350
WRA BCS BCA
Yield stress, Pa

300

250
2
R = 0.95 (for SP)
200

150

100

50

0
0.00 0.05 0.10 0.15 0.20 0.25
Normalized time, t/tf

Fig. 4-18 Responses of yield stresses of mortars at normalized time

151
Chapter 4 Results and Discussion

10

8
Plastic Viscosity, Pa.s

W/C = 0.34
PCE SNF
4
PLS UNA

2
0 10 20 30 40 50 60
Time, min

Fig. 4-19 Plastic viscosity response on mortars with time (w/c = 0.34)

10

W/C = 0.40
8
PCE SNF
Plastic Viscosity, Pa.s

PLS UNA

6 BCS BCA

2
0 10 20 30 40 50 60
Time, min

Fig. 4-20 Plastic viscosity response on mortars with time (w/c = 0.40)

152
Chapter 4 Results and Discussion

150

PCE SNF
140
PLS UNA

130
BCS BCA
Flow Value, %

120

110

100

90

80
10 30 60
Time, minutes

Fig. 4-21 Change in flow value of mortars with time (w/c = 0.40)

150

PCE SNF
140
PLS UNA

130 -0.2319
Power: y = 349.19x BCS BCA
2
R = 0.91
Flow value, %

120

110

Linear: y = -0.1079x + 128.26


2
100 R = 0.95

90

80
0 50 100 150 200 250 300 350 400 450 500
Yield stress, Pa

Fig. 4-22 Relationship between yield stress and flow value

153
Chapter 4 Results and Discussion

UNA 875 85

PLS 665 85

PCE 300 75

SNF 250 70 Initial


Final - Initial

0 240 480 720 960 1200 1440


Setting Time, min

Fig. 4-23 Initial and final setting times of prepared mortars (w/c = 0.34)

BCS 1075 300

BCA 850 160

UNA 745 95

PLS 560 100

PCE 330 90

Initial
SNF 260 75 Final - Initial

0 240 480 720 960 1200 1440


Setting Time, min

Fig. 4-24 Initial and final setting times of prepared mortars (w/c = 0.40)

154
Chapter 4 Results and Discussion

1500

0.34, initial 0.40, initial


2
1200 R = 0.91
0.34, final 0.40, final 2
R = 0.88
Setting Times of Prepared Mortar, min

2
R = 0.88
2
R = 0.95
900

600

300

0
0 100 200 300 400 500 600 700 800
Time to Start Acceleration Period in a Heat Curve, min

Fig. 4-25 Relationship between the initial and final setting times of prepared mortars
and time to start the acceleration period in heat curves

0.30

1d
0.25
Cumulative Pore Volume (Intrusion) mL/g

0.20
3d

7d
28d critical
0.15 diameter, dc
91d

0.10

threshold
0.05 diameter, dt

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Pore Diameter, µm

Fig. 4-26 A typical MIP graph (w/c = 0.40, paste with PCE superplasticizer)

155
Chapter 4 Results and Discussion

50
PCE SNF
PLS UNA
40
Total Porosity, %

30

20

10

0
1d 3d 7d 28d 91d
Age

Fig. 4-27 Total porosity of pastes with w/c = 0.34 at various ages
50

45 PCE SNF PLS

40 UNA BCS BCA


35
Total Porosity, %

30

25

20

15

10

0
1d 3d 7d 28d 91d

Fig. 4-28 Total porosity of pastes with w/c = 0.40 at various ages

156
Chapter 4 Results and Discussion

50

(a) W/C = 0.34 G M L

40

30
Porosity, %

L, 27.2
L, 22.3 L, 24.7
L, 25.1
20

10
M, 11.0 M, 9.2 M, 10.8
M, 8.1

G, 1.5 G, 2.4 G, 2.9 G, 2.3


0
PCE SNF PLS UNA

50
(b) W/C = 0.40 G M L

40

30
Porosity, %

L, L, L,
32.9 35.8 L, 33.4
39.2 L,
L,
39.7
20 39.6

10

M, 9.2 M, 7.0 M, 8.0


M, 2.9
M, 0.8 M, 1.7
G, 1.5 G, 2.1 G, 3.0 G, 2.5
0 G, 1.2 G, 1.3
PCE SNF PLS UNA BCS BCA

Fig. 4-29 Total porosities and pore size distributions of the pastes at 1 day

157
Chapter 4 Results and Discussion

50

(a) W/C = 0.34 G M L

40

30
Porosity, %

20 L, 19.0 L, 18.3 L, 22.6


L, 19.2

10
M, 8.6 M, 8.0 M, 7.4 M, 6.3

G, 2.3 G, 2.8 G, 2.3 G, 2.8


0
PCE SNF PLS UNA

50
(b) W/C = 0.40 G M L

40

30 L,
Porosity, %

L, L, 23.5
21.2 L, L, 26.6 L,
25.4 28.9 27.6
20

10 M, 11.9
M, 10.9
M, 10.1
M, 8.7 M, 6.9
M, 7.5

G, 4.0 G, 2.9 G, 3.2 G, 4.5


G, 2.7 G, 2.2
0
PCE SNF PLS UNA BCS BCA

Fig. 4-30 Total porosities and pore size distributions of the pastes at 3 days

158
Chapter 4 Results and Discussion

50

(a) W/C = 0.34 G M L

40

30
Porosity, %

20 L, 13.4 L, 13.1
L, 16.8
L, 15.8

10
M, 10.3 M, 11.1
M, 7.0 M, 6.4

G, 2.2 G, 3.2 G, 3.0 G, 3.2


0
PCE SNF PLS UNA

50
(b) W/C = 0.40 G M L

40

30
Porosity, %

L, L,
L, L, L, L, 19.2
18.9 21.5
19.4 21.8 20.2
20

10 M, 10.0 M, 11.3
M, 8.7 M, 9.0
M, 8.6 M, 9.8

G, 4.4 G, 4.4 G, 3.1 G, 2.8 G, 3.6 G, 3.9


0
PCE SNF PLS UNA BCS BCA

Fig. 4-31 Total porosities and pore size distributions of the pastes at 7 days

159
Chapter 4 Results and Discussion

50

(a) W/C = 0.34 G M L

40

30
Porosity, %

20
L, 13.4 L, 13.0
L, 16.1
L, 12.7
10
M, 9.1 M, 9.1
M, 4.8 M, 4.8
G, 2.3 G, 1.9 G, 3.0 G, 2.6
0
PCE SNF PLS UNA

50
(b) W/C = 0.40 G M L

40

30
Porosity, %

L, L, L, L,
L, 18.4 L,
14.1 16.5 15.8
20 16.9 16.8

10 M, 9.3
M, 9.8 M, 9.3
M, 9.5 M, 10.0 M, 8.6

G, 5.8 G, 4.5
G, 3.6 G, 3.1 G, 3.2 G, 3.8
0
PCE SNF PLS UNA BCS BCA

Fig. 4-32 Total porosities and pore size distributions of the pastes at 28 days

160
Chapter 4 Results and Discussion

50

(a) W/C = 0.34 G M L

40

30
Porosity, %

20
L, 12.5
L, 11.6
L, 16.2
L, 12.5
10
M, 8.1 M, 9.2
M, 4.1 M, 4.4
G, 2.9 G, 1.9 G, 3.0 G, 2.6
0
PCE SNF PLS UNA

50
(b) W/C = 0.40 G M L

40

30
Porosity, %

L, L, L, L,
20 16.1 L,
13.9 14.3 L, 12.5
13.2 12.8

10
M, 8.4 M, 9.8 M, 8.8
M, 9.0 M, 9.2 M, 8.4

G, 4.5 G, 3.3 G, 3.3 G, 2.9 G, 4.4


G, 2.7
0
PCE SNF PLS UNA BCS BCA

Fig. 4-33 Total porosities and pore size distributions of the pastes at 91 days

161
Chapter 4 Results and Discussion

80

PCE SNF
PLS UNA
60
Compressive Strength, MPa

40

20

0
1d 3d 7d 28d 91d
Age

Fig. 4-34 Compressive strength of 50mm mortar cubes (w/c = 0.34)

80

PCE SNF PLS


UNA BCS BCA
60
Compressive Strength, MPa

40

20

0
1d 3d 7d 28d 91d
Age

Fig. 4-35 Compressive strength of 50mm mortar cubes (w/c = 0.40)

162
Chapter 5 Conclusions and Recommendations

5.1 Conclusions

The focus of this study is a newly-developed LS based superplasticizer (PLS). The

main objectives are to investigate this LS superplasticizer (PLS) in comparison with

polycarboxylate (PCE), naphthalene (SNF) and the other modified LS (UNA) based

superplasticizers and with regular LS water reducing admixtures (BCS and BCA).

The mixtures were designed to achieve similar workability, thus the type and dosage

of the admixtures have influence on various properties. Based on the laboratory

results, the following conclusions may be drawn:

1. It is clear that the water reducing admixtures and superplasticizers delayed

cement hydration for both w/c ratios at early ages. However, the degree of

retardation is different for admixtures. The retardation of the pastes are in

the order of SNF < PCE < PLS < UNA < BCA < BCS based on the rate of

heat and length of the induction period. This retardation order of pastes

with admixtures at early ages is supported by cumulative heat, C3S

consumption and CH precipitation.

163
Chapter 5 Conclusions and Recommendations

2. The test results, including C3S consumption, CH precipitation and non-

evaporable water content and compressive strength, suggest that the

admixtures did not have significant effect on the cement hydration at late

ages, from 7 to 91 days.

3. From yield stress determined by BML coaxial rheometer, the workability

loss of mortars with LS superplasticizers (PLS and UNA) were similar

within the first hour, but were less than those with SNF and PCE

superplasticizers. The workability loss of the cement pastes with two

regular water reducing admixtures (BCS and BCA) was more significant

than those with the superplasticizers in the same period of time. This

corresponds to the observation made from the heat development that the

BCS and BCA admixtures accelerated cement hydration at the very

beginning.

4. For the two modified LS based superplasticizers, plastic viscosities remain

relatively unchanged within the first hour for both w/c ratios whereas

mortars with PCE and SNF superplasticizers had some variation in plastic

viscosities. Nevertheless, the changes of plastic viscosity with time were

much less than those of the yield stress.

5. The order of setting times of the mortars with admixtures agree with the

length of induction periods in the heat curves of the respective pastes, i.e.

SNF < PCE < PLS < UNA< BCA < BCS. The admixtures have strong

influence on the initial setting times. However, once the mortars reached

the initial setting, the final setting was not significantly affected by the

admixtures.

6. For the four superplasticizers, better workability retention corresponds to

164
Chapter 5 Conclusions and Recommendations

longer setting time. However, the two regular LS based water reducing

admixtures had longer setting time, but poor workability retention which is

probably related to their acceleration of cement hydration during the first

hour according to the rate of heat curves.

7. At 28 and 91 days, the porosity of the paste with the LS superplasticizers

at w/c of 0.34 was similar to that with the SNF admixture, but was higher

than that with the PCE admixture. At w/c of 0.40, the total porosity of the

pastes with different admixtures was not significantly different at 28 days.

8. Pore size distributions of the pastes changed with time due to cement

hydration and they differed with respect to w/c ratios and admixtures. In

general, the proportions of small capillary pores in the pastes investigated

were not significantly different, and the differences were mainly in the

large and medium capillary pores. The pastes with LS superplasticizers

had similar large pores at 91 days compared to the paste with PCE

admixture, but less large pores compared to the paste with SNF admixture

at both w/c ratios. However, the pastes with LS superplasticizers had more

medium pores compared to the pastes with PCE and SNF admixtures. The

pastes with the regular LS admixtures (BCS and BCA) appeared to have

less large capillary pores at 91 days compared to those with the

superplasticizers. The threshold and critical pore diameters of the pastes

were not significantly affected by the admixtures.

9. The chemical admixtures investigated affected early compressive strength

of the mortars due to their different retarding effects. However, the

strength of the mortars was not significantly affected by the admixtures

beyond 7 days.

165
Chapter 5 Conclusions and Recommendations

10. The performances of regular and modified LS based admixtures were

different at the early ages but similar at 28 and 91 days. The regular LS

admixtures retarded cement hydration much more than the modified LS

superplasticizers. It was reflected on the heat generation of cement pastes

and setting times of respective mortars. On the other hand, the mortars

with regular LS admixtures had much higher rate of workability loss than

those with the modified LS superplasticizers.

5.2 Recommendations

Further studies are recommended in the following areas:

1. The current studies, including the heat evolution, C3S consumption, CH

production, NEW content; were concentrated on degree of hydration. It is seen

that water reducing admixtures delayed cement hydration and it also

demonstrated that these admixtures delayed C3S hydration to different degrees.

But it remains unclear how they influence the hydration rate of other clinker

phases, particularly C3A in cement at early age. Further studies of the

influences of these admixtures on the C3A and other relevant phases at early

age are recommended.

2. Microstructures of fresh cement paste with and without admixtures are of

great importance to understand cement hydration. Further study of the early-

age microstructures including the morphology – the shape and the size, would

provide a clear image of the anhydrated cement particles and hydration

products such as ettringite.

3. A study of the microstructures of hardened cement paste with and without

admixtures would be of great help to better understand cement hydration at

166
Chapter 5 Conclusions and Recommendations

late ages. Microstructures at late stages include distribution of hydration

products and the pore structure of the paste samples.

4. Since the admixtures affect the pore structure of the cement pastes as

discussed before, it is important to determine how the differences in the pore

structure influence permeability and resistance of concrete to chloride ion

penetration, which affect long-term durability of reinforced concrete structures.

167
References

Agarwal S. K., Masood I., Malhotra S. K. (2000), Compatibility of superplasticizers


with different cements. Construction and Building Materials, Vol. 14, No. 5, pp. 253-
259.

Aïtcin P.-C., Sarkar S. L., Regourd M. and Volant D. (1987), Retardation effect of
superplasticizer on different cement fractions. Cement and Concrete Research, Vol. 17,
No. 6, pp. 995-999.

Alarcon-Ruiz L., Platret G., Massieu E. and Ehrlacher A. (2005), The use of thermal
analysis in assessing the effect of temperature on a cement paste. Cement and
Concrete Research, Vol. 35, No. 3, pp. 609-613.

American Society for Testing and Materials, ASTM C109 / C109M–99 Standard test
method for compressive strength of hydraulic cement mortars, Washington, D. C..

American Society for Testing and Materials, ASTM C128–01 Standard test method
for density, relative density (specific gravity), and absorption of fine aggregate,
Washington, D. C..

American Society for Testing and Materials, ASTM C136–01 Standard test method
for sieve analysis of fine and coarse aggregates, Washington, D. C..

American Society for Testing and Materials, ASTM C138-92 Standard test method for
Unit Weight, Yield, and Air Content (Gravimetric) of Concrete, Washington, D. C..

American Society for Testing and Materials, ASTM C150-01, Standard specification
for Portland cement, Washington, D. C..

American Society for Testing and Materials, ASTM C230/C230M-03, Standard


specification for Flow Table for Use in Tests of Hydraulic Cement, Washington, D. C..

American Society for Testing and Materials, ASTM C403/C403M-99, Standard test
method for time of setting of concrete by penetration resistance, Washington, D. C..

American Society for Testing and Materials, ASTM C494, Standard specification for
chemical admixtures for concrete, Washington, D. C..

American Society for Testing and Materials, ASTM C1017/C1017M-98, Standard


specification for Chemical Admixtures for Use in Producing Flowing Concrete,
Washington, D. C..

American Society for Testing and Materials, ASTM C1437-99, Standard test method
for flow of hydraulic cement mortar, Washington, D. C..

168
References

Balasubramanian K., Krishnamoorthy T. S. and Sellevold E. J. (1997), “Effects of


drying procedure on the capillary suction in concrete and upon the exchange of pore
water with ethanol”, The Indian Concrete Journal, Vol. 71, No. 2, pp.105-109.

Banfill P. F. G. (2003), The rheology of fresh cement and concrete – a review. The 11th
International Cement Chemistry Congress, Durban.

Beaudoin J. J., Gu P., Marchand J., Tamtsia B., Myers R. E. and Liu Z. (1998),
Solvent Replacement Studies of Hydrated Portland Cement Systems: The Role of
Calcium Hydroxide. Advanced Cement Based Materials, Vol. 8, No. 2, pp. 56-65.

Bensted J. (1987), Some applications of conduction calorimetry to cement hydration.


Advances in Cement Research, Vol. 1, No. 1, pp. 35-44.

Berger R. L. and McGregor J. D. (1972), Influence of admixtures on the morphology


of calcium hydroxide formed during tricalcium silicate hydration. Cement and
Concrete Research, Vol. 2, pp. 43-55.

Bishop M. and Barron A. R. (2006), Cement Hydration Inhibition with Sucrose,


Tartaric Acid, and Lignosulfonate: Analytical and Spectroscopic Study. Industrial &
Engineering Chemistry Research, Vol. 45, pp.7042-7049.

Björnström J. and Chandra S. (2003), Effect of superplasticizers on the rheological


properties of Cements. Materials and Structures, Vol. 36, No. 264, pp. 685-692.

Bonen D. and Sarkar S. L. (1995), The superplasticizer adsorption capacity of cement


pastes, pore solution composition, and parameters affecting flow loss. Cement and
Concrete Research, Vol. 25, No. 7, pp. 1423– 1434.

Borregaard (2006), The admixture handbook – Borregaard’s handbook on


lignosulfonates and concrete admixtures. Borregaard LignoTech, Sarpsborg, Norway.

Bragg W. L. (1914), The Diffraction of Short Electromagnetic Waves by a Crystal.


Proceedings of the Cambridge Philosophical Society, Vol. 17, pp. 43–57.

Carazeanu I., Chirila E. and Georgescu M. (2002), Investigation of the hydration


process in 3CaO·Al2O3-CaSO4 · 2H2O-plasticizer-H2O systems by x-ray diffraction.
Talanta, Vol. 57, No. 4, pp. 617-623.

Chan Y. N., Feng N.-Q. and Tsang K. C. (1996), Workability retention of high-
strength/superplasticized concrete. Magazine of Concrete Research, Vol. 48, No. 177,
pp. 301-309.

Chandra S. and Björnström J. (2002), Influence of superplasticizer type and dosage on


the slump loss of Portland cement mortars—Part II. Cement and Concrete Research,
Vol. 32, No. 10, pp. 1613–1619.

Chen C.-T. (2007), Interactions between Portland cements and carboxylated and
naphthalene-based superplasticizers. Ph.D DISSERTATION, University of Illinois at
Urbana-Champaign, 426 p.

169
References

Chotard T., Gimet-Breart N., Smith A., Fargeot D., Bonnet J. P. and Gault C. (2001),
Application of ultrasonic testing to describe the hydration of calcium aluminate
cement at the early age. Cement and Concrete Research, Vol.31, No.3, pp.405-412.

Ciach T. D. and Swenson E. G. (1971), Morphology and microstructure of hydrating


Portland cement and its constituents IV. Changes in hydration of A C3S, C3S, C3A,
C4AF and gypsum paste with and without the admixtures triethanolamine and calcium
lignosulphonate. Cement and Concrete Research, Vol. 1, No. 4, pp. 367-383.

Collepardi M. (1993), The world of chemical admixtures. 18th congress on our world
in concrete and structures, Singapore, pp. 63-71.

Collepardi M. (1998), Admixtures used to enhance placing characteristics of concrete.


Cement and Concrete Composites, Vol. 20, pp. 103–112.

Collepardi M., Corradi M. and Baldini G. (1980), hydration of C3A in the Presence of
Lignosulfonate-Carbonate System or Sulfonated Naphthalene Polymer. 7th
International Congress on the Chemistry of Cement, pp. 524-528.

ConTec Ltd. (2003), The ConTec BML Viscometer 4 & 5 Operating Manual, 123 p.

Cook R. A. (1991), Fundamentals of mercury intrusion porosimetry and its


application to concrete materials science, Master of Science Thesis, Cornell
University.

Cook R. A. and Hover K. C. (1999), Mercury porosimetry of hardened cement pastes.


Cement and Concrete Research, Vol. 29, No. 6, pp. 933–943.

Coole M. J. (1984), Calorimetric Studies of the Hydration Behaviour of Extended


Cements, British Ceramic Proceedings, No. 35, pp. 385-401.

Day R. L. (1981), Reactions between methanol and Portland cement paste. Cement
and Concrete Research, Vol. 11, No. 3, pp. 341-349.

Daugherty K. E. and Kowalewski Jr. M. J (1968), Effects of Organic Compounds on


the Hydration Reactions of Tricalcium Aluminate. Proceedings of the 5th International
Symposium on the Chemistry of Cement, Vol. IV, Tokyo, pp. 42-52.

Diamond S. (2000), Mercury porosimetry - An inappropriate method for the


measurement of pore size distributions in cement-based materials. Cement and
Concrete Research, Vol. 30, No. 10, pp. 1517-1525.

Diamond S. and Leeman M. E. (1995), Pore size distributions in hardened cement


paste by image analysis, in Microstructure of Cement Based Systems/Bonding and
Interfaces in Cementitious Materials. Materials Research Society Symposium
Proceedings, Vol. 370, pp. 217-226.

Dodson V.H. (1967). Proceedings of the International Symposium on Admixtures for


Mortar and Concrete, Brussels, pp. 59–64.

170
References

Emoto T. and Bier T. A. (2007), Rheological behavior as influenced by plasticizers


and hydration kinetics. Cement and Concrete Research, Vol. 37, No. 5, pp. 647–654.

Falikman V. R., Sorokin Y.V., Vainer A. Y. and Bashlykov N. F. (2005), New High
Performance Polycarboxilate Superplasticizers based on Derivative Copolymers of
Maleinic Acid in Admixture – enhancing concrete performance (Eds. Ravindra et al).
Proceedings of the International Conference, University of Dundee, UK, pp. 41-46.

Farrington S. A. (2007), Evaluating the effect of mixing method on cement hydration


in the presence of a polycarboxylate high-range water reducing admixture by
isothermal conduction calorimetry. 12th International Congress on the Chemistry of
Cement.

Ferraris C. F. and Martys N. S. (2003), Relating fresh concrete viscosity


measurements from different rheometers. Journal of Research of the National Institute
of Standards and Technology, Vol. 108, No. 3, pp. 229-234.

Ferraris C. F. and Brower L. E. (Eds., 2001), Comparison of concrete rheometers:


International tests at LCPC (Nantes, France) in October 2000. National Institute of
Standards and Technology, NISTIR 6819.

Ferraris C. F. and de Larrard F. (1998), Testing and Modelling of Fresh Concrete


Rheology, National Institute of Standards and Technology, NISTIR 6094.

Flatt R. J., Houst Y. F., Bowen P. and Hofmann H. (2000), Electro steric repulsion
induced by superplasticizers between cement particles – an overlooked mechanism?,
Proceedings of the 6th CANMET/ACI International Conference on Superplasticizer
and Other Chemical Admixtures in Concrete, Nice, France. SP 195-33, pp. 29-42.

Ftikos C. and Philippou T. (1990), Preparation and hydration study of rich C2S
cements.
Cement and Concrete Research, Vol. 20, No. 6, pp. 934-940.

Gallé C., (2001) Effect of drying on cement-based materials pore structure as


identified by mercury intrusion porosimetry. A comparative study between oven-,
vacuum-, and freeze-drying. Cement and Concrete Research, Vol. 31, No. 10,
pp.1467-1477.

Gjørv O. E. (1994), Workability: A New Way of Testing. Concrete International, Vol.


20, No. 9, pp. 57-60.

Gołaszewski J. G. and Szwabowski J. (2004), Influence of superplasticizers on


rheological behaviour of fresh cement mortars. Cement and Concrete Research, Vol.
34, No. 2, pp. 235–248.

Govin A., Peschard A. and Guyonnet R. (2006), Modification of cement hydration at


early ages by natural and heated wood. Cement and Concrete Composites, Vol. 28, No.
1, pp. 12-20.

Grattan-Bellew P. E. (1996), Microstructural investigation of deteriorated Portland

171
References

cement concretes. Construction and Building Materials, Vol. 10, No. 1, pp. 3-16.

Gruskovnjak A., Lothenbach B., Holzer L. and Winnefeld F. (2006), Hydration of


alkali-activated slag: comparison with ordinary Portland cement. Advances in Cement
Research, Vol. 18, No. 3, pp. 119-128.

Gu D., Xiong D. and Lu Z. (1982), Mode of mechanism of naphthalene series on


water reducing agents. Journal of the American Concrete Institute, Vol. 69, pp. 378-
386.

Gu P., Xie P., Beaudoin J. J. and Jolicoeur C. (1994), Investigation of the retarding
effect of superplasticizers on cement hydration by impedance spectroscopy and other
methods. Cement and Concrete Research, Vol. 24, No. 3, pp. 433-442.

Haines P. J. (2002), Principles of thermal analysis and calorimetry. Cambridge: Royal


Society of Chemistry, 220 p.

Halamickova P., Detwiler R. J., Bentz D. P. and Garboczi E. J. (1995), Water


permeability and chloride ion diffusion in portland cement mortars: relationship to
sand content and critical pore diameter. Cement and Concrete Research, Vol. 25, No. 4,
pp. 790-802.

Hanehara S. and Yamada K. (1999), Interaction between cement and chemical


admixture from the point of cement hydration, absorption behaviour of admixture,
and paste rheology. Cement and Concrete Research, Vol.29, No. 8, pp. 1159-1165.

Hanehara S. and Yamada K. (2007), Rheology and Early Age Properties of Cement
Systems – Part 1. 12th International Congress on the Chemistry of Cement.

Hansen W. C. (1959), Actions of Calcium Sulfate and Admixtures in Portland Cement


Pastes, in Proceedings of the Symposium on Effect of Water-Reducing Admixtures
and Set-Retarding Admixtures on Properties of Concrete, ASTM STP-266, pp. 3-25.

Hansen T. C. (1986), Physical Structure of Hardened Cement Paste - A Classical


Approach. Materials and Structures, Vol. 19, No. 114, pp. 423-436.

He J., Pang H., Zhang X.-W. and Liao B. (2005), A new polycarboxylic acid-based
superplasticizer with poly (ethylene oxide) graft chains. Polymeric Materials Science
and Engineering, Vol. 21, No. 5, pp. 44-50. (In Chinese)

Heikal M., Morsy M. S. and Aiad I. (2006), effect of polycarboxylate superplasticizer


on hydration characteristics of cement pastes containing silica fume. Ceramics −
Silikáty, Vol. 50, No. 1, pp. 5-14.

Hekal E. E. and Kishar E. A. (1999), Effect of sodium salt of naphthalene-


formaldehyde polycondensate on ettringite formation. Cement and Concrete Research,
Vol. 29, pp. 1535–1540.

Hewlett P. C. (1998), Lea's Chemistry of Cement and Concrete, 4th Edition. London:
Arnold, 1053 p.

172
References

Houst Y. F., Bowen P. and Perche, F. (2005), Towards tailored superplasticizers.


Admixtures - Enhancing Concrete Performance (Dhir R. K., Hewlett P. C., Newlands
M. D., Eds), Thomas Telford, London, pp. 11-20.

Houst Y. F., Flatt R. J., Bowen P., Hofmann H. (1999), Optimisation of


Superplasticisers: From Research to Application. Modern Concrete materials: binders,
additions and Admixtures (Eds. Dhir R. K. and Dyer T. D., pp. 445-456.

http://ciks.cbt.nist.gov/bentz/phpct/database/thermal.html, accessed on 7 July 2007.

http://durhamgeo.com/testing/concrete/cement-flowtable.htm, accessed on 6 May


2007.

Hwang L. and Lee J. C. (1989), The effect of NF superplasticizer on the micro and
macro properties of concrete material. 3rd CANMET/ACI international conference on
superplasticizers and other admixtures in concrete (Ed. Malhotra V. M.). SP-119.

Illston J. M. and Domone P. L. J. (2001), Construction Materials: their nature and


behavior. Taylor & Francis Group, 584 p.

Jolicoeur C. and Simard M.-A. (1998), Chemical admixture-cement interactions:


Phenomenology and physico-chemical concepts. Cement and Concrete Composites,
Vol. 20, Nos. 2-3, pp.87-101.

Khalil Kh. A. (1999), Surface area and pore structure of hardened Portland
cement/silica fume pastes containing a superplasticizer. Adsorption Science and
Technology, Vol.17, No. 7, pp. 557-563.

Khalil S. M. and Ward M. A. (1973), Influence of a lignin based admixture on the


hydration of portland cements. Cement and Concrete Research, Vol. 3, No. 6, pp. 677-
688.

Khoury G. A. (1992), Compressive strength of concrete at high temperatures: a


reassessment. Magazine of Concrete Research, Vol. 44, pp. 291-309.

Kjellsen K. O. and Diamond S. (2007), Investigations into the microstructure of fresh


Portland cement mortar. 12th International Congress on the Chemistry of Cement,
Montreal, Canada.

Koizumi K., Umemura Y. and Tsuyuki N. (2007), Effects of chemical admixtures on


the silicate structure of hydrated Portland cement. 12th International Congress on the
Chemistry of Cement.

Kreppelt F., Weibel M., Zampini D., Romer M.(2002), Influence of solution chemistry
on the hydration of polished clinker surfaces—a study of different types of
polycarboxylic acid-based admixtures. Cement and Concrete Research, Vol. 32, No. 2,
pp.187–198.

Lamond J. F. and Pielert J. H. (Eds., 2006), Significance of tests and properties of


concrete and concrete-making materials. West Conshohocken, PA, revised ed., 664 p.

173
References

Law Chee Meng (2004), Effect of water reducing admixtures on the resistance of
concrete to chloride ion penetration. Final Year Thesis, Department of Civil
Engineering, National University of Singapore.

Lei W.-G. and Struble L. (1997), Microstructure and flow behavior of fresh cement
paste, Journal of American Ceramic Society, Vol. 80, No. 8, pp. 2021–2049.

Li C., Feng N., Wang D. and Huo Y. (2005), Preparation and characterization of
comb-like polycarboxylic water-reducers and its function mechanism. Journal of the
Chinese Ceramic Society, Vol. 33, No. 1, pp. 87-92. (In Chinese)

Lilkov V., Dimitrova E. and Petrov O. E. (1997) Hydration process of cement


containing fly ash and silica fume: The first 24 hours. Cement and Concrete Research,
Vol. 27, No. 4, pp. 577-588.

Lim G.-G., Hong S.-S., Kim D.-S., Lee B.-J. , Rho J.-S. (1999), Slump loss control of
cement paste by adding polycarboxylic type slump-releasing dispersant. Cement and
Concrete Research, Vol. 29, No. 2, pp. 223-229.

Lombois-Burger H., Guillot L., and Haehnel C. (2006), SP-239-24: Interaction


between Cements and Superplasticizers. 8th CANMET/ACI International Conference
on Superplasticizers and Other Chemical Admixtures in Concrete, pp. 357-373.

Lothenbach B., Winnefeld F. and Figi R. (2007), The influence of superplasticizers on


the hydration of Portland cement. 12th International Congress on the Chemistry of
Cement.

Malhotra V. M. (Ed., 1997), 5th CANMET/ACI International Conference on


Superplasticizers and Other Chemical Admixtures in Concrete (SP-173), ACI,
Farmington Hill, Mi, USA.

Mansoutre S. and Lequex N. (1996), Quantitative phase analysis of Portland cements


from reactive powder concretes by X-ray powder diffraction. Advances in Cement
Research, Vol. 8, No. 32, pp. 175–182.

Marchand J. (1993), Contribution to the study of the scaling deterioration of concrete


in presence of deicing salts, Ph. D. Thesis, École Nationale des Ponts et Chaussées
(ENPC), Paris, France, 326 p.

Mikanovic N., Simard M. A. and Jolicoeur C. (2000) Interaction between PNS-type


superplasticizers and cement during initial hydration, 6th CANMET/ACI International
Conference on Superplasticizers and Other Chemical Admixtures in Concrete, Nice,
ACI SP-195, pp. 561-576.

Mindess S., Young J. F. and Darwin D. (2003), Concrete, Prentice-Hall, 2nd ed. 644 p.

Mollah M.Y.A., Yu W., Schennach R., Cocke D. L. (2000), A Fourier transform


infrared spectroscopic investigation of the early hydration of Portland cement and the
influence of sodium lignosulphonate. Cement and Concrete Research, Vol. 30, No. 2,
pp. 267-273.

174
References

Mollah M.Y.A., Palta P., Hess T. R., Vempati R. K. and Cocke D. L. (1995), Chemical
and physical effects of sodium lignosulphonate superplasticizer on the hydration of
portland cement and solidification/stabiization consequences. Cement and Concrete
Research, Vol. 25, No. 3, pp. 671-682.

Myrvold B. O. (2006), SP-239-19: Adsorption of Lignosulphonates on Cement and


the Hydration Products of Cements. 8th CANMET/ACI International Conference on
Superplasticizers and Other Chemical Admixtures in Concrete (Ed.: V.M. Malhotra),
pp. 285-296.

Myrvold B. O. (2007), Interactions between lignosulfonates and clinker minerals and


the hydration products of clinker minerals. 12th International Congress on the
Chemistry of Cement.

Nakajima Y. and Yamada K. (2004), The effect of the kind of calcium sulfate in
cements on the dispersing ability of poly h-naphthalene sulfonate condensate
superplasticizer. Cement and Concrete Research, Vol. 34, No. 5, pp. 839–844.

Nawa T., Ichiboji H., and Knoshita M. (2000), Influence of Temperature on Fluidity
of Cement Paste Containing Superplasticizer with Polyethylene Oxide Graft Chains.
The Sixth CANMET/ACI Conference on Superplasticizers and Other Chemical
Admixtures in Concrete, SP-195, pp. 195-210.

Newman J. and Choo B. S. (2003), Advanced Concrete Technology 1: Constituent


Materials. Butterworth-Heinemann, Elsevier, Ltd.

Nkinamubanzi P.-C. and Aïtcin P.-C. (2004), Cement and superplastcizer


combinations: compatibility and robustness. Cement, Concrete and Aggregates, Vol.
26, no. 2, pp. 102-109.

Nuffield E. W. (1966), X-ray Diffraction Methods, Wiley, New York, pp. 137– 149.
Odler I. and Becker Th. (1980), Effect of some liquefying agents on properties and
hydration of portland cement and tricalcium silicate pastes. Cement and Concrete
Research, Vol. 10, No. 3, pp. 321-331.

Odler I. and Samir A.-M. (1987), Effect of chemical admixtures on Portland cement
hydration. Cement Concrete and Aggregates, Vol. 9, No. 1, pp. 38-43.

Ouyang X., Qiu X. and Chen P. (2006), Physicochemical characterization of calcium


lignosulfonate—A potentially useful water reducer. Colloids and Surfaces A:
Physicochemical and Engineering Aspects, Vol. 282–283, pp. 489–497.

Pang Y., Lou H., Qiu X. and Yang D. (2005), Influences of modified lignosulfonate
superplasticizer on cement hydration and the durability of concrete. Journal of
Sichuan University (Engineering Science Edition), Vol. 37, No. 1, pp. 74-77. (In
Chinese)

Pang Y., Qiu X. and Yang D. (2005a), Influence of calcium lignosulfonate on cement
hydration. Journal of the Chinese Ceramic Society, Vol. 33, No. 4, pp. 477-83. (In
Chinese)

175
References

Parrott L. J. and Killoh D. C. (1984), Prediction of cement hydration, the chemistry


and chemically related properties of cement. British Ceramic Proceedings, No. 35, pp.
41–53.

Parrott L. J., Geiker M., Gutteridge W. A. and Killoh D. (1990), Monitoring Portland
cement hydration: comparison of methods. Cement and Concrete Research Vol. 20,
No. 6, pp. 919-926.

Peschard A., Govin A., Grosseau P., Guilhot B. and Guyonnet R. (2004), Effect of
polysaccharides on the hydration of cement paste at early ages. Cement and Concrete
Research, Vol. 27, No. 4, pp. 2153-2158.

Petit J.-Y., Khayat K. H. and Wirquin E. (2006), Coupled effect of time and
temperature on variations of yield value of highly flowable mortar. Cement and
Concrete Research, Vol. 36, No. 5, pp. 832 – 841.

Pourchet S., Comparet C., Nonat A., and Maitrasse P.(2006), SP-239—11: Influence
of Three Types of Superplasticizers on Tricalciumaluminate Hydration in Presence of
Gypsum. 8th CANMET/ACI International Conference on Superplasticizers and Other
Chemical Admixtures in Concrete (Ed.: V.M. Malhotra), pp. 151-167.

Prince W., Edwards-Lajnef M. and Aïtcin P.-C. (2002), Interaction between ettringite
and a polynaphthalene sulfonate superplasticizer in a cementitious paste. Cement and
Concrete Research, Vol. 32, No. 1, pp. 79-85.

Prince W., Espagne M. and Aïtcin P.-C. (2003), Ettringite formation: A crucial step in
cement superplasticizer compatibility. Cement and Concrete Research, Vol. 33, No. 5,
pp. 635–641.

Puertas F., Santos H., Palacios M. and Martinez-Ramirez S. (2005), Polycarboxylate


superplasticiser admixtures: effect on hydration, microstructure and rheological
behaviour in cement pastes. Advances in Cement Research, Vol. 17, No. 2, pp. 77–89.

Ramachandran V. S. (1979), Differential thermal method of estimating calcium


hydroxide in calcium silicate and cement pastes. Cement and Concrete Research, Vol.
9, pp. 677-684.

Ramachandran V. S. (1993), Recent developments in concrete admixtures


formulations, Il Cemento, Vol. 90, pp. 11-24.

Ramachandran V. S. (1995), Concrete admixtures handbook: properties, science, and


technology. Park Ridge, N.J., U.S.A. : Noyes Publications, 2nd ed., 1153 p.

Ramachandran V. S. and Beaudoin J. J. (1999), Handbook of analytical techniques in


concrete science and technology. Norwich, N.Y.: Noyes Publications, 964 p.

Ramachandran V. S., Malhotra V. M., Jolicoeur C., and Spiratos N. (1998),


Superplasticizers: Properties and Applications in Concrete. CANMET Publication
MTL 97-14, CANMET, Ottawa, 404 p.

176
References

Ramachandran V. S., Paroli R. M., Beaudoin J. J., Delgado A. H. (2003), Thermal


analysis of construction materials. William Andrew Publishing/Noyes Publications,
680 p.

Reiner M. (1949), Deformation and flow: H.K. Lewis and Co., London, 346 p.

Reknes K. (2004), The performance of a new lignosulphonate plasticizing admixture


and its sensitivity to cement composition, 29th Conference on Our World In Concrete
and Structures, Singapore, pp. 441-447.

Reknes K. and Peterson B. G. (2003), Self-Compacting Concrete with


Lignosulphonate Based Superplasticizer. 3rd International Symposium on Self-
Compacting Concrete, Reykjavik, pp.184-189.

Rickert J. and Thielen G. (2004), Influence of a long-term retarder on the hydration of


clinker and cement. Cement, Concrete, and Aggregates, Vol. 26, No. 2, pp. 92-101.

Ridi F., Dei L., Fratini E., Chen S.-H., and Baglioni P. (2003), Hydration Kinetics of
Tri-calcium Silicate in the Presence of Superplasticizers. The Journal of Physical
Chemistry B, Vol. 107, No. 4, pp. 1056-1061.

Rixom M. R. and Mailvaganam N. (1999), Chemical admixtures for concrete. London:


E & FN Spon; New York: Routledge. 3rd ed., 437 p.

Roncero, J., Valls, S., and Gettu, R. (2002), Study of the influence of superplasticizers
on the hydration of cement paste using nuclear magnetic resonance and X-ray
diffraction techniques. Cement and Concrete Research, Vol. 32, No. 1, pp. 103-108.

Rößler C., Eberhardt A., Kučerová H. and Möser B. (2007), Influence of Hydration on
the Fluidity of Normal Portland Cement Pastes. 12th International Congress on the
Chemistry of Cement.

Sakai E., Kasuga T., Sugiyama T., Asaga K. and Daimon M. (2006), Influence of
superplasticizers on the hydration of cement and the pore structure of hardened
cement. Cement and Concrete Research, Vol. 36, No. 11, pp. 2049–2053.

Sandberg P. (2004), Optimization of Cement Sulfate, Part II - Cement with admixture.


Thermometric, Application Note AN314-06.

Sarkar S. L. and Xu A. (1992), Preliminary study of very early hydration of


superplasticized C3A+gypsum by environmental SEM. Cement and Concrete
Research, Vol.22, No.4, pp. 605-608.

Scarlett N. V. Y., Madsen I. C., Manias C. and Retallack D. (2001),On-line X-ray


diffraction for quantitative phase analysis: Application in the Portland cement industry.
Powder Diffraction, Vol. 16, No. 2, pp. 71-80.

Scrivener K. L. (1997), Microscopy methods in cement and concrete science. World


Cement Research and Development, September, 1997, pp. 92-112.

177
References

Scrivenera K. L., Fullmanna T., Galluccia E., Walentab G. and Bermejob E. (2004),
Quantitative study of Portland cement hydration by X-ray diffraction/Rietveld
analysis and independent methods. Cement and Concrete Research, Vol. 34, No. 9, pp.
1541–1547.

Simard M.-A., Nkinamubanzi P.-C., Jolicoeur C., Perraton D. and Aïtcin P.-C. (1993),
Calorimetry, rheology and compressive strength of superplasticized cement pastes.
Cement and Concrete Research, Vol. 23, No. 4, pp. 939-950.

Sugamata T., Edamatsu Y., Ouchi M. (2003), A study of particle dispersing retention
effect of polycarboxylate-based superplasticizers. Proceedings of the 3rd International
RILEM Symposium (Eds. O. Wallevik and I. Nielsson), pp. 420-431.

Tagnit H. A. and Sarkar S. L. (1990), Influence of varying sulphur content on the


microstructure of commercial clinkers and the properties of cement. World Cement,
Vol. 21, No. 9, pp. 389-393.

Taylor H. F. W. (1997), Cement chemistry. London: T. Telford, 2nd ed., 459 p.

Taylor H. F. W. and Turner A. B. (1987), Reactions of tricalcium silicate paste with


organic liquids. Cement and Concrete Research, Vol. 17, No. 4, pp. 613-623.

Uchikawa H., Hanehara S. and Sawaki D. (1997), The role of steric repulsive force in
the dispersion of cement particles in fresh paste prepared with organic admixture.
Cement and Concrete Research, Vol. 27, No. 1, pp. 37−50.

Uchikawa H., Hanehara S., Shirasaka T. And Sawaki D. (1992), Effect of admixture
on hydration of cement adsorptive behavior of admixture and fluidity and setting of
fresh cement paste. Cement and Concrete Research, Vol. 22, No. 6, pp. 1115-1129.

Uchikawa H., Ogawa K. and Uchida S. (1983), Effects of admixtures on the rheology
of fly ash cements, Japan Cement Association, Cement and Concrete, Vol. 37, pp. 53–
56.

Uchikawa H., Sawaki D. and Hanehara S. (1995), Influence of kind and added timing
of organic admixture on the composition, structure and property of fresh cement paste.
Cement and Concrete Research, Vol. 25, No. 2, pp. 353-364.

Uchikawa H., Uchida S., Ogawa K. and Hanehara S. (1984), Influence of


CaSO4.2H2O, CaSO4.0.5H2O, CaSO4 on the initial hydration of clinker having
different burning degree. Cement and Concrete Research, Vol.14, No. 5, pp. 645-656.

Vikan H. (2005), Rheology and reactivity of cementitious binders with plasticizers.


Doctoral thesis at Norwegian University of Science and Technology (NTNU), 332 p.

Vikan H. and Justnes H. (2007), Influence of cement and plasticizer type on the heat
of hydration. 12th International Congress on the Chemistry of Cement.

Wallevik J. E. (2003), Rheology of Particle Suspensions - Fresh Concrete, Mortar and


Cement Paste with Various Types of Lignosulfonates, (Ph.D. Dissertation);

178
References

Department of Structural Engineering, Norwegian University of Science and


Technology (NTNU).

Wang K., Ge Z., Grove J., Ruiz J. M. and Rasmussen R. (2006), Developing a simple
and rapid test for monitoring the heat evolution of concrete mixtures for both
laboratory and field applications. Center for Transportation Research and Education
Iowa State University.

Washburn E. W. (1921), The dynamics of capillary flow. Physical Review Letters, Vol.
17, pp. 273–283.

Watanabe Y., Suzuki S. and Nishi S. (1969), Influence of Saccharides and Other
Organic Compounds on the Hydration of Portland Cement. Journal of Research 11, pp.
184-196.

Webb P. A. (2001), An Introduction to the Physical Characterization of Materials by


Mercury Intrusion Porosimetry with Emphasis on Reduction And Presentation of
Experimental Data. Micromeritics Instrument Corp., Norcross, Georgia.

Wiles D.B. and Young R.A. (1981), A new computer program for Rietveld analysis of
X-ray powder diffraction patterns, Journal of Applied Crystallography, Vol.14,
pp.149-151.

Williams P. J., Biernacki J. J., Bai J. and Rawn C. J. (2003), Assessment of a


synchrotron X-ray method for quantitative analysis of calcium hydroxide. Cement and
Concrete Research, Vol. 33, No. 10, pp. 1553–1559.

Winslow D. N. and Diamond S. (1970), A Mercury Porosimetry Study of the


Evolution of Porosity in Portland Cement. Journal of Materials, Vol. 5, No. 3, pp. 564
- 585.

Xu G. and Beaudoin J. J. (2000), Effect of Polycarboxylate Superplasticizer on


Contribution of Interfacial Transition Zone to Electrical Conductivity of Portland
Cement Mortars. ACI Materials Journal, Vol. 97, No. 4, pp. 418-424.

Yilmaz V. T. and Glasser F. P. (1989), Influence of sulphonated melamine


formaldehyde superplasticizer on cement hydration and microstructure. Advances in
Cement Research, Vol. 2, No. 7, pp. 111–119.

Yoshioka K., Tazawa E., Kawai K., Enohata T. (2002), Adsorption characteristics of
water-reducers on cement component minerals, Cement and Concrete Research, Vol.
32, No. 10, pp. 1507-1513.

Yousuf M., Mollah A., Palta P., Hess T. R., Vempati R. K. and Cocke D. L. (1995),
Chemical and physical effects of sodium lignosulfonate superplasticizer on the
hydration of Portland cement and solidification/stabilization consequences. Cement
and Concrete Research, Vol. 25, No. 3, pp. 671-682.

Young J. F. (1972), A review of mechanisms of set-retardation in Portland cement


pastes containing organic admixtures. Cement and Concrete Research, Vol. 2, pp. 415-

179
References

433.

Zhang L. and Glasser F. P., (2000). Critical examination of drying damage to cement
pastes. Advances in Cement Research, Vol. 12, No. 22, pp. 79-88.

Zhang W., Wang H. and Ye J. (2006), Effect of polycarboxylate-type superplasticizer


on microstructure of calcium silicate hydrates. Journal of the Chinese Ceramic
Society, Vol. 34, No. 5, pp. 546-550. (In Chinese)

Zhor J. (2006), SP-239—33: Effect of Functional Groups on the Performance of


Lignosulfonates in Cement-Water Systems. 8th CANMET/ACI International
Conference on Superplasticizers and Other Chemical Admixtures in Concrete (Ed.:
V.M. Malhotra), pp. 507-523.

Zhor J. and Bremner T. W. (1999), Role of Lignosulfonates in High Performance


Concrete. Proceedings of the International Symposium on The Role of Admixtures in
High Performance Concrete, Edited by J. G. Cabrera and R. Rivera-Villarreal,
Monterrey, Mexico, pp. 143-165.

Zhou Q. and Glasser F. P. (2001), Thermal stability and decomposition mechanisms of


ettringite at <120°C. Cement and Concrete Research, Vol. 31, No. 9, pp. 1333-1339.

180
Appendix

Table A-1 Intensity ratios of CH, C3S to anatase in pastes from XRD (w/c = 0.34)

I/Io CTR 2h 4h 8h 12h 1d 3d 7d 28d 91d


C3S 29.4° 0.75 0.71 0.48 0.39 0.25 0.19 0.17 0.15 0.15
C3S 51.7° 0.38 0.32 0.21 0.16 0.09 0.05 0.04 0.03 0.03
I/Io PCE 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.95 0.86 0.74 0.52 0.32 0.21 0.17 0.16 0.15
C3S 51.7° 0.48 0.39 0.34 0.24 0.12 0.06 0.05 0.04 0.03
I/Io SNF 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.86 0.70 0.58 0.43 0.28 0.21 0.18 0.17 0.14
C3S 51.7° 0.39 0.34 0.25 0.18 0.11 0.06 0.04 0.04 0.00
I/Io PLS 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.86 0.84 0.79 0.58 0.31 0.22 0.15 0.14 0.14
C3S 51.7° 0.39 0.39 0.38 0.26 0.10 0.05 0.04 0.03 0.00
I/Io UNA 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.85 0.84 0.81 0.66 0.35 0.19 0.18 0.17 0.16
C3S 51.7° 0.40 0.40 0.39 0.32 0.15 0.05 0.04 0.04 0.00

181
Appendix

Table A-2 Intensity ratios of CH, C3S to anatase in pastes from XRD (w/c = 0.40)

I/Io CTR 2h 4h 8h 12h 1d 3d 7d 28d 91d


C3S 29.4° 0.93 0.88 0.69 0.60 0.39 0.27 0.19 0.18 0.14
C3S 51.7° 0.48 0.46 0.33 0.22 0.15 0.07 0.04 0.02 0.00
I/Io PCE 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.95 0.88 0.85 0.72 0.38 0.22 0.16 0.15 0.14
C3S 51.7° 0.46 0.45 0.41 0.26 0.13 0.06 0.00 0.00 0.00
I/Io SNF 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.92 0.89 0.72 0.52 0.35 0.24 0.20 0.16 0.14
C3S 51.7° 0.46 0.44 0.36 0.21 0.13 0.06 0.04 0.00 0.00
I/Io PLS 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.88 0.86 0.84 0.62 0.44 0.23 0.17 0.15 0.14
C3S 51.7° 0.44 0.43 0.39 0.23 0.18 0.07 0.04 0.00 0.00
I/Io UNA 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.93 0.86 0.85 0.80 0.49 0.25 0.19 0.18 0.13
C3S 51.7° 0.44 0.42 0.41 0.42 0.21 0.07 0.04 0.00 0.00
I/Io BCS 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.86 0.81 0.79 0.77 0.52 0.24 0.18 0.16 0.13
C3S 51.7° 0.42 0.39 0.38 0.41 0.25 0.08 0.00 0.00 0.00
I/Io BCA 2h 4h 8h 12h 1d 3d 7d 28d 91d
C3S 29.4° 0.92 0.86 0.83 0.80 0.57 0.26 0.18 0.17 0.11
C3S 51.7° 0.42 0.39 0.39 0.38 0.26 0.07 0.03 0.00 0.00

Table A-3 Non-evaporable water content in pastes from furnace burning (w/c = 0.34)

2h 4h 8h 12h 1d 3d 7d 28d 91d


CTR 2.18 3.54 7.28 8.36 11.58 13.75 14.43 15.23 17.54
PCE 1.44 1.66 3.50 6.53 10.35 13.40 14.21 14.85 15.14
SNF 2.05 2.97 7.62 7.80 10.92 13.48 14.19 15.35 15.82
PLS 1.30 1.45 2.48 6.07 11.75 13.55 14.75 15.67 17.68
UNA 1.75 1.83 2.19 4.06 9.86 13.34 14.70 15.52 17.25

182
Appendix

Table A-4 Non-evaporable water content in pastes from furnace burning (w/c = 0.40)

2h 4h 8h 12h 1d 3d 7d 28d 91d


CTR 1.96 3.89 6.20 8.42 9.14 12.07 14.22 15.56 16.71
PCE 1.28 1.66 3.59 7.35 10.06 13.09 13.82 16.15 16.67
SNF 1.31 1.81 3.06 6.31 9.48 12.26 14.23 15.99 16.75
PLS 1.57 2.82 3.87 7.77 8.95 11.88 13.81 15.88 16.92
UNA 1.82 1.91 3.45 6.34 7.62 12.85 14.79 15.21 16.84
BCS 1.72 1.97 2.90 3.05 6.37 12.40 15.46 16.17 17.62
BCA 1.79 1.87 2.35 2.45 6.54 12.13 14.73 16.51 17.94

Table A-5 Calcium hydroxide content in pastes from TG (w/c = 0.34)

CH, % 2h 4h 8h 12h 1d 3d 7d 28d 91d


CTR 2.36 5.28 7.82 11.45 12.20 14.49 15.21 15.97
PCE 5.59 9.94 12.31 14.69 14.98 16.29
SNF 4.37 7.21 10.96 13.58 14.08 14.74 16.70
PLS 2.52 10.71 12.67 13.41 14.15 15.51
UNA 10.44 12.92 13.20 14.01 14.65

Table A-6 Non-evaporable water content in pastes from TG (w/c = 0.34)


NEW, % 2h 4h 8h 12h 1d 3d 7d 28d 91d
CTR 3.84 5.88 8.78 10.75 14.01 16.70 17.78 18.82 19.45
PCE 3.58 4.08 6.82 9.49 13.19 17.00 17.22 18.23 18.53
SNF 4.23 6.27 10.19 10.56 14.59 16.42 16.93 18.75 20.15
PLS 3.38 4.63 6.44 10.03 13.07 16.85 17.91 18.57 19.70
UNA 4.09 5.23 6.04 6.47 11.89 16.97 17.83 18.76 19.67

Table A-7 Calcium hydroxide content in pastes from TG (w/c = 0.40)

CH, % 2h 4h 8h 12h 1d 3d 7d 28d 91d


CTR 1.78 5.09 8.12 10.70 12.23 13.49 16.69 17.39
PCE 4.75 10.52 12.95 15.54 16.18 16.56
SNF 3.31 7.10 10.96 13.66 14.45 14.67 16.49
PLS 8.83 14.12 15.54 15.84 16.85
UNA 6.80 12.13 15.26 16.18 17.20
BCS 5.23 13.75 14.57 17.21 17.32
BCA 5.75 12.49 13.51 15.49 17.16

183
Appendix

Table A-8 Non-evaporable water content in pastes from TG (w/c = 0.40)


NEW % 2h 4h 8h 12h 1d 3d 7d 28d 91d
CTR 6.04 6.25 9.87 12.99 13.33 18.25 18.44 20.22 22.39
PCE 3.85 4.40 5.76 9.87 13.93 16.87 18.39 19.36 21.76
SNF 6.12 6.34 6.69 11.20 13.03 17.84 18.22 19.76 20.98
PLS 5.13 5.15 5.81 6.65 10.54 17.24 18.77 19.94 21.20
UNA 5.18 6.13 6.18 6.86 12.77 15.73 18.28 21.46 21.76
BCS 4.77 5.57 5.80 7.24 9.02 17.11 18.51 21.34 22.48
BCA 4.50 4.82 5.46 6.21 10.38 16.39 18.54 20.47 22.12

Table A-9 Yield stress, plastic viscosity and flow values of mortars with time (w/c =
0.34)
Rheological Parameters Flow Value
Yield Stress Plastic Viscosity 10 drops 25 drops
PCE τ0, Pa µ, Pa.s % %
10 min 70.5(1) ± 5.0(2) 7.0 ± 2.6 n.m.(3) Overflow
30 min 178 6.2 n.m. 120
60 min 264 7.2 n.m. 106
SNF τ0, Pa µ, Pa.s % %
10 min 68.5 ± 0.7 8.4 ± 0.5 n.m. Overflow
30 min 255 7.8 n.m. 118
60 min 470 6.3 n.m. 100
PLS τ0, Pa µ, Pa.s % %
10 min 77.0 ± 18.4 11.9 ± 3.2 n.m. Overflow
30 min 83 9.3 140 Overflow
60 min 147 9.5 102 132
UNA τ0, Pa µ, Pa.s % %
10 min 70.3 ± 17.0 11.1 ± 1.8 n.m. Overflow
30 min 79 9.3 140 Overflow
60 min 108 9.8 112 144
(1) (2) (3)
average value of the three data points standard deviation n.m. = not measured

184
Appendix

Table A-10 Yield stress, plastic viscosity and flow value of mortars with time (w/c =
0.40)

Rheological Parameters Flow Value


Yield Stress Plastic Viscosity 10 drops 25 drops
PCE τ0, Pa µ, Pa.s % %
(1) (2)
10 min 77.3 ± 6.1 4.1 ± 0.1 120 n.m.(3)
30 min 100 3.6 118 140
60 min 194 3.8 114 138
SNF τ0, Pa µ, Pa.s % %
10 min 73.7 ± 5.8 3.5 ± 0.2 116 n.m.
30 min 261 3.7 98 134
60 min 380 4.3 86 120
PLS τ0, Pa µ, Pa.s % %
10 min 76.0 ± 8.7 3.3 ± 0.2 144 n.m.
30 min 94 3 114 n.m.
60 min 162 3 112 142
UNA τ0, Pa µ, Pa.s % %
10 min 78.7 ± 6.8 4.7 ± 0.6 140 n.m.
30 min 97 3.5 122 146
60 min 148 3.5 110 140
BCS τ0, Pa µ, Pa.s % %
10 min 126 5.7 98 124
30 min 390 7.1 86 96
60 min n.m. n.m. n.m. n.m.
BCA τ0, Pa µ, Pa.s % %
10 min 75 ± 6.2 3.7 ± 0.2 122 n.m.
30 min 232 3.3 102 140
60 min 467 2.6 80 120
(1) (2) (3)
average value of the three data points standard deviation n.m. = not measured

185
Appendix

Table A-11 Capillary pore size distribution and total porosity of pastes (w/c = 0.34)

PCE SNF
Pore Volume,
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
mL/g
Gel
0.006 0.010 0.011 0.013 0.016 0.014 0.016 0.018 0.011 0.011
(3.8~10nm)
Medium
Capillary 0.047 0.038 0.033 0.027 0.022 0.047 0.045 0.036 0.027 0.025
(10~50nm)
Large
Capillary 0.095 0.084 0.075 0.071 0.068 0.146 0.104 0.093 0.090 0.092
(50~10µm)
Total Porosity
34.8 29.9 25.0 19.8 19.4 35.6 29.1 26.4 22.8 22.6
(%)
PLS UNA
Pore Volume,
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
mL/g
Gel
0.015 0.013 0.017 0.016 0.013 0.013 0.012 0.017 0.016 0.014
(3.8~10nm)
Medium
Capillary 0.048 0.043 0.057 0.049 0.035 0.063 0.028 0.061 0.054 0.049
(10~50nm)
Large
Capillary 0.129 0.110 0.074 0.072 0.050 0.158 0.099 0.072 0.076 0.066
(50~10µm)
Total Porosity
36.8 28.9 26.8 25.5 22.7 40.3 31.7 27.4 24.7 24.2
(%)

186
Appendix

Table A-12 Capillary pore size distribution and total porosity of pastes (w/c = 0.40)

PCE SNF
Pore Volume,
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
mL/g
Gel (3.8~10nm) 0.0097 0.0233 0.0249 0.0156 0.0172 0.0132 0.0146 0.026 0.0201 0.0182
Medium
Capillary 0.059 0.0638 0.0495 0.0251 0.0323 0.0431 0.0476 0.0588 0.0545 0.0498
(10~50nm)
Large Capillary
0.2118 0.1242 0.111 0.0382 0.0531 0.2212 0.1391 0.1113 0.0923 0.0892
(50~10000nm)
Total Porosity
43.6 36.1 32.4 29.1 26.8 44.9 36.7 33.3 29.9 28.4
(%)
PLS UNA
Pore Volume,
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
mL/g
Gel (3.8~10nm) 0.0167 0.0153 0.0221 0.0181 0.0165 0.0154 0.0169 0.0185 0.0194 0.0157
Medium
Capillary 0.0166 0.0526 0.0622 0.0555 0.0492 0.0492 0.0584 0.0641 0.0601 0.0535
(10~50nm)
Large Capillary
0.2214 0.202 0.1579 0.0982 0.0718 0.2062 0.1534 0.1411 0.1113 0.0769
(50~10000nm)
Total Porosity
45.1 38.6 33.5 29.5 27.3 43.9 39.6 34.1 31.6 25
(%)
BCS BCA
Pore Volume,
1d 3d 7d 28d 91d 1d 3d 7d 28d 91d
mL/g
Gel (3.8~10nm) 0.0075 0.0182 0.0212 0.0178 0.0181 0.0077 0.0254 0.0221 0.0271 0.0261
Medium
Capillary 0.0052 0.0388 0.0525 0.0407 0.0519 0.0103 0.0663 0.064 0.0558 0.0523
(10~50nm)
Large Capillary
0.2501 0.1549 0.1177 0.0791 0.0794 0.2382 0.1314 0.109 0.0948 0.0741
(50~10000nm)
Total Porosity
41.6 37.7 32.9 29.2 24.1 42.7 39.9 34.3 29.6 25.8
(%)

187
Appendix

Table A-13 Average & standard deviation of mortar compressive strengths (w/c = 0.34)

fcu* PCE SNF PLS UNA


1d 35.4 37.5 24.8 17.4
3d 53.2 54.1 49.7 45.7
7d 61.9 66.0 61.6 56.4
28d 69.7 72.6 70.7 64.8
91d 73.1 78.6 74.8 70.9
STD**
1d 0.0 0.4 0.5 0.6
3d 0.4 1.8 0.9 1.3
7d 0.3 0.3 1.0 1.4
28d 2.4 0.9 1.2 2.5
91d 3.8 3.3 4.6 1.5
* = average of compressive strength of three 50mm mortar cubes
** = standard deviation of the three fcu of the mortar cubes

Table A-14 Average & standard deviation of mortar compressive strengths (w/c = 0.40)

fcu* PCE SNF PLS UNA BCS BCA


1d 24.0 28.7 21.3 16.9 0.0 8.9
3d 45.3 47.0 43.0 40.2 40.5 40.0
7d 52.9 56.5 51.1 52.0 54.8 51.7
28d 70.8 67.9 62.9 66.7 73.0 67.0
91d 73.3 75.5 68.9 69.8 75.7 72.7
STD**
1d 0.3 1.1 0.8 0.2 0.0 0.6
3d 0.6 2.1 0.8 1.4 1.0 0.6
7d 1.1 1.3 2.0 0.8 0.5 1.3
28d 1.7 2.9 2.7 1.1 1.1 1.8
91d 0.6 2.2 1.6 2.3 1.3 1.0
* = average of compressive strength of three 50mm mortar cubes
** = standard deviation of the three fcu of the mortar cubes

188
Appendix

Fig. A-1 TG curves of the control paste (w/c = 0.34)

Fig. A-2 TG curves of PCE paste (w/c = 0.34)

189
Appendix

Fig. A-3 TG curves of SNF paste (w/c = 0.34)

Fig. A-4 TG curves of PLS paste (w/c = 0.34)

190
Appendix

Fig. A-5 TG curves of UNA paste (w/c = 0.34)

Fig. A-6 TG curves of the control paste (w/c = 0.40)

191
Appendix

Fig. A-7 TG curves of PCE paste (w/c = 0.40)

Fig. A-8 TG curves of SNF paste (w/c = 0.40)

192
Appendix

Fig. A-9 TG curves of PLS paste (w/c = 0.40)

Fig. A-10 TG curves of UNA paste (w/c = 0.40)

193
Appendix

Fig. A-11 TG curves of BCS paste (w/c = 0.40)

Fig. A-12 TG curves of BCA paste (w/c = 0.40)

194
Appendix

0.25

0.2
Cumulative Intrusion, mL/g

0.15 I-1d
I-3d

I-7d
0.1 I-28d

I-91d

0.05

0
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-13 MIP curves for PCE paste (w/c = 0.34)

0.25

0.20 II-1d
Cumulative Intrusion, mL/g

0.15 II-3d
II-7d
II-28d
II-91d
0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-14 MIP curves for SNF paste (w/c = 0.34)

195
Appendix

0.25

III-1d
0.20
Cumulative Intrusion, mL/g

0.15 III-3d
III-7d
III-28d

0.10 III-91d

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-15 MIP curves for PLS paste (w/c = 0.34)


0.25

IV-1d
0.20
Cumulative Intrusion, mL/g

IV-3d

0.15
IV-28d
IV-7d

IV-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-16 MIP curves for UNA paste (w/c = 0.34)

196
Appendix

0.30

A-1d
0.25
Cumulative Intrusion, mL/g

0.20
A-3d

A-7d
A-28d
0.15
A-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-17 MIP curves for PCE paste (w/c = 0.40)

0.30

B-1d
0.25
Cumulative Intrusion, mL/g

B-3d
0.20
B-7d

B-28d
0.15
B-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-18 MIP curves for SNF paste (w/c = 0.40)

197
Appendix

0.30
C-1d

0.25 C-3d

C-7d
Cumulative Intrusion, mL/g

0.20

C-28d
0.15 C-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-19 MIP curves for PLS paste (w/c = 0.40)

0.30
D-1d

0.25 D-3d
Cumulative Intrusion, mL/g

0.20 D-7d

D-28d
0.15

D-91d
0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-20 MIP curves for UNA paste (w/c = 0.40)

198
Appendix

0.30

E-1d
Cumulative Intrusion, mL/g 0.25

E-3d
0.20

E-7d

0.15 E-28d

E-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-21 MIP curves for BCS paste (w/c = 0.40)

0.30

F-1d
0.25

F-3d
Cumulative Intrusion, mL/g

0.20
F-7d
F-28d
0.15 F-91d

0.10

0.05

0.00
0.001 0.01 0.1 1 10 100 1000
Mean Diameter, µm

Fig. A-22 MIP curves for BCA paste (w/c = 0.40)

199

You might also like