THESIS - C. JAYASREE - Jayasree Chakkamalayath

Download as pdf or txt
Download as pdf or txt
You are on page 1of 282

STUDY OF CEMENT-SUPERPLASTICIZER INTERACTION AND

ITS IMPLICATIONS FOR CONCRETE PERFORMANCE

A THESIS

submitted by

C. JAYASREE

for the award of the degree

of

DOCTOR OF PHILOSOPHY

BUILDING TECHNOLOGY AND CONSTRUCTION MANAGEMENT DIVISION


DEPARTMENT OF CIVIL ENGINEERING
INDIAN INSTITUTE OF TECHNOLOGY MADRAS
CHENNAI 600 036

DECEMBER 2008
THESIS CERTIFICATE

This is to certify that the thesis entitled “STUDY OF CEMENT-

SUPERPLASTICIZER INTERACTION AND ITS IMPLICATIONS FOR

CONCRETE PERFORMANCE” submitted by C. Jayasree to the Indian Institute

of Technology Madras for the award of the degree of Doctor of Philosophy is a

bonafide record of research work carried out by her under my supervision. The

contents of this thesis, in full or in parts, have not been submitted to any other

Institute or University for the award of any degree or diploma.

Chennai 600 036 Prof. Ravindra Gettu


Date: (Research Guide)
Professor
Department of Civil Engineering
Indian Institute of Technology Madras
ACKNOWLEDGEMENTS

I am deeply indebted to my research guide Prof. Ravindra Gettu, Department of Civil


Engineering for his timely guidance, integral view of research, uncompromising
standards for quality and inestimable help right from the initial stage of the research
having enabled me to complete the thesis within the stipulated time. His calmness,
patience, positive and constructive criticisms, willingness to spend long hours on
discussions and his commitment to the job at hand are qualities to emulate. He always
made me feel welcome in his office with a smiling face, no matter how busy he was,
or how silly my question. The experience has been enriching and enlightening and I
thank God almighty for giving me an opportunity to work as a research scholar with
him at Indian Institute of Technology, Madras.

I also wish to place on record my sincere gratitude to Prof. K. Ramamurthy,


Department of Civil Engineering, who had guided me during the beginning of my
research programme and always kept an eye on the progress of my work by giving
suggestions as a doctoral committee member.

I am extremely thankful to Dr. J. Muralikrishnan, Assistant Professor, Department of


Civil engineering, for the great interest he took to help me with valuable suggestions
at certain stages of the thesis work. My sincere thanks are also due to Dr. Manu
Santhanam, Assistant Professor, Department of Civil engineering for the creative
discussions and suggestions at different stages of the thesis work.

I express my sincere thanks to the present Head of Civil Engineering Department,


Prof. K. Rajagopal and the former head, Prof. S. Mohan for all the administrative
support extended to me for carrying out this work. I would like to thank my doctoral
committee members Dr. Ligy Philip, Dr. G. Sekhar and Dr. Anuradha Banerjee for
the suggestions at different stages of my work.

I take this opportunity to extend my sincere gratitude to all faculty members of the
Building Technology and Construction Management Division, Prof. A.
Ramachandriah, Prof. M. S. Mathews, Prof. K. N. Sathyanarayana, Prof. Koshy
Varghese, Dr. K. Ananthanarayanan and Dr. Ashwin Mahalingam for all the
encouragement and support.

i
I also wish to thank Prof. C. Vijayan, Department of Physics, Dr. Abhijit Deshpande,
Associate Professor, Department of Chemical Engineering and Prof. Kesavan Nair,
Department of Metallurgical Engineering for supporting me with their valuable time
for discussions and permitting me to use the laboratory facilities.

I am pleased to record my gratitude to the Management of Amrita Vishwa Vidya


Peetham University, Coimbatore for sponsoring me for the Ph. D programme under
Q.I.P. scheme. My sincere thanks to Prof. C. R. Ramakrishnan, Prof. T.R.
Padmanabhan, Dr. K.M. Mini, Mrs. Sunitha and other colleagues in the Department
of Civil Engineering, Amrita School of Engineering for extending a helping hand at
crucial times.

The staff of concrete lab will always have a special place in my heart for all the efforts
they have put in towards the successful completion of my experimental work. Ms. A.
Malarvizhi, Messers M. Soundarapandiyan, Murthy, Krishnan, Subramaniam,
Dhanasekaran, Watson all have contributed their might for the smooth conducting of
the experimental work. I am also thankful to all the staff members of the Civil
Engineering Department Office, Stores section and DCF for their excellent
cooperation.

I also wish to thank Palani, Premraj, Premkumar, Vijayakumar, Krishna and Prasad
who admirably and unstintingly helped me physically through all the laboratory work.

I would like to thank the chemical admixture companies, BASF Construction


Chemicals, Sika, Cera-Chem and, Burzin and Leons for giving the product
information that was needed as well as the products that they donated for use in this
thesis. I also wish to thank Grassim Industries for supplying cement for the study.

I have been blessed with a friendly and cheerful group of fellow research scholars-
Cindrela, Glory, Geetha, Indu, Rakesh, Ganesh, Prakash, Anand, Senthil Kumar,
Rajesh, Elson, Paul, Venkatesan, Shajatnan, Rajasekar, Boeing, Liju, Arun, Dr.
Manikandan, Dr. Uma, Dr. Ramesh Babu, Dr. Kanagasabhapathi, Dr. Harikrishnan,
Dr. Nambiar, Dr. Anil kumar and many others. I express my gratitude to all of them
for making my life at IIT pleasant and memorable.

ii
I would like to acknowledge my indebtedness to several of the individuals who have
been instrumental in this research- Simi, Vinoj, Anbudayanidhi and Diwakar had
helped me with suggestions. Many thanks to Jennifer, Sansu, Shanthi and Mohan for
helping me in the laboratory while conducting experiments. I am grateful to
Ms. Nigiya, Dr. Sindhu and their families for their affection and help during my stay
at IIT.

I am deeply indebted to my in-laws who have encouraged me in fulfilling my dream


of doing research at IIT. I am grateful to my husband, Girish who in his own special
way has supported me and my work in the last three years. To my daughter,
Sreelakshmi, for her patience, understanding and interest in my studies. My heartfelt
thanks to Ms. Subhadra, who was with me all the three years for making my life at IIT
pleasant and smooth. Finally, I feel a deep sense of gratitude to my mother, sister and
other family members who have supported me in all possible ways and have made me
what I am today.

C. JAYASREE

iii
ABSTRACT

Chemical admixtures, particularly superplasticizers have become indispensable

because of their beneficial effects in fresh and hardened concrete. However, the

compatibility between the cement and admixture, which is influenced by the

characteristics of cement paste, and the type and dosage of admixtures, could affect

the benefits of incorporating a superplasticizer. It is, therefore, essential to understand

the mechanisms that influence the cement-superplasticizer interaction for selecting an

appropriate type of admixture at an optimum dosage. The current study addresses

these and other specific issues pertaining to the performance of concrete incorporating

chemical and mineral admixtures, especially superplasticizers.

A suitable mixing method for the preparation of paste is first selected so that the paste

prepared will be representative of that of concrete. In order to evaluate the influence

of superplasticizers and cement characteristics, studies are conducted with four

families of superplasticizers and four types of cements on fluidity, loss of fluidity and

setting behaviour. The relative fluidity of different superplasticized pastes are studied

with the Marsh cone, mini-slump and viscometric tests for evaluating the best

combination. The Marsh cone flow time, yield stress and plastic viscosity decreases

whereas the mini-slump spread increases with an increase in the dosage of the

superplasticizer upto the saturation dosage, after which they remain constant. An

experimental and analytical validation of the Marsh cone test is done using

rheological models to understand its effectiveness in capturing the relative fluidity.

The correlation between the flow behaviour of paste with normal concrete and self

compacting concrete shows that the saturation dosages obtained from paste and

mortar are comparable whereas a dosage slightly higher than the saturation dosage of

iv
the paste is required for adequate workability in concrete. The setting behaviour is

studied through Vicat penetration and electrical conductivity methods. Both tests

confirm that superplasticizers retard the setting of cement paste; however, the extent

of retardation varies with the type of superplasticizer.

Another aspect studied in this work is the influence of the superplasticizer on the

hydration processes and on the development of microstructure of hydrated cement

paste. With this aim, cement pastes are analyzed using several characterization

techniques (X-ray diffraction, scanning electron microscopy, thermal analysis and Si

nuclear magnetic resonance). The test results clearly indicate the influence of

superplasticizers in delaying the hydration. The superplasticizer delays the formation

of ettringite as well as the silicate polymerization; however, it produces a dense and

uniform microstructure, enhancing the strength development.

The absorbance study using a UV spectrometer gives insight into the influence of

cement and superplasticizer characteristics on the absorbance of the superplasticizers

and its variation with time. The study of the influence of other admixtures on the

cement-superplasticizer interaction shows that the addition of metakaolin and

viscosity modifying admixtures increases the saturation dosage of the superplasticizer

and the viscosity of the paste.

Finally, based on the test results, a simple methodology is recommended to select a

compatible cement-superplasticizer combination is recommended. The use of Marsh

cone and Vicat penetration tests in the paste phase, and final trials on concrete for

strength and cost evaluation are suggested as part of this methodology.

Keywords: Superplasticizer; Marsh cone; mini-slump; rheology; loss of


fluidity; compatibility.

v
TABLE OF CONTENTS

Title Page No.

ACKNOWLEDGEMENTS ....................................................................................... i
ABSTRACT ............................................................................................................... iv
LIST OF TABLES ................................................................................................... xii
LIST OF FIGURES ................................................................................................ xiv

CHAPTER 1 INTRODUCTION

1.1 General ...........................................................................................................1


1.2 Objectives and Scope .....................................................................................3
1.3 Structure of the Thesis....................................................................................4

CHAPTER 2 STATE-OF THE-ART REPORT ON CEMENT-


SUPERPLASTICIZER INTERACTION

2.1 General ...........................................................................................................7


2.2 Cement Chemistry ..........................................................................................8
2.2.1 Raw Materials Used in Cement Manufacture ................................................8
2.2.2 Cement Composition ......................................................................................9
2.2.3 Hydration of Cement ....................................................................................13
2.2.3.1 Hydration of Cement compounds ................................................................13
2.2.3.2 Hydration of the Portland Cement Grain .....................................................15
2.3 Water Reducers and Superplasticizers .........................................................18
2.3.1 Characteristics and Chemistry of Superplasticizers .....................................19
2.3.2 Mechanisms of Action of Superplasticizers .................................................21
2.3.3 Factors Affecting the Cement-Superplasticizer Interaction ........................28
2.3.3.1 Effect of Chemical Structure of Superplasticizer ........................................28
2.3.3.2 Effect of Calcium Sulphate .........................................................................33
2.3.3.3 Effect of Soluble Alkalis ..............................................................................35
2.3.3.4 Effect of C3A ................................................................................................36
2.3.3.5 Other Factors Affecting the Cement-Superplasticizer Interaction ...............38
2.3.4 Effect of Superplasticizer on the Morphology of Hardened Cement Paste .39
2.3.5 Effect of Superplasticizer on Properties of Concrete .................................41
2.3.6 Admixture–Admixture Interaction ...............................................................43
vi
Table of Contents (Contd.,) Page No.

2.3.6.1 Cement superplasticizer Interaction in the Presence of Mineral


Admixtures ...................................................................................................43
2.3.6.2 Influence of other Chemical Admixtures on Cement-Superplasticizer
Interaction.....................................................................................................44

2.4 Need for the Present Study ...........................................................................44

CHAPTER 3 STUDY OF THE FLOW BEHAVIOUR OF


SUPERPLASTICIZED CEMENT PASTE

3.1 General ................................................................................................ 47


3.2 Materials Used in the Study .........................................................................48
3.2.1 Cement .........................................................................................................48
3.2.1.1 Determination of the Water Demand for Standard Consistency
of Cement ................................................................................................... 50
3.2.1.2 Determination of Initial and Final Setting Times.........................................50
3.2.1.3 Determination of the Fineness of Cement ....................................................51
3.2.2 Superplasticizers ..........................................................................................51
3.3 Test Methods Used for Studying the Flow Behaviour of the
Cement Paste ...............................................................................................52
3.3.1 Marsh Cone Test .........................................................................................53
3.3.2 Mini-slump Test ...........................................................................................56
3.3.3 Viscometer Test............................................................................................57
3.4 Selection of Mixing Method ........................................................................62
3.4.1 Materials Used in the Selection of Mixing Method .....................................63
3.4.2 Experimental Details ...................................................................................63
3.4.3 Ball Milling ..................................................................................................64
3.4.4 Results and Discussions ..............................................................................66
3.4.5 Conclusions ..................................................................................................69
3.5 Influence of Type and Dosage of Superplasticizer on the
Flow Behaviour of Cement Paste .................................................................70
3.5.1 Testing Details..............................................................................................70
3.5.2 Results and discussions ................................................................................71
3.5.3 Correlations between the Results from Empirical Tests and
the Viscometer..............................................................................................78
3.5.4 Influence of Type and Dosage of the Superplasticizer on the
Non-Newtonian Characteristics of Cement Paste ........................................82

vii
Table of Contents (Contd.,) Page No.

3.5.5 Change of Flow Behaviour with Time .........................................................82


3.5.6 Summary ......................................................................................................85

CHAPTER 4 CORRELATIONS BETWEEN THE FRESH AND HARDENED


PROPERTIES OF SUPERPLASTICIZED PASTE, MORTAR
AND CONCRETE

4.1 General .........................................................................................................87


4.2 Experimental Details ....................................................................................88
4.2.1 Materials ......................................................................................................88
4.2.2 Test Procedures ............................................................................................90
4.2.2.1 Tests on Paste ...............................................................................................90
4.2.2.2 Tests on Mortar ............................................................................................90
4.2.2.3 Tests on Concrete .........................................................................................91
4.2.3 Mixing Methods for Paste, Mortar and Concrete .........................................92
4.3 Results and Discussions ...............................................................................93
4.3.1 Comparison of the Flow Behaviour of Paste, Mortar and Concrete ............93
4.3.1.1 Flow Behaviour of Paste ..............................................................................93
4.3.1.2 Flow Behaviour of Mortar............................................................................94
4.3.1.3 Flow Behaviour of Concrete ........................................................................96
4.3.2 Correlation between the Fluidity of Paste, Mortar and Concrete .................98
4.4 Loss of Fluidity ..........................................................................................101
4.4.1 Loss of Fluidity of Paste.............................................................................101
4.4.2 Loss of Fluidity of Mortar ..........................................................................103
4.4.3 Loss of Fluidity of Concrete.......................................................................105
4.5 Effect of Superplasticizer on Setting of Cement Paste and Concrete ........106
4.6 Analysis of Setting Behaviour of Superplasticized Cement Paste Using
Electrical Resistivity Method ....................................................................109
4.6.1 Background ................................................................................................109
4.6.2 Experimental Procedure .............................................................................110
4.6.3 Results and Discussions .............................................................................111
4.7 Effect of Superplasticizer on Compressive Strength of Concrete ..............116
4.8 Summary ....................................................................................................117

viii
Table of Contents (Contd.,) Page No.

CHAPTER 5 EFFECT OF SUPERPLASTICIZER ON PROPERTIESS OF


HARDENED CEMENT PASTE

5.1 General .......................................................................................................120


5.2 X-ray Diffraction (XRD) Study .................................................................121
5.2.1 Background ................................................................................................121
5.2.2 Materials Used and Sample Preparation ....................................................122
5.2.3 Experimental procedure .............................................................................122
5.2.4 Results and discussions ..............................................................................123
5.3 Scanning Electron Microscopy (SEM) ......................................................128
5.3.1 Background .................................................................................................128
5.3.2 Sample Preparation .....................................................................................129
5.3.3 Comparison of the Microstructure of Superplasticized Paste and
Reference paste .........................................................................................130
5.4 Thermal (DTA/TG/ DTG) Analysis ...........................................................135
5.4.1 Background ................................................................................................135
5.4.2 Results and Discussions .............................................................................135
5.5 Nuclear Magnetic Resonance Spectroscopy (Si NMR) .............................142
5.5.1 Background ................................................................................................142
5.5.2 Sample Preparation ....................................................................................143
5.5.3 Results and Discussions .............................................................................143
5.6 Summary ....................................................................................................146

CHAPTER 6 CEMENT-SUPERPLASTICIZER COMPATIBILITY

6.1 General .......................................................................................................148


6.2 Materials Selection and Characterisation ...................................................149
6.2.1 Types of cements ........................................................................................149
6.2.2 Superplasticizers.........................................................................................150
6.3 Experimental Procedure .............................................................................151
6.4 Results and discussions ..............................................................................151
6.4.1 Influence of Cement Composition on the Flow Behaviour of
Superplasticized Pastes ..............................................................................151
6.4.2 Influence of Cement Composition on the Setting of
Superplasticized Pastes ..............................................................................160

ix
Table of Contents (Contd.,) Page No.

6.5 Study of the Adsorption of superplasticizers .............................................162


6.5.1 Background ................................................................................................162
6.5.2 Materials Used............................................................................................163
6.5.3 Experimental Procedure .............................................................................163
6.5.4 Results and Discussions of Absorbance Tests ...........................................165
6.5.4.1 Variation in Superplasticizer Absorbance with type of cement .................168
6.6 Identifying Compatible Cement-Superplasticizer Combinations ..............168
6.7 Summary ....................................................................................................176

CHAPTER 7 FLOW BEHAVIOUR OF SUPERPLASTICIZED PASTES


INCORPORATING OTHER ADMIXTURES

7.1 General .......................................................................................................177


7.2 Materials Used in the Study .......................................................................178
7.3 Flow Behaviour of Pastes with Metakaolin ...............................................179
7.3.1 Effect of Metakaolin in Paste Incorporating PCE based Superplasticizer .180
7.3.2 Effect of Metakaolin in Paste Incorporating SNF based Superplasticizer .183
7.4 Influence of VMA on the Flow Behaviour of Superplasticized Paste .......185
7.4.1 Characteristics of VMA .............................................................................185
7.4.2 Experimental Details ..................................................................................187
7.4.2.1 Study of the Interaction between Superplasticizer and VMA ....................187
7.4.2.2 Effect of Time on the Flow Behaviour of Paste
Incorporating Superplasticizer and VMA .................................................191
7.5 Comparison of the Flow Behaviour of Paste and Self
Compacting Concrete .................................................................................193
7.5.1 Mix Design and Testing of SCC ................................................................193
7.6 Summary ....................................................................................................200

CHAPTER 8 CONCLUSIONS AND RECOMMENDATIONS FOR


FURTHER RESEARCH

8.1 General Conclusions ..................................................................................202


8.2 Specific Conclusions ..................................................................................204
8.2.1 Influence of Mixing Method on the Flow Behaviour of Paste ...................204
8.2.2 Influence of Superplasticizers on the Flow Behaviour of Paste.................205

x
Table of Contents (Contd.,) Page No.

8.2.3 Comparison between Marsh Cone Flow time and


Rheological Parameters ..............................................................................206
8.2.4 Effect of Type and Dosage of superplasticizer on the Retardation of
Setting.........................................................................................................207
8.2.5 Correlation between Paste, Mortar and Concrete.......................................207
8.2.6 Effect of Superplasticizer on the Hydration Process and on the
Microstructural Development ....................................................................209
8.2.7 Influence of Chemical Composition of Cement on Compatibility ............210
8.3 Guidelines for Selecting the Best Combination of Cement and
Superplasticizer ..........................................................................................211
8.4 Recommendations for Further Research ....................................................211

APPENDICES .........................................................................................................214
REFERENCES ........................................................................................................243
PUBLICATIONS BASED ON THE THESIS ......................................................263

xi
LIST OF TABLES
 

Table No. Title Page No.

2.1 Typical Portland Cement Composition Conforming


to ASTM C 150 .............................................................................................10
2.2 Requirements of Different Grades of Cement as per IS codes ......................11
2.3 Chemical Analyses of Some CEM I Portland Cements of UK......................12
2.4 Chemical Structure of Superplasticizers ........................................................20
2.5 Structural Aspects of the PCE that Affect the Fluidity of Cement Paste .......32
3.1 Chemical properties of the cement and the corresponding clinker ................49
3.2 Bogue composition of the clinker ..................................................................49
3.3 Physical properties of cement 1 ....................................................................50
3.4 Properties of superplasticizers........................................................................52
3.5 Optimum superplasticizer dosage and range of unit weight for different
mixing methods ...............................................................................................67
3.6 Rheological characteristics and Marsh cone flow times ................................77
3.7 Test data for pastes with different superplasticizers at 0 and 60-minutes......83
4.1 Summary of properties of aggregates ............................................................89
4.2 Summary of Marsh cone data and concrete cost............................................94
4.3 Loss of fluidity of mortar .............................................................................104
4.4 Initial and final setting time of paste and concrete ......................................109
4.5 Compressive strength of concrete for different combinations .....................117
5.1 Relative evolution of crystalline phases with age ........................................127
5.2 Summary of mass loss in different pastes ....................................................142
5.3 Ranges of 29 Si chemical shifts of the Qn tetrahedrons.................................143
6.1 Chemical and Bogue composition of different cements ..............................150
6.2 Physical properties of different types of cements ........................................150
6.3 Test data for pastes with different superplasticizers for cement C1 ............155
6.4 Test data for pastes with different superplasticizers for cement C2 ............156
6.5 Test data for pastes with different superplasticizers for cement C3 ............157
6.6 Test data for pastes with different superplasticizers for cement C4 ............158
6.7 Comparison of the saturation dosages at 0-minutes for different cements ..159
6.8 Summary of absorbance test results .............................................................167
6.9 Summary of Superplasticized paste combinations at 0-minutes ..................173
xii
List of Tables (Contd.,) Page No.

6.10 Summary of Superplasticized paste combinations at 60-minutes ................174


7.1 Properties of viscosity modifying agents .....................................................179
7.2 Flow characteristics of pastes with different types and dosages of VMA ...191
7.3 Influence of time on SP-VMA mix ..............................................................192
7.4 Acceptance criteria for SCC ........................................................................195
7.5 Proportion of materials used in the SCC trials .............................................198
7.6 Test results of SCC with various SP-VMA combinations ...........................199
A.1 Test results of three trials in the selection of mixing method ......................214
A.2 Comparison of superplasticizer dosages (data from Marsh cone and viscometer
test results) ....................................................................................................214
C.1 Mix details for steady shear experiments .....................................................224
C.2 Work done in different loading-unloading cycles ........................................227
C.3 Shear thinning indices for different types and dosages of superplasticizers 229
C.4 Mix details for creep and recovery and stress relaxation experiments ........230
C.5 Details of the different data acquisition frequencies ....................................230
C.6 Superposition of separate responses ............................................................237

xiii
LIST OF FIGURES

Figure Title Page No

2.1 Microstructure of hydrated portland cement ..................................................16


2.2 Physical Action of Superplasticizers ..............................................................24
2.3 Chemical action of superplasticizers ..............................................................26
3.1 Geometry and dimensions (in millimetres) of the Marsh cone ......................54
3.2 Flow time curve showing the determination of optimum Superplasticizer
dosage (according to the method of Gomes et al., 2001) ...............................55
3.3 Marsh cone test showing continuous and discontinuous flow .......................55
3.4 Mini-slump cone ............................................................................................56
3.5 Mini-slump test ..............................................................................................56
3.6 Brookfield HA DV II +Pro viscometer ..........................................................58
3.7 Details of co-axial cylinder setup ..................................................................58
3.8 Shear history of rheological testing ...............................................................59
3.9 Typical graph showing hysteresis cycles for a superplasticized
cement Paste ...................................................................................................59
3.10 Fits of experimental data with Bingham and Herschel-Bulkley Models ......62
3.11(a) Hobart mixer with B-flat beater ...................................................................64
3.11(b) Hobart mixer with D-wire whip .....................................................................64
3.12 Test set up for ball milling .............................................................................65
3.13 Typical flow time curves obtained with Marsh cone for different mixing
methods .........................................................................................................67
3.14 Typical flow curves obtained with mini-slump for different mixing
methods ..........................................................................................................68
3.15 Marsh cone flow time for different superplasticizers ....................................71
3.16 Mini-slump spread for different superplasticizers .........................................72
3.17 Correlation between Marsh cone flow time and mini-slump spread .............73
3.18 Typical graphs showing hysteresis cycles for a PCE based superplasticized
cement paste ..................................................................................................75
3.19 Comparison of data from different tests ........................................................79
3.20 Prediction of yield stress from mini-slump spread using Roussel-Coussot
model ..............................................................................................................80

xiv
List of Figures (Contd.,) Page No.

3.21 Flow time curve showing the determination of optimum superplasticizer


dosage according to Aïtcin (1998) .................................................................84
4.1 Particle size distribution of fine aggregates (0-4.75mm) ...............................89
4.2 Particle size distribution of coarse aggregates (4.75-20 mm) ........................89
4.3 Flow table test for mortar ...............................................................................90
4.4 Tests for workability of concrete ...................................................................91
4.5 Marsh cone flow time curves for superplasticized cement paste ...................93
4.6 Marsh cone flow times of cement mortar ......................................................95
4.7 Flow table spreads for superplasticized mortars ............................................95
4.8 Variation of concrete slump with dosage of superplasticizer ........................97
4.9 Variation of flow table spread with dosage of superplasticizer .....................97
4.10 Correlation between paste, mortar and concrete in terms of flow time and
slump for SNF-S1 ..........................................................................................99
4.11 Correlation between paste, mortar and concrete in terms of flow time and
slump for SNF-D2 ..........................................................................................99
4.12 Correlation between paste, mortar and concrete in terms of flow time and
slump for SMF-S1 ........................................................................................100
4.13 Correlation between paste, mortar and concrete in terms of flow time and
slump for PCE-D1 ........................................................................................100
4.14 Correlation between slump of concrete, and the yield stress and plastic
viscosity of paste ..........................................................................................101
4.15 Loss of fluidity of paste ...............................................................................102
4.16 Loss in slump of concrete with time ............................................................106
4.17 Vicat penetration measurements for paste with different superplasticizers .107
4.18 Penetration test on concrete .........................................................................108
4.19 Setting of concrete at saturation dosage of superplasticizer ........................108
4.20 Comparative study on setting behaviour of paste with Vicat apparatus and
LCR meter ....................................................................................................111
4.21 Evolution of conductivity for different superplasticizers.............................114
4.22 Electrical conductivity graphs for different pastes along with beginning and
end of setting defined by Vicat penetration measurements .........................115
5.1 X-ray diffractograms of cement paste and superplasticized cement paste ..126

xv
List of Figures (Contd.,) Page No.

5.2 BSE (250X) Micrograph of pure cement paste and superplasticized cement
paste .............................................................................................................133
5.3 BSE micrograph along with EDAX for a dark grey (C-S-H) particle in the LS
based paste ...................................................................................................134
5.4 DTA analysis of cement paste and superplasticized paste at 3 days............137
5.5 TG and DTG analysis of pastes at 3 days ....................................................139
5.6 TG and DTG analysis of pastes at 28 days ..................................................141
5.7 Details of Si NMR spectrum of different pastes for different ages .............145
5.8 Determination of area of peaks of Si NMR spectrum ..................................145
5.9 Comparison of area of the Q1 peak (A1) with reference to the area of the Q0
peak (A0) at different ages............................................................................146
6.1 Marsh cone flow times of different cement-superplasticizer combinations 154
6.2 Influence of type of cement on setting .........................................................162
6.3 Absorbance spectra for the different superplasticizers ................................165
6.4 Absorbance spectrum of the diluted aqueous solutions of superplasticized
pastes extracted after mixing ........................................................................166
6.5 Absorbance spectrum of cement C2 and LS-C1 ..........................................168
6.6 Methodology to select a compatible cement-superplasticizer combination 172
6.7 Classification of cement based on its chemical composition .......................176
7.1 Marsh cone flow time curves for pastes with metakaolin and PCE based
superplasticizer.............................................................................................180
7.2 Variation of mini-slump spread with dosage of PCE for metakaolin-PCE
combination ..................................................................................................181
7.3 Variation of yield stress with dosage of PCE for metakaolin-PCE
combination ..................................................................................................182
7.4 Variation of plastic viscosity with dosage of PCE for metakaolin-PCE
combination...................................................................................................182
7.5 Marsh cone flow time curves for pastes with metakaolin and SNF-D2 ......184
7.6 Variation of mini-slump spread for pastes with metakaolin and SNF-D2....184
7.7 Variation of yield stress for paste with metakaolin and SNF-D2 ................185
7.8 Variation of plastic viscosity for paste with metakaolin and SNF-D2.........185
7.9 Marsh cone flow time for different viscosity modifying agents ..................188
7.10 Mini slump spread for different viscosity modifying agents .......................189
xvi
List of Figures (Contd.,) Page No.
7.11 Values of yield stress and plastic viscosity for different viscosity modifying
agents ..........................................................................................................189
7.12 Determination of saturation dosage of superplasticizer at different
dosages of VMA ..........................................................................................190
7.13 Effect of time on Marsh cone flow curves of SP-VMA mix .......................192
7.14 Determination of aggregate proportions for concrete ..................................194
7.15 Slump test .....................................................................................................196
7.16 J-Ring test ....................................................................................................197
7.17 V-Funnel test ................................................................................................197
A.1 Correlation between Marsh cone flow time and rheological parameters for
different compositions ..................................................................................216
A.2 Correlation between Marsh cone flow time and mini-slump spread with
rheological parameters for different compositions ......................................216
C.1 Anton Paar Rheometer with parallel plate attachment .................................224
C.2 Loading-unloading cycles for the superplasticizer, PCE-D1 at the dosage
of 0.1% .........................................................................................................225
C.3 Loading-unloading cycles for the superplasticizer, PCE-D1 at the dosage
of 1% ............................................................................................................226
C.4 Creep and recovery test for pure cement paste ............................................232
C.5 Creep and recovery test result of cement paste and with constant
dosage of different superplasticizers ............................................................233
C.6 Variation in creep and recovery behaviour with different dosages of
PCE-D1 ........................................................................................................234
C.7 Creep test result at saturation dosage of superplasticizers ...........................235
C.8 Test for linear scaling ...................................................................................236
C.9 Stress relaxation of pure cement paste .........................................................238
C.10 Stress relaxation test results for pastes with 0.05% dosage for different
superplasticizers ............................................................................................239
C.11 Stress relaxation at different dosages of PCE-D1 ........................................240
C.12 Stress relaxation of pastes at saturation dosage of superplasticizers ...........241

xvii
CHAPTER 1

INTRODUCTION

1.1 GENERAL

Recent advances in concrete technology have generally resulted in better performance

of concrete in terms of strength, workability and durability over the past few decades.

The enhancement of fresh and hardened state properties has been achieved especially

through the incorporation of chemical and mineral admixtures. The multifunctional

benefits of chemical admixtures, such as the high range water reducers (HRWRs) or

superplasticizers, have made them essential components of modern high quality

concrete. Today’s high performance concretes require superplasticizers in order to

reduce the water content while maintaining high workability, so that better

mechanical integrity and lower permeability are obtained when compared to

conventional concretes.

Though the primary purpose of adding superplasticizers is to provide high

workability, their incorporation in concrete could also affect, sometimes negatively,

the hardening and hardened state properties (Ramachandran, 2002; Mailvaganam,

2001). Moreover, phenomena such as loss in the workability of concrete, rapid/slow

setting, air entrainment, excessive bleeding and segregation could result from the

improper use of superplasticizers. Therefore, a good understanding of the cement-

superplasticizer interaction is required for better utilization of high performance

concrete, and for the further improvement of the properties of the superplasticizers

and concrete.

The performance of a superplasticizer in concrete depends on its properties, as well as

that of the cement, in addition to the concrete mix proportions and environmental
conditions (Jolicoeur and Simard, 1998). The major characteristics of cement that

could affect the performance of chemical admixtures are the C3A/SO3 ratio, alkali

content and nature of sulphates whereas the chemical structure and molecular weight

of polymers are the principal characteristics of superplasticizers (Rixom and

Mailvaganam, 1999; Ramachandran, 2002). The main action of the superplasticizers

is electrostatic repulsion and steric hindrance that result in the dispersion of the

cement particles (Ramachandran et al., 1998). Nevertheless, the four different families

- lignosulphonates, melamines, naphthalenes and polycarboxylates - each has different

distinguishing characteristics and effects (Rixom and Mailvaganam, 1999).

The proposed study focuses on the improvement of the flow behaviour and its

characterization, and the understanding of the microstructure of the superplasticized

cement-based materials. More specifically, the optimization of the superplasticizer

dosage, determination of the rheological behaviour, setting and evolution of the

hydration products are evaluated from tests of cement paste. The study also addresses

the issues of loss of workability and set retardation, which could arise as a result of

cement-superplasticizer incompatibility.

The relevance of the present work in the Indian context can be appreciated by

considering the following points. The types and number of superplasticizers available

in the market have increased significantly over the past few years. The usage of

products without adequate knowledge of their effects, along with the use of high

superplasticizer dosages, has led to significant problems during construction, some of

which can be attributed to cement-admixture incompatibility. The current knowledge

of materials and practices involved in the use of superplasticized concrete is mostly

based on foreign experience, which is not always directly applicable to construction in

India today. For example, the tendency to use synthetic gypsum instead of natural
2
gypsum during cement production results in the increase of superplasticizer dosage

and setting time. Also, ready-mix concrete in most construction sites is expected to be

placeable even 3-5 hours after mixing.

The final aim of the work is to identify the factors that affect the cement-

superplasticizer interaction and to propose remedial measures for the concrete user to

avoid incompatibilities through a materials science approach. Commercially-available

superplasticizers of different types, several cements, mineral admixtures and viscosity

modifying admixtures are used in the study in order to span the wide spectrum of

products used in the manufacture of high performance concrete.

1.2 OBJECTIVES AND SCOPE

The present thesis focuses on the study of the mechanisms that affect the cement-

superplasticizer interaction, which in turn influence the fresh and hardened state

properties of the cement based material. The main objectives of the research are:

• To identify a suitable mixing method for the preparation of pastes such that it

will represent the paste phase of concrete.

• To evaluate the flow behaviour of superplasticized cement paste using simple

methods, and correlating it with rheological characteristics for validating the

use of simple test methods.

• To analyze the correlation between the fluidity of cement paste, mortar and

concrete, in order to validate the optimization of the cement paste composition

as the first step in the mix design of high performance concrete.

• To study the influence of superplasticizer on the hydration processes and on

the microstructural development of cement paste.

3
• To identify the characteristics of cement and superplasticizer that can affect

the behaviour of the cement-superplasticizer system and propose guidelines to

avoid incompatibilities.

The scope of the study is limited to the following with respect to main components

and methods adopted:

• Four types of cements available in India and products from four families of

superplasticizers.

• Empirical correlation between the behaviour of cement paste and concrete is


made through experimental studies.

1.3 STRUCTURE OF THE THESIS

The research strategy that was followed in order to achieve the objectives specified in

the previous section and the results obtained are described in the following chapters of

the thesis.

A detailed review of the influence of superplasticizers on the hydration of cement is

given in Chapter 2, along with appropriate comparisons with the hydration of pure

cement. The characteristics, chemistry and mechanisms of action of superplasticizers

are also presented. The factors affecting the cement-superplasticizer interaction,

mainly, the chemical structure of the superplasticizer and the cement characteristics

are reviewed, and the need and motivation for conducting this study are emphasized

in the same Chapter.

Chapter 3 begins with the selection of a suitable procedure for the preparation of

superplasticized cement paste that simulates the paste phase of concrete. The

determination of the saturation dosage of superplasticizer with the Marsh cone is

explained in detail. The influence of the type and dosage of superplasticizer on the

4
flow behaviour of paste is evaluated through the Marsh cone, mini-slump and

viscometer tests. In order to validate the use of simple engineering-level test methods,

the Marsh cone flow time and mini-slump spread are shown to correlate well with

rheological parameters through appropriate analytical models. This chapter also

analyses the influence of saturation dosage of superplasticizer on the loss of fluidity of

different cement-superplasticizer combinations. Appendix A contains the results for

establishing the correlation between simple tests with rheological parameters. The

matlab programme for predicting the Marsh cone flow time using rheological model

is presented in Appendix B. The viscoelastic characterization of pure cement paste

and superplasticized cement paste, using creep and recovery and stress relaxation

tests, is given in Appendix C. The influence of type and dosage of superplasticizers on

the change in rheological response, from viscoelastic to viscous nature, is also

explained.

Chapter 4 deals with the correlation between the flow behaviour of superplasticized

cement paste, mortar and concrete. The setting behaviour of paste and the set

retardation due to the use of superplasticizers are studied with the Vicat apparatus and

electrical conductivity measurements. The responses of paste, mortar and concrete in

terms of loss of fluidity and setting are compared using appropriate tests. The

influence of superplasticizer dosage on the compressive strength of concrete and its

evolution with age has also been studied.

The study of the effect of the superplasticizer on the hydration processes, as well as

on the microstructural development, through several complementary physico-

chemical characterization techniques like X-ray diffraction, scanning electron

microscopy, thermal methods and nuclear magnetic resonance is presented in

Chapter 5. It is intended to provide an overall picture of the influence of different


5
families of superplasticizers on the types of hydrated products as well as their rate and

amount of formation.

Chapter 6 presents the evaluation of the compatibility between different cements and

superplasticizers through flow and setting tests. The variations of the dosages and loss

of fluidity for different superplasticizers is explained based on an absorbance study

using a UV spectrophotometer. Subsequently, a methodology based on the test results,

for selecting a compatible combination of cement and superplasticizer is

recommended. The studied combinations are evaluated with the recommended

methodology and finally the same is validated with other approaches for selecting the

compatible combination.

Chapter 7 deals with the influence of mineral admixtures and viscosity modifying

admixtures on the cement-superplasticizer interactions. A mix design procedure for

self compacting concrete incorporating VMA is suggested in line with previous

studies. The validation of the use of the saturation superplasticizer dosage obtained

from tests on pastes is illustrated with the characterization of self compacting concrete

through the slump flow, V-Funnel and J-Ring tests.

The conclusions drawn from the research work are summarized and the guidelines for

selecting the superplasticizers are presented in Chapter 8. Also, the scope for further

research is highlighted.

  

6
CHAPTER 2

STATE-OF THE-ART REPORT ON CEMENT-


SUPERPLASTICIZER INTERACTION

2.1 GENERAL

Admixtures, both mineral and chemical, have become essential components of high

quality concrete and are added to concrete at the mixing stage to modify its properties

in the fresh and hardened state. ACI 116R (2000) defines the admixture as ‘a material

other than water, aggregates, hydraulic cement and fiber reinforcement, used as an

ingredient of concrete or mortar, and added to the batch immediately before or during

its mixing’. Common mineral admixtures are silica fume, fly ash, ground granulated

blast furnace slag and metakaolin, the incorporation of which can improve strength, as

well the durability of the concrete. The incorporation of chemical admixtures can vary

significantly with the application; e.g., air entraining admixtures are very common in

colder regions where the concrete is prone to freeze-thaw damage, set accelerators are

essential in wet-mix shotcreting, and retarders are common when the ready-mix

concrete has to be transported over long distances. However, the superplasticizer or

water reducer is an admixture that has become an important component of all high

quality concretes; the addition of the superplasticizer improves the workability at low

water content, and consequently leading to good fluidity in the fresh state, reduction

in concrete porosity and compositional stability when the concrete is being

transported and placed. The use of superplasticizers/high range water reducers in

concrete has played a central role in the development of high strength and high

performance concretes, without which such concretes would not be technically

feasible.

7
However, the performance of an admixture in concrete is dependent on many factors,

like the nature and amount of admixture, nature of cement and aggregates, water-

cement ratio and environmental conditions. In addition, some problems could

commonly arise as a result of the incompatibility between cement and water reducers,

such as rapid loss of workability, rapid/retardation of setting and low rates of strength

gain. Sometimes, the incompatibility could even occur between a particular admixture

and only some batches of the same brand of cement (probably made at different

plants), suggesting that the problem is complex from the user's point of view.

Therefore, the understanding of the effects of superplasticizers on the properties of

concrete is essential for further improvement of the properties and behaviour of both

the superplasticizer and concrete.

The objective of this chapter is to summarise the various factors responsible for the

physico-chemical reactions in order to have a better understanding of cement-

superplasticizer interaction. Some relevant aspects of cement composition and cement

hydration reactions are also described in order to understand the action of the

superplasticizer in a cementitious system. A detailed review of the different families

of superplasticizers, mechanisms of action and their characteristics that could control

their compatibility with cement is also given.

2.2 CEMENT CHEMISTRY

2.2.1 Raw Materials Used in Cement Manufacture

The raw materials used for the production of cement include calcareous materials and

argillaceous materials. CaO is obtained from limestone, principally calcite; the lime

can be either quick lime, dolomite lime or magnesium lime. In quick lime, the major

part is CaO or MgO. Dolomite lime indicates that there is 35-45% magnesium

8
carbonate while the magnesium lime indicates the presence of 5-35% of magnesium

carbonate in the limestone. Silica (SiO2) exists naturally in the fine state as different

crystalline polymorphs (e.g., α-quartz, cristobalite, tridymite) and as impure poorly

crystallized or amorphous minerals (e.g., opal and flint). It is introduced through the

aluminosilicate minerals of the shale or clay component of kiln feed and combine with

CaO during clinkering process to form the impure calcium silicates, alite and belite.

The other two significant oxides, namely aluminium oxide that occurs in nature as α-

Al2O3 (Corundum) and iron oxide are derived from clay and shale (Hewlett, 2004).

Gypsum and anhydrite (i.e., anhydrous calcium sulphate) are found naturally in many

parts of the World; in Asia, India has the largest reserves of natural gypsum, 95% of

which is found in Rajasthan (Shah, 2006). Natural gypsum used for cement

production in India could be from marine or desert sources. Calcium sulphate is also

available as a byproduct of the acid based manufacture of fertilizers from phosphate

ore. However, when this is used in cement production, it causes an increase in setting

time due to the presence of water soluble fluorine, water soluble phosphates and

phosphates in the crystal lattice of gypsum (Hewlett, 2004). Nevertheless, such

synthetic gypsum of good quality is readily available and, consequently, there is a

growing in tendency to use it instead of natural gypsum in the Indian cement

manufacturing industry.

2.2.2 Cement Composition

Portland cements are multicomponent multiphase inorganic materials and consist of

five major constituents in varying amounts. In the early ages of a cement-based

material, the two calcium silicates – tricalcium silicate (C3S) and dicalcium silicate

(C2S) control the strength and its evolution, and the two aluminates - tricalcium

9
aluminate (C3A) and tetra calcium alumino ferrite (C4AF) and the sulphate phases

(i.e., gypsum) influence the setting behaviour (Hewlett, 2004). In addition to these

main compounds, there are minor compounds, such as MgO, TiO2, Mn2O3, K2O and

Na2O. Two of the minor compounds, i.e., sodium and potassium oxides, known as

alkalis, play an important role in the cementitious system in terms of the rheological

behaviour, and cement-superplasticizer interactions.

Standards such as those of ASTM, BIS and CEN give the acceptable ranges of

components for different classes of cements. For example, the ASTM C 150 (2007)

standard, limits the tricalcium silicates to 35%, tricalcium aluminates to 7%, and tetra

calcium alumina ferrite to 25%, and specifies a minimum dicalcium silicate content of

40%, along with a maximum alkali content of 0.6% (Na2O)eq (Hewlett, 2004); a

typical cement corresponding to this standard could have the composition as shown in

Table 2.1.

Table 2.1 Typical portland cement composition


conforming to ASTM C 150

Type of compound Composition (%)


C3S and C2S 70-80
C3A 0-16
C4AF 1-17
MgO 0.5 -6
Na2SO4, K2 SO4 0.5-3
Free lime 0.2-4
Gypsum 4
Low alkali cement < 0.6 (Na2O)eq

The chemical and physical requirements of different grades of cement as per the IS

codes are given in Table 2.2. It should be noted that the IS codes do not specify the

limits for alkalis (as done by ASTM C 150 (2007))which lead to significantly

10
different rheological behaviour especially if the alkali content varies considerably.

Similarly, the code does not prescribe upper limits of the alumina ratio and specific

surface area, both of which can affect the performance of the cement in the fresh and

setting stages. It can be seen from the table that the only differences between the different

grades of cement are in the rate of compressive strength gain and the 28-day value.

Table 2.2 Requirements of different grades of cement as per IS codes.

Requirements
33 grade (as per 43 grade (as per 53 grade (as per
Characteristics IS 269-1989) IS 8112-1989) IS 12269-1987)
Lime saturation Not greater than Not greater than Not greater than
factor * 1.02 and not less 1.02 and not less 1.02 and not less
than 0.66 than 0.66 than 0.8
Alumina ratio** Not less than 0.66 Not less than 0.66 Not less than 0.66
Insoluble residue Not more than Not more than 2% Not more than 2%
by mass 4%
Magnesia, % by Not more than Not more than 6% Not more than 6%
mass 6%
Total sulphur Not more than 2.5 Not more than 2.5 Not more than 2.5
content calculated and 3.0 when C3A and 3.0 when C3A and 3.0 when C3A
as sulphuric % by mass is 5 or % by mass is 5 or % by mass is 5 or
anhydride (SO3 ), less and greater less and greater less and greater
% by mass than 5, than 5, than 5,
respectively respectively respectively
Total loss on Not more than Not more than 5% Not more than 4%
ignition 5%
Fineness Specific surface Specific surface Specific surface
shall not be less shall not be less shall not be less
than 225 m2/kg than 225 m2/kg than 225 m2/kg
Initial setting time Not less than 30 Not less than 30 Not less than 30
minutes minutes minutes
Final setting time Not more than Not more than 600 Not more than
600 minutes minutes 600 minutes
Compressive For, 72 ± 1h, not For, 72 ± 1h, not For, 72 ± 1h, not
strength less than 16 MPa less than 23 MPa less than 27 MPa
For, 168 ± 2h, not For, 168 ± 2h, not For, 168 ± 2h, not
less than 22 MPa less than 33 MPa less than 37 MPa
For, 672 ± 4h, not For, 672 ± 4h, not For, 672 ± 4h, not
less than 33 MPa less than 43 MPa less than 53 MPa
*Lime saturation factor = Ratio of percentage of lime to percentage of silica,
alumina and iron oxide represented as (CaO - 0.7 SO3)/ (2.8 SiO2 + 1.2 Al2O3 +
0.65 Fe2O3).
**Alumina ratio = Ratio of percentage of alumina to iron oxide.

11
Though cements are classified by national or international standards, as seen earlier,

the chemical composition of a type of cement can vary within certain limits between

brands and even batches, depending on the raw materials and production processes.

For example, in Table 2.3 (from Hewlett, 2004), which gives the chemical analyses of

six different CEM I portland cements available in the UK in 1987, it can be seen that

the compound composition, fineness and particle size distribution vary significantly

from one cement to another, though they all belong to the same “standard” type.

Consequently, different cements of the same type/grade can have significantly

different characteristics that can lead to differences in the interactions between the

cement and admixture.

Table 2.3 Chemical analyses of some CEM I portland cements of UK (Hewlett, 2004)

Type of DO DOX EO FO GO HO
compound (%)
Si O2 22.2 21.8 20.9 20.1 19.5 20.2
Al2 O3 4.2 4.1 4.6 5.0 6.2 5.3
Fe2 O3 2.4 3.3 2.1 2.8 2.0 2.2
Mn2 O3 0.08 0.07 0.06 0.16 0.03 0.04
P 2 O5 0.15 0.1 0.23 0.1 0.17 0.07
Ti O2 0.27 0.23 0.2 0.2 0.31 0.23
Ca O 63.9 64.2 66.3 63.2 64.7 66.4
Mg O 1.0 1.4 0.8 2.2 1.2 0.9
SO3 2.7 2.5 2.8 3.0 3.0 2.5
CO2 0.79 0.3 0.48 ----- ---- ----
K2 O 0.74 0.61 0.8 0.78 0.43 0.56
Na2 O 0.14 0.15 0.16 0.13 0.24 0.23
Free lime 1.5 1.4 1.4 2.1 0.9 1.2
Loss on ignition 1.7 1.5 0.75 1.0 1.1 0.7
Insoluble residue 1.7 0.8 1.1 0.84 0.85 0.5
Na2O equivalent 0.63 0.55 0.69 0.64 0.52 0.6
C3 S 46 51 63 50 58 66
C2 S 29 24 12 20 12 8
C3 A 7.1 5.3 8.6 8.5 13 10.3
C4AF 7.3 10 6.4 8.5 6.1 6.7
Density(kg/m3) 3130 3130 - 3150 3110 3140
Surface area 383 380 - 443 394 344
(m2/kg)
Particles coarser 13.4 12.5 - 9.0 12.3 17.4
than 45µm

12
2.2.3 Hydration of Cement

2.2.3.1 Hydration of Cement Compounds

Hydration of cement is the key to concrete performance in terms of setting, durability

and strength. It includes the hydration of calcium silicates, calcium aluminates and

calcium aluminoferrite. When water is added to cement, each of the compounds

undergoes reactions but with different characteristic contributions to the setting,

strength gain and other properties of the hardened cement paste; e.g., tricalcium

silicate hydration affects the early age response while the hydration of dicalcium

silicate affects the strength at later ages. The relevant aspects of the hydration of the

various components of cement are explained in the following paragraphs. It should be

noted that, by volume, calcium silicate hydrate (C-S-H) constitutes 60-70% of the

hydration products, calcium hydroxide constitutes 20-25% and 5-15% is made up of

other minor phases.

Upon the addition of water, tricalcium silicate (C3S) rapidly reacts to release calcium

ions, hydroxide ions and a large amount of heat. The pH quickly raises to over 12

because of reaction alkaline hydroxide (OH-) ions. This initial hydrolysis slows down

quickly resulting in a decrease in the heat evolved. The reaction continues slowly,

producing calcium and hydroxide ions until the system saturates. Once this occurs,

Ca(OH)2 starts to crystallize. Simultaneously, calcium silicate hydrate begins to form.

Subsequently, calcium and hydroxide ions precipitate out of solution accelerating the

hydration of tricalcium silicate. The C-S-H crystals grow thicker making it more

difficult for water molecules to reach the unhydrated tricalcium silicate.

Consequently, the speed of reaction is controlled by the rate at which water molecules

diffuse through the C-S-H coating on the surface of the unhydrated cement particles.

As the coating thickness increases, the production of C-S-H becomes slower and

13
slower (Jolicoeur and Simard, 1998). The overall chemical reaction and the heat

evolved (∆H) are given in the following equation:

2C3S+6H => C3S2H3+ 3 CH (∆H = - 520 kJ/kg) (2.1)

Dicalcium silicate reacts with water in a similar manner as C3S but much more

slowly, according to the reaction and heat of hydration shown below:

2C2S+4H => C3S2H3+ CH (∆H = - 260 kJ/kg) (2.2)

The hydration of tricalcium aluminate (C3A) in the presence of gypsum occurs as

follows:

2 C3A+ H+ CS => C6AS3 H 32+ 3CH (∆H = - 910 kJ/kg) (2.3)

Calcium aluminate hydrate or ettringite (AFt) needles cover the surface of hydrating

calcium aluminate particle and prevent further hydration. If the supply of sulphate

from gypsum is exhausted before C3A is completely hydrated, a second reaction can

occur resulting in the formation of calcium monosulphoaluminate (AFm). This

reaction may occur before the formation of the ettringite if the availability of gypsum

is lower than that required by the rate of reaction of C3A and sulphate ions. If

monosulphoaluminate is exposed subsequently to sulphate ions, new reactions will

occur leading to the formation of more ettringite.

Calcium aluminoferrite forms the same hydration products as C3A, with a slow rate of

reaction that is decreased further by gypsum and higher iron oxide content. The

reaction is as given by the following equation:

C4AF + CS H2 +12H => C4 (A,F) CSH12 (∆H = - 420 kJ/kg) (2.4)

The relative reactivity of different mineral phases with water can be given as C3A>

C3S> C2S> C4AF; accordingly, the aluminate phases and their hydration products play

an important role in the early hydration processes and cement-admixture interactions.

14
2.2.3.2 Hydration of the Portland Cement Grain

The development of the microstructure of a typical polymineralic cement grain is

shown in Figure 2.1, and the major unhydrated compounds present are shown in

Figure 2.1 (a). After the first contact of cement with water, various reactions occur

that result in a typical heat evolution pattern with time, through which four stages can

be identified in the early hydration process: initial hydration period, dormant period,

acceleration and deceleration period. The basic mechanisms mostly influenced by the

presence of admixtures, particularly those occurring in the early hydration processes

are described below (Mehta and Monteiro, 2005; Jolicoeur and Simard, 1998; Taylor,

1997; Neville, 2000; Hewlett, 2004):

Initial hydration (0-15 minutes): During this period, rapid heat generation, called as

heat of wetting, occurs on mixing cement with water. Easily soluble components like

alkalis, calcium, sulphate phases and free lime are dissolved in the surrounding water

and the solubilization of variety of ionic species like Na+, K+, Ca2+, SO42-, and OH-

ions takes place. Hence, the wetting of cement particles and subsequent dissolution of

ions lead to the formation of an amorphous gel layer over the surface of grains.

Beyond the solubilization process, the formation of solid hydration product is

governed by the nucleation process. The initial nucleation process involves the

formation of calcium sulphoaluminate or ettringite from Ca2+, SO42-and Al(OH)4- ions

on the outer surface of layer of gel, as shown in Figure 2.1 (b). This reaction is the

most important as far as the rheology of the fresh cement paste is concerned. The

initial C3A dissolution may be increased by the presence of more dissolved alkalis

(Spiratos et al., 2003). After some time, the cement grains are coated with a protective

layer of hydration products. At this stage, the rate of reactions sharply reduces

resulting in less heat of hydration.

15
The presence of organic admixtures can interfere with these nucleation and growth

processes resulting in the change of the hydration reactions and hydration products.

Also, after the first few minutes, the chemical admixtures interact with these

hydration products instead of hydrating compounds.

(a) (b) (c) (d) (e) (f)

Fig. 2.1 Microstructure of hydrated portland cement (Taylor, 1997)

Dormant period (2-4 hours): During this period, the mix remains plastic and very

low heat generation takes place, indicating that all the reactions are slow. This is due

to two reasons: (1) the gel formed in the first stage acts as a diffusion barrier around

the C3S; (2) the SiO2 rich electrical double layer formed around C3S and the

supersaturation of Ca2+, OH- and SO42- in solution prevents further dissolution of ions.

As the thickness increases, the time it takes for water to penetrate the coating

increases; thus the reaction becomes diffusion controlled.

In the presence of an adequate concentration of SO42- ions, continued growth of

ettringite crystals takes place. Simultaneously, the reaction of C3S produces ‘outer

product’ C-S-H on the AFt, leaving space between the grain and hydrated shell, as

shown in Figure 2.1 (c). An inadequate concentration of SO42- ions leads to the

16
formation of calcium aluminate hydrate (C-A-H) resulting in flash set. If the SO42-

concentration is too high due to the presence of hemihydrates and alkali sulphates,

massive nucleation and growth of gypsum crystals (false set) occurs.

The interaction of a chemical admixture with any of these reactions or reactive species

significantly affects the behaviour of cement pastes and concretes during the induction

period.

Acceleration period (4-8 hrs): Several theories have been put forward to explain the

end of the dormant period or the start of the acceleration period: Weakening of the

double layer by further hydration, increased diffusion and weakening of ionic strength

around the hydrating particle all result in the end of the dormant period and the

beginning of a new stage of high chemical activity, called the acceleration period.

However, during this period, the suspension loses its plasticity and is converted to a

stiffer matrix, which is no longer fluid. Subsequently, the secondary hydration of C3A

produces long rods of AFt and the intense hydration of C3S during the acceleration

period results in the formation of ‘inner product’ C-S-H and precipitation of C-H, as

shown in Figure 2.1 (d). Later, C2S starts to hydrate and, to a lesser extent, C4AF

continues to hydrate. During the acceleration period, the calcium and sulphate ion

concentrations in the pore water decrease due to the ettringite formation.

The chemical admixtures may influence the formation and properties of the protective

hydrate layer in the acceleration stage. Also, the admixtures remaining in the aqueous

solution further influence the nucleation and growth of the hydration products.

Deceleration period (8-24 hours): The reactions are slow and completely diffusion

dependent in this period, which is reflected by the hardening of the cement paste. Due

17
to the depletion of sulphate ions in the pore water, C3A reacts with the AFt forming

hexagonal plates of monosulphate (AFm), as shown in Figure 2.1 (e). The pore

volume decreases with increasing time and decreasing free water content as hydration

progresses with the formation of C-S-H gel and C-H. This is followed by a steady

state (12-24 hours) in which temperature has little influence on hydration.

Later hydration results in the formation of sufficient ‘inner’ C-S-H to fill the space

between grain and shell, as shown in Figure 2.1 (f). The outer C-S-H becomes more

fibrous during this period and the ettringite is completely converted to

monosulphoaluminate within 1 to 3 days.

2.3. WATER REDUCERS AND SUPERPLASTICIZERS

Water reducers are classified as normal, medium and high range depending on the

level of water reduction possible due to its addition in concrete. Water reducers

belong to the dispersant family; dispersants are long chain organic molecules having a

polar hydrophilic end and non-polar hydrophobic groups that are adsorbed on the

cement particles. Superplasticizers are high range water reducers and are incorporated

in concrete either to maintain the same workability at reduced water cement ratio or to

increase the workability by maintaining constant w/c. These admixtures are synthetic

high molecular weight water-soluble polymers that are effective in dispersing the

flocculated cement grains. Solubility is achieved by the presence of adequate

hydroxyl, sulphonate or carboxylate groups attached to the main organic unit which is

usually anionic (Rixom and Mailvaganam, 1999; Ramachandran, 2002; Bentur,

2002).

After the introduction of the first generation of superplasticizers in 1960s, these

products have become essential for the placement of concrete in areas with low

18
accessibility or with high density of reinforcement, as well as to provide an increase in

the pumpability of concrete. In general, however, the reduction of w/c that can be

obtained with the use of superplasticizers leads to higher strength and durability,

making them common components of high performance and high quality concretes.

Also, newly developed concretes such as self compacting concrete require the

presence of a superplasticizer to achieve the desired properties.

2.3.1 Characteristics and Chemistry of Superplasticizers

Superplasticizers belong to four different families: modified lignosulphonates (LS),

sulphonated naphthalene formaldehyde (SNF), sulphonated melamine formaldehyde

condensates (SMF), and the new generation of superplasticizers, which is made up of

copolymers that include polycarboxylates (PCE), polyphosphonates, polyacrylates

and monovinyl alcohols (Rixom and Mailvaganam, 1999; Ramachandran, 2002). The

different families of superplasticizers are generally based on chemicals of the type

presented in Table 2.4.

The lignosulphonate based admixtures are normal water reducers and at higher

dosages of such admixtures, excessive retardation of set and entrainment of air occurs,

which limits their dosage in practice (Ramachandran, 2002). Nevertheless, modified

lignosulphonates, in which the sugar content is reduced, exhibit less retardation

(Aïtcin, 1998). At higher molecular weights, they can be considered as

superplasticizers and used for obtaining workable concrete, with some retardation of

initial set (Ramachandran, 2002). It has also been found that sodium lignosulphonate

is a better plasticizer than calcium lignosulphonate because the former has higher

solubility.

19
Table 2.4 Chemical structure of superplasticizers (Rixom and Mailvaganam, 1999)

Class Origin Structure (typical repeated unit)


Lignosulphonates Derived from
(LS) neutralization,
precipitation, and CH2OH
fermentation OH
H O CH CH
processes of the
waste liquor SO3M n
O
obtained during Me
production of Me =CH3 , M =Na
paper-making
pulp from wood
Sulphonated Manufactured by
melamine normal H H
formaldehyde resinification of
HO CH N N N CH2 H
(SMF) melamine -
formaldehyde N N
n
HNCH2SO3M
M = Na

Sulphonated Produced from


naphthalene naphthalene by
formaldehyde oleum or SO3 R R
(SNF) sulphonation;
subsequent CH2 H
reaction with
formaldehyde n
leads to
polymerization SO3M SO3M
R =H,CH3,

CH
Polycarboxylic Free radical
ether (PCE) mechanism using
peroxide initiators CH2 CH CH2 CH2
is used for
C=O C=O
polymerization
process in these OCH3 OCH2CH2(EO)12CH2CH2
n
systems
EO: Ethylene oxide

The first admixtures to be classified as superplasticizers were salts of SNF and SMF.

All of these have the same functional groups, namely sulphonates. The use of SNF

and SMF based superplasticizers resulted in significant water reduction at low

20
dosages. The SNF based superplasticizers appear to yield the best fluidizing effect in

the molecular weight range of 4000-40000 gm/mole, while at higher ranges the effect

decreases (Ramachandran, 2002). SMF based superplasticizers generally have poorer

slump retention characteristics, when compared with the SNF based products, and are,

consequently, used more extensively in the pre-cast industry.

In the last few years, several polymeric surfactants with carboxyl, hydroxyl, and

phosphonate functional groups have been introduced in the market as third generation

superplasticizers, and are generically known as polycarboxylates. These are comb

polymers and are characterized by long chains and side chains, in which the

hydrophilic groups are in contact with water and hydrophobic groups are attached to

cement particles (Tanaka et al., 1999; Sakai et al., 2003). The dispersion due to these

superplasticizers depends on the length of the main chain, and the length and number

of the side chains. In general, the polycarboxylates are sufficient in low dosages due

to better dispersing action than first and second generation superplasticizers.

Moreover, it appears that they can be tailored to fulfill the requirements of the

application (Magarotto et al., 2003; Houst et al., 2005; Falikman et al., 2005). The

mechanisms of action of superplasticizers are explained with more detail in the

following section.

2.3.2 Mechanisms of Action of Superplasticizers

The rheology of cement paste is governed by van der Waal's attractive forces between

the cement particles and the electrostatic repulsion due to the surface charges on the

cement particles. However, as the former force is larger than the latter, cement

particles tend to eventually flocculate. The action of the superplasticizer is to prevent

or delay the flocculation and to disperse the cement particles within the paste. The

21
different interactions between the cement and superplasticizer, which are often

classified as physical or chemical interactions, are described in detail in the following

paragraphs.

Flatt and Houst (2001) proposed that the behaviour of the superplasticizer in the

cement paste is made up of three components, which was later supported by Banfill

(2003) and Yamada et al. (2006). The first part of the behaviour is the absorption of

superplasticizer by intercalation (i.e., the incorporation of molecules of

superplasticizer into the precipitates formed as the cement hydrates), co-precipitation

(i.e., precipitation of tiny solid nuclei of the superplasticizer along with the

precipitation of the cement hydrates) or micellization (i.e., groups of polymer

molecules are trapped within the hydrating cement making them unavailable for

active dispersion), leading to the formation of an organo-mineral phase (Banfill,

2003). These phenomena that occur during the formation of ettringite and C-S-H

decrease the amount of admixture available for dispersing the flocs of cement later on.

Consequently, for the same superplasticizer dosage, a cement-admixture combination

with a high degree of absorption will yield less fluidity than another system with a

lower degree of absorption. Nevertheless, the consumption of the superplasticizer by

absorption is generally small, especially when the superplasticizer addition is delayed

(i.e., the superplasticizer is added at least one minute after the addition of water to the

cement) (Banfill, 2003).

The second part of the superplasticizer is adsorbed on the cement particles, resulting

in their dispersal due to the formation of an electrical double layer and consequent

electrostatic repulsion between the particles. The chemical nature of the cement

compounds that form the solid particle makes them undergo ionisation on the surfaces

22
that are in contact with water. The electrical double layer (Uchikawa et al., 1997;

Young, 2008) appears as a consequence of the adsorption of anionic molecules of

superplasticizers on the cationic cement particles. The organic superplasticizer

molecules, having charged groups (SO3-, COO-) interact with the particle surfaces,

imparting negative charges to the mineralogical phases of cement (see Figure 2.2a),

which results in the electrostatic repulsive forces (Aïtcin, 1998; Rixom and

Mailvaganam, 1999; Yoshioka et al., 2002). Mollah et al. (1995) extended the double

layer theory by considering that the cement particle is covered by a layer formed by

Ca2+ ions resulting in a trilayer of a “diffuse ion swarm”, which also delays the

hydration reactions. Due to the dispersive action of the superplasticizer, water that

would be otherwise entrapped within the flocs is released for providing fluidity to the

paste. In other words, the presence of the superplasticizer reduces the amount of water

needed for the effective dispersion of the cement particles. This type of action is

observed in the cases of lignosulphonates, melamines and naphthalenes. The

electrostatic repulsion generated depends on the composition of the solution phase and

the adsorbed amount of the superplasticizer (i.e., greater the adsorption, better the

repulsion and higher the fluidity) (Nakajima and Yamada, 2004). As the electrostatic

potential diminishes, flocculation of cement particles occurs and the cement paste

eventually loses fluidity.

Another type of physical interaction, also based on adsorption, that is observed in the

case of polycarboxylates, in addition to electrostatic repulsive forces, is that the side

chain polymer molecules cause steric hindrance between the cement particles,

preventing flocculation and resulting in their dispersion (Rixom and Mailvaganam,

1999; Corradi et al., 2005; Collepardi, 2005) (see Figure 2.2b). The steric repulsion

depends on the length of main chain, and the length and number of side chains

23
(Yamada et al., 2001; Sugiyama et al., 2003). This action last for a longer time than

the electrostatic repulsion and hence leads to better performance.

Cement particle Cement particle

(a) Electrostatic repulsion

Cement particle Cement particle

(b) Steric hindrance


Fig. 2.2 Physical action of superplasticizers (Aїtcin, 1998)

When electrostatic repulsion is the main mechanism of dispersion, the zeta potential

represents a measure of the electrostatic interaction between the cement particles. The

superplasticizer, being anionic, leads to a higher negative zeta potential that results in

strong superplasticizer adsorption (Krishna, 1993, 1996; Houst et al., 1999a). It

should be noted that the different phases of cement have differences in zeta potential

(Yoshioka et al., 2002) that result in preferential adsorption of the superplasticizers on

some phases such as C3A. Similarly, the zeta potentials of early hydration products

can differ and influence the adsorption of the superplasticizer (Plank and

Hirsch, 2007). For example, the zeta potential of ettringite is more positive than that

of calcium monosulphate (Plank and Hirsch, 2007), which causes more

24
superplasticizer to be adsorbed on the former. Nevertheless, the maximum amount of

superplasticizer adsorbed, in addition to the zeta potential, is also dependent on the

type and morphology of the phase or hydration product. Accordingly, Jolicoeur and

Simard (1998) and Ramachandran (2002) have determined that the amount of

superplasticizer adsorbed on the monosulphate is higher than on the ettringite (though

the zeta potential of the latter is higher), probably due to the higher surface area of the

former.

The scientific literature on superplasticizer actions generally associates the fluidity of

the paste with the amount of superplasticizer adsorbed by the cement grains. It has

been stated that optimal fluidity of the paste occurs when the maximum adsorption is

reached on the cement surfaces (Houst et al., 1999a; Roncero, 2000; Griesser, 2002).

As mentioned earlier, the maximum adsorption depends on the chemical composition

of the cement, as well as the chemical structure of the superplasticizer. It should be

noted that when a higher amount of superplasticizer is adsorbed, there is better initial

fluidity but the fluidity may not be maintained sufficiently over time (Nkinamubanzi

et al., 2000; Li et al., 2003; Corradi et al., 2005). In such cases, the dosage of the

superplasticizer has to be increased to provide an additional amount for maintaining

the fluidity.

In addition to the physical effects of the adsorption, there are chemical reactions

between the adsorbed superplasticizer and cement grains that occur mainly at the most

reactive sites of cement particle surface (Jolicoeur and Simard, 1998). For example,

polynaphthalene sulphonate reacts with C3A in competition with SO42- ions. Jolicoeur

and Simard (1998) further points out that many organic admixtures solubilize the

ionic species, through association or complexation, which delays the precipitation of

25
the hydrated products. Also, the presence of organic molecules in the solid solution

inhibits crystal nucleation, and causes slower growth of the hydration products and a

change in their morphology. This is attributed to the blocking of the pores or gaps in

the initial protective hydrate or the inhibition of the more reactive sites by the low

molecular weight superplasticizer molecules, as shown in Figure 2.3.

High molecular weight Low molecular weight

Surface sites having affinities for RSO4 –-- or SO4--- ions

Fig.2.3 Chemical action of superplasticizers (Aïtcin, 1998)

Furthermore, Gaidis and Gartner (1991) suggest that the strong adsorption of the

superplasticizer on the cement grain affects the surface charge that exerts an

immediate deactivating effect on the C3A, which helps in fluidifying the cement paste.

However, after the admixture is completely removed from the solution, the C3A

resumes a more normal hydration.

The third part of the superplasticizer is non-adsorbed species that may be adsorbed

subsequently maintaining the dispersion of cement particles leading to flow retention.

The unadsorbed molecules also lead to additional mechanisms of action of the

superplasticizer including the dispersion of cement particles by reducing the surface

tension of the mixing water, the decrease in frictional resistance owing to the lining-

up of linear polymers along the flow direction, and the lubrication produced by the

26
low molecular weight polymers between the particles (Uchikawa et al., 1995; Tanaka

et al., 1999).

Furthermore, some additional mechanisms supposedly enhance the performance of

PCEs in terms of flowability and segregation resistance, as discussed by Tanaka et al.

(1999). The ingress of non-adsorbed low molecular weight superplasticizer molecules

between the cement grains causes additional dispersion due to volume repulsion

(depletion effect) whereas the higher molecular weight polymers that cannot ingress

between the cement particles coagulate (depletion coagulating effect) to improve the

segregation resistance. Even the higher molecular weight fractions may not be

completely adsorbed on the cement surface beyond a certain critical concentration

(called as the critical micelle concentration for surfactants; Porter, 1994) giving rise to

the effects given above.

The formation of air bubbles in the cement paste due to the incorporation of the

superplasticizer can also help in its fluidification but also could reduce the strength

and durability if the air content is high. The different types of admixtures have

differing influences on air entrainment in concrete (Mailvaganam, 1999; Rixom and

Mailvaganam, 1999; Ramachandran, 2002). Lignosulphonates incorporate significant

amounts of air, which limits their dosage in concrete, in addition to other reasons

already discussed earlier. The SNF and polycarboxylates can also incorporate air in

the concrete due to which deairing agents are sometimes added to the commercial

product. The SMF based superplasticizer, on the other hand, does not have any

significant air entraining effect. Nevertheless, some of the air entrained due to the

admixture incorporation is released when the concrete is vibrated, especially when it

is highly workable.

27
2.3.3 Factors Affecting the Cement-Superplasticizer Interaction

Though there is significant knowledge about the mechanisms of action of

superplasticizers, there still exist gaps in the comprehension of why occasionally these

chemicals do not work as intended. This is generally termed as incompatibility

between the cement and superplasticizer; the term incompatibility refers to the

adverse effect on the performance when a specific combination of cement and

superplasticizer is used. Common problems of incompatibility include rapid setting,

delayed setting, rapid slump loss, improper strength gain, etc. These problems can in

turn affect the productivity in the construction site and the hardened properties of

concrete, primarily strength and durability.

The cement-superplasticizer interactions have many dimensions. On the one hand,

there is the influence of the type of superplasticizer, and at the other end of the

spectrum are the effects of the composition of cement, particularly the relative

proportions of C3A, alkalis and C3S in the cement. In addition, the type of gypsum

incorporated in the cement (gypsum, hemihydrate or anhydrite) has an important role

to play. The fineness of cement could also affect its compatibility with a particular

admixture. It is, therefore, necessary to understand the factors governing the

interactions and the reasons for incompatibilities. This is all the more important since

the type and number of superplasticizers available in the market has increased, along

with the use of high superplasticizer dosages. The various factors affect cement-

superplasticizer interactions are discussed in further detail in the following sections.

2.3.3.1 Effect of Chemical Structure of Superplasticizer

The type and dosage, degree of polymerization, degree of sulphonation, the position

of functional group in the benzene ring, the molecular weight distribution of the

28
polymer, the addition rate and the time of addition of superplasticizer influences the

interaction between cement and sulphonate based superplasticizers, namely, SMF,

SNF and LS (Aїtcin et al., 2001).

In the case of the lignosulphonates (Mollah et al., 1995; Rixom and Mailvaganam,

1999), the presence of low molecular weight constituents is known to cause excessive

air entrainment leading to loss of strength. In addition, the high sugar content of these

admixtures could cause unnecessary retardation, especially at high dosages. A further

unpredictability might arise depending on whether the lignosulphonate is a sodium or

a calcium salt; Ramachandran (2002) states that the sodium lignosulphonate as a

better plasticizer than calcium lignosulphonate. This difference in effectiveness may

be due to the presence of alkalies or due to the higher solubility of former. In order to

be more effective, lignosulphonates are modified - the sugars are removed by

fermentation, and low molecular weight matter is removed by centrifuging - and used

as superplasticizers. Lignosulphonate admixtures easily produce a complex salt with

Ca2+, thus decreasing the Ca2+ concentration in the liquid phase, resulting in a delay in

the hydration of alite and causing set retardation.

The effect of superplasticizer on the fluidity is related with adsorption and zeta

potential, as mentioned earlier, which in turn depends on the molecular weight of

polymer. There is no consensus, however, on the exact relations among the molecular

weight, adsorption and effectiveness in terms of dispersion. Andersen et al. (1987)

suggested that superplasticizer with smallest molecular weight is the most adsorbed

whereas the polymer that is least adsorbed gives the highest negative zeta potential.

Later, Basile et al. (1989) reported that the effectiveness of SNF in increasing the

fluidity is increased when the monomer content is reduced and molecular weight is

29
increased. They also found that the polymers with higher molecular weight caused

more retardation of the cement hydration at early ages. Similarly, Bonen and Sarkar

(1995) suggested that SNF with higher molecular weight led to more fluidity and

retardation in setting. Rixom and Mailvaganam (1999) have clarified that for SNF, the

influence of its molecular weight on the rheology, as well as the hydration of the

cement paste, depends on the alkali content of the cement. In the case of high alkali

cement, SNF superplasticizers with higher molecular weight (e.g., 16000 g/mol) yield

better fluidity in the cement paste due to the higher surface charge provided by the

larger molecules leading to higher adsorption (Kim et al., 1999; Kim and Aїtcin,

2003; Page et al., 1999). For low alkali cement, an SNF with high molecular weight

does not retard the setting significantly but the incorporation of a lower molecular

weight results in a considerable increase in the induction period, possibly due to lower

adsorption of the latter (Rixom and Mailvaganam, 1999).

Another factor affecting SNF effectiveness is the location of the sulphonate (-HSO3)

group in the naphthalene structure. It is well accepted that the presence of the

sulphonate group in the β-position leads to a high polymer charge and better

electrostatic repulsion (Aïtcin, 1998). So, the degree of sulphonation is an important

parameter for the effectiveness of the SNF based superplasticizer.

The action of SMF based superplasticizers is similar to that of SNF based products.

Nevertheless, there are some differences. The adsorption characteristics of SMF based

superplasticizers are such that they have higher affinity for C3A than C4AF and C3S

(Ramachandran et al., 1998). The stability of SMF appears to vary with the

temperature and could be reason why it is not widely used, especially in countries

with warmer climates. Torresan and Khurana (1998) state that if the SMF is stored at

30
20°C, it could have a shelf life of more than 1.5 years whereas at 40°C, its shelf life is

reduced to 5-7 months. This is attributed to further condensation at higher temperature

and consequent increase in viscosity.

In the case of polycarboxylates, the backbone chain consists of acrylic or methacrylic

copolymers and various functional groups (polar or ionic, carboxyl or hydroxyl

groups) can be grafted on the backbone chain as side chains (e.g., polyethyleneoxide

graft chains). PCE type superplasticizers with high molecular weight (1400-88000

gm/mole) can be obtained with sizes of upto 30-150 nm (Uchikawa, et al.,1997).

However, it appears that the normal length of main chain is 20 nm and that of the side

chain is 7 nm (Ohta et al., 2000; Sakai et al., 2003). The variations in the type and

length of the main and side chains of the PCE type superplasticizers have yielded a

broad variety of new products with very different properties (Yamada et al., 2000;

Maeder and Schober, 2003; Magarotto et al., 2003; Sakai et al., 2003; Sugiyama et al.,

2003; Houst et al., 2005).

The differences in the molecular structure influence the amount of adsorption as well

as the initial rate of adsorption of PCE, which in turn affect the initial fluidity and

flow retention (Uchikawa et al., 1997; Maeder and Schober, 2003; Sugiyama et al.,

2003; Yamada and Hanehara, 2003). In general, higher molecular weight and

reduction of fractions with lower molecular weight leads to better flow in pastes

(Magarotto et al., 2003). Recently, Winnefeld et al. (2007) have proposed the use of

lower side chain densities and shorter side chains to obtain higher fluidity through

better adsorption, and the use of longer side chains for reducing the set retardation,

implying that the optimization of the molecular weight distributions is needed for the

best performance. They have also stated that polymers with higher charge density are

31
adsorbed more strongly and, consequently, result in better initial fluidity. For comb-

type PCE-based superplasticizers, the charge density is higher when more free

carboxylic groups are present (Winnefeld et al., 2007). The Table 2.5 summarises the

relations between the structural factors of the PCE and the dispersibility parameters

(Sakai et al., 2003; Sugiyama et al., 2003).

The adsorption of PCE could also depend on the amount of superplasticizer available

in the solution, and consequently on the dosage of superplasticizer. More precisely, at

low dosages, all fractions of the polymer are adsorbed whereas beyond a certain

dosage, only fractions of superplasticizer with larger molecular weights will be

preferentially adsorbed (Flatt et al., 1998; Winnefeld et al., 2007). This can lead to

differences in the effectiveness of different products since they could have fractions of

varying molecular weights and larger molecular fractions of varying effectiveness,

especially when steric repulsion is the dominant mechanism, as shown in a study of

Na-polycarboxylate-polysulphonate superplasticizers by Flatt et al. (1998).

Table 2.5 Structural aspects of the PCE that affect the fluidity of cement paste

Parameter Structural aspects


Relative chain Relative side Relative
length of back chain length number of side
bone polymer chains
Low dispersibility and short Long Short Large
dispersibility retention
High dispersibility Short Long Small
Long dispersibility retention Shorter Long Large

Compared with the polycondensates (SMF and SNF), the PCE based superplasticizer

is adsorbed much less due to its lower charge density (Houst, et al., 2005; Plank and

Hirsch, 2007). This is attributed to the fact that the functional groups of the PCE type

32
superplasticizers are weaker acids compared to sulphonate groups containing SMF

and SNF type superplasticizers, and therefore, PCE type superplasticizers are weaker

electrolytes and show a lower electronic activity. Among the polycarboxylates,

adsorption is more when the charge density is higher, giving better initial fluidity.

Generally, there is a problem in achieving high initial fluidity and maintaining the

same over a long period of time when using products that are currently available in

the market. Consequently, it has been suggested that the combination of polymers

with different rates of adsorption could lead to the retention of high slump in concrete.

Corradi et. al. (2005) have used this concept in the development of new PCE-based

superplasticizers that are supposed to have both water reduction and flow retention

properties. This is attributed to the addition of a new functional monomer that has a

lower adsorption rate than the more conventional component.

Furthermore, the response of the superplasticizers to the variation in chemical

composition of cement also depends on the type of superplasticizer. According to

Nkinamubanzi and Aïtcin (2004), the polycarboxylate based superplasticizer is less

sensitive to the alkali and soluble sulphate content of the cement than sulphonated

superplasticizers such as lignosulphonates and SNF. Further details about the

dependence of the chemical composition of the cement on the cement-superplasticizer

interaction are given in the following sections.

2.3.3.2 Effect of Calcium Sulphate

In order for normal set to occur in cement paste, sufficient amounts of soluble

sulphate and calcium ions should be available in the solution to form calcium

aluminosulphate (ettringite). When gypsum is incorporated in the portland cement

there is, generally, enough soluble sulphate for hydration to proceed normally.

33
However, if the gypsum is replaced partially or completely with natural anhydrite,

compatibility problems like rapid set and accelerated slump loss can occur especially

in pastes with low w/c and with superplasticizers having sulphonate functional groups

and polycarboxylates (Dodson and Hayden, 1989; Jiang et al., 1999; Mailvaganam,

1999; Nkinamubanzi and Aïtcin, 2004). This is because the rate of solution of natural

anhydrite is slower than gypsum and it is further reduced possibly due to the

adsorption of superplasticizer on the surface of natural anhydrite particles (Jolicoeur

and Simard, 1998; Hanna et al., 2000; Ramachandran, 2002). Consequently, the

availability of soluble sulfate ions in the aqueous phase is not sufficient to keep up

with that needed by the C3A, resulting in the stiffening of the cement paste due to

accelerated setting caused by the formation of monosulphate, which is converted later

to ettringite (Jolicoeur and Simard, 1998). Such incompatible behaviour is not

expected to occur when hemihydrate or soluble anhydrite or gypsum is used in the

cement since their solubility rates are not reduced by the addition of superplasticizer

(Dodson and Hayden, 1989). However, Fernon et al. (1997) report that, in lime

saturated solution, Na-SNF enhances the solubility of gypsum and hemihydrates

significantly and that of the anhydrite only slightly, whereas the Ca-SNF does not

affect the solubilities except that of the anhydrite, which is decreased. The increased

solubility of gypsum or hemihydrates in the presence of Na-SNF is attributed to the

formation of calcium ion complexes by the ionized sulphonates in an alkaline

medium.

Since the adsorption of the superplasticizer on the C3A and AFm is much more than

on ettringite, in competition with the SO42- ions, a sufficient amount of

superplasticizer is, consequently, not available in the solution for providing fluidity to

the paste. This results in the loss of fluidity or an increase in the superplasticizer

34
demand to compensate for this effect (Jolicoeur and Simard, 1998; Ramachandran,

2002). Alternatively, the superplasticizer dosage needs to be increased to provide

sufficient superplasticizer in solution so that paste fluidity is adequate. Fernon et al.

(1997) also caution that an excess of dissolved calcium sulphate can lead to rapid

slump loss due to the alteration of the hydration process because the formation of

ettringite is inhibited.

In India, there is an increasing tendency to use synthetic gypsum instead of natural

gypsum during cement production. This is because of the better quality and higher

SO3 content of synthetic gypsum. However, the rate of solubility of synthetic gypsum,

and its effects on the initial adsorption and consequent fluidity, are not clear from

literature and have to be investigated. Moreover, synthetic gypsum could contain

soluble phosphates (in the range of 0.08-0.18 %) that increase the setting time from 3

to 10 hours (Hewlett, 2004). Such retardation can be eliminated if the phosphogypsum

is neutralized with lime, and content of water soluble phosphates and fluorine is

reduced to below 0.02% (Hewlett 2004; Diop et al., 2005).

2.3.3.3 Effect of Soluble Alkalis

The alkali compounds have an active role in the adsorption and consequently in the

interaction. It has been observed that a higher alkali content in the aqueous solution of

cement paste increases the solubility of sulphate ions and hence decreases the loss of

fluidity with SNF due to lower initial adsorption of the superplasticizer (Dodson and

Hayden, 1989; Chandra and Bjornstrom, 2002b). Note that the soluble alkali content

is not necessarily related to the total alkali content of the cement. If cement has low

soluble alkali content, the adsorption of SNF will be more and that remaining in

solution will be less resulting in loss of fluidity (Dodson and Hayden, 1989;

35
Nkinamubanzi et al., 2000). This is because the superplasticizers act as sulphate ion

providers and interact with C3A instead of dispersing it. Consequently, cements with

high alkali contents are expected to have good slump retention and to be more robust

than cements with low alkali contents, when used in combination with SNF

(Nkinamubanzi and Aïtcin, 2004).

The problem of inadequate fluidity for low alkali cement-SNF combinations can be

overcome by adding soluble alkalis, such as Na2SO4 (Jiang et al., 1999; Li et al.,

2003; Spiratos et al., 2003); Jiang et al. (1999) reports that 0.4-0.5% Na2O soluble

equivalent is the optimum soluble alkali content in the cement to maximize initial

fluidity and to minimize the loss of fluidity. However, the addition of sodium sulphate

could increase the incompatibility problem with some polysulphonates, as it alters the

formation of the ettringite (Li et al., 2003).

In the case of the PCE, the excess of alkali sulphates results in the lowering of fluidity

due to the shrinking of the side chains, thereby decreasing the steric repulsion

(Hanehara and Yamada, 1999). Also, efflorescence problems are observed when

naphthalene and melamine based superplasticizers are used with cements having high

alkali oxide contents (Na2O + K2O >0.75%) (Akman and Cavdar, 1999).

2.3.3.4 Effect of C3A

The interaction of the superplasticizer with the C3A phase of the cement has the prime

influence on the rheological properties of the fresh mix due to its higher positive zeta

potential and reactivity compared to other cement compounds. The C3A content or

more specifically SO3/C3A ratio and the reactivity of C3A controls the adsorption of

the superplasticizer, which in turn affects the early hydration behaviour and

workability of cement paste (Aïtcin et al., 2001). The calcium sulphate reactivity

36
determines the availability of soluble sulphates whereas the C3A reactivity determines

the rate of consumption of these ions. Therefore, the amount of soluble sulphates

along with the reactivity of C3A influences the hydration reactions of the aluminate

phases resulting in the setting of cement paste as either normal, quick, flash or false

set as described by several authors (Mehta and Monteiro, 2005; Spiratos, et al., 2003).

The effect of the superplasticizer on C3A hydration depends on the type of

superplasticizer; the adsorption of SNF and SMF based superplasticizers retard C3A

hydration and delay the conversion of ettringite to monosulphonate. This is because

C3A and C4AF surfaces rapidly adsorb large amounts of SNF and SMF. Also, in the

presence of sulphate ions, the SNF forms an organo-mineral compound with C3A and

alters the morphology of the hydration products. However, high molecular weight

SNF shows less retardation of C3A than that with lower molecular weight (Rixom and

Mailvaganam, 1999).

The interaction of PCE with C3A depends on the form of C3A, cubic or orthorhombic,

with the orthorhombic form being more reactive than cubic form. Hence, a high cubic

C3A content results in low flow retention as the PCE will be adsorbed first on the C3A

surface and the ettringite will be formed later. Later, the admixture will be covered by

the hydrated products and will not be available for further flow retention (Magarotto

et al., 2003).

In the case of cements with less C3A, adsorption of the superplasticizer will be more

on the C3S and C2S, preventing their hydration, resulting in the reduction of early

strength (Roberts, 1995). Also, crystallization of gypsum needles and consequent false

set could occur if the C3A content is not enough to consume the dissolved sulphates

(Ramachandran, 2002). The C3A phase in cement has very high surface area

37
compared to its hydration product, ettringite, and consequently adsorbs large amounts

of admixture resulting in loss of fluidity. Hence, the delayed addition of

superplasticizer is advantageous as the hydration of C3A takes place and ettringite is

formed before the superplasticizer is added (Roberts, 1995).

2.3.3.5 Other Factors Affecting the Cement-Superplasticizer Interaction

There are several other factors that can influence the effectiveness of the

superplasticizer and the interaction between the superplasticizer and cement, in

general. Some important factors that have not been discussed earlier are explained in

the following paragraphs.

The specific surface area influences the effectiveness of the superplasticizer since it

increases the water demand, as well as the amount of superplasticizer for a given

workability. Hence, the amount of superplasticizer adsorbed depends on the fineness,

with finer cements leading to higher superplasticizer dosages (Aïtcin, 1998; Jolicoeur

and Simard, 1998; Erdogdu, 2000; Vikan et al., 2007).

As is well known, the w/c ratio affects the performance of superplasticizer in terms of

the flow of the cement paste and workability of concrete. Moreover, most of the

incompatibility problems occur at low w/c, where the lower interparticle separation

between the cement particles makes the system to be more sensitive to the loss of

water by evaporation or hydration reactions resulting in higher loss of fluidity with

time (Collepardi, 1998).

The placing of concrete in different environmental conditions also affects the

superplasticizer performance. Temperature is the major environmental factor affecting

cement-superplasticizer interaction. Generally low temperatures decrease the fluidity,

which cannot be compensated with the addition of superplasticizer beyond the

38
saturation dosage (Gettu et al., 1997; Roncero et al., 1999). On the other hand, high

temperatures increase superplasticizer adsorption, which increases the initial fluidity

and increases the superplasticizer demand for flow retention over a certain range of

temperatures (Roncero et al., 1999; Greisser, 2002). Also, the temperature increase

causes an increase in the reactivity of C3A, which causes a more rapid loss of fluidity

and formation of higher ettringite contents with finer morphology in the presence of

superplasticizer (Greisser, 2002; Spiratos et al., 2003).

The effect of temperature on mixes with PCE based superplasticizers is dependent on

its chemical structure (e.g., the carboxylates have the highest fluidity loss compared to

other functional groups) (Felekoglu and Sarikahya, 2007; Grzeszczyk and Sudol,

2005), with the increase in temperature leading to a lower increase of flow in the paste

and lower flow loss for longer polyethylene oxide graft chains (Nawa et al., 2000).

This is attributed to the higher adsorption of the graft chains with an increase in

temperature, as a result of which the steric repulsion of the longer graft chains is less

affected.

Roncero et al. (1999), Roncero (2000), and Fernandez-Altable and Casanova (2006)

have shown that though the fluidity increases with the temperature, it has no practical

effect on the saturation dosage of superplasticizer, which is attributed to the

dependence of the saturation on the available surface area of the cement particles.

2.3.4 Effect of Superplasticizer on the Morphology of Hardened Cement Paste

The morphology of the hardened cement paste could be affected by the alteration of

the cement hydration reactions by the retarding effect of the superplasticizers as well

as the intercalations with the early hydration products, as described in the previous

sections. The polysulphonates, can combine with the interstitial phase of the cement

39
(C3A+C4AF) to form an organo mineral composite resembling ettringite (Aïtcin et al.,

2001). They could also modify the morphology of ettringite produced during

hydration; i.e., ettringite crystallizes in small and thick crystals rather than

conventional needle shaped crystals (Aïtcin et al., 1994; Hanna et al., 2000; Prince et

al., 2002). The addition of superplasticizer also modifies the morphology of the

portlandite; i.e., the size of portlandite crystals decrease and also their form changes

from block-like to thin plates (Grabiec, 2004).

In addition to modifying the morphology, the delay in the release of SO42- (due to the

presence of anhydrite) results in the formation of monosulphoaluminate, first and then

ettringite is formed (Prince et al., 2003). The addition of lignosulphonates further

reduces the solubility of anhydrite due to the reduction in the rate of diffusion of ions

at solid-liquid interphase, and further promotes this phase transformation (Prince et

al., 2003).

In general, the change in morphology due to the incorporation of superplasticizer

improves the rheological properties by yielding smaller hydrate particles and

preventing hydration products from bridging neighbouring cement particles. This

results in substantial reduction in the yield stress of paste and increase in fluidity.

There is also a difference in porosity and pore size distribution of superplasticized

concrete compared to normal concrete. Cements with high C3A and low SO3 content

have larger pores that decrease in size with higher SO3 contents whereas the pore size

in low C3A cements is independent of the SO3 content. Generally, higher numbers of

smaller pores are produced in superplasticized mixtures (Roncero, 2000; Khatib and

Mangat, 1999). This has been attributed to better silicate polymerization leading to the

refinement of pores and consequent reduction in porosity (Puertas et al., 2005).

40
2.3.5 Effect of Superplasticizer on Properties of Concrete

As intended, the incorporation of superplasticizers increases the workability of

concrete that permits the use of comparatively lower w/c leading to higher strengths.

In addition to strength, the properties of concrete that are affected include slump, air

content, bleeding, segregation and setting time in the fresh state, and properties like

permeability, abrasion resistance and durability in the hardened state. The type of

superplasticizers and cement can cause variations in these properties and in certain

cases the incompatibility between the two leads to adverse effects like loss of fluidity,

retardation in setting, air entrainment and reduction in strength gain, as explained in

the previous sections. Aïtcin et al. (2001) observed that cements with a C3A content of

6% to 10%, a soluble alkali content of 0.4 % to 0.5% give cement-polysulphonate

combinations that are both compatible and robust.

Concrete with high workability containing a superplasticizer generally shows more

rapid slump loss than that with the same workability and without superplasticizer.

This could be attributed to the loss of consistency due to the low w/c during the

dormant stage and the increased adsorption of the superplasticizers by the aluminate

phase (Ramachandran et al., 1998). In addition, the higher slump loss is related with

the C3S phase of cement also; i.e., the hydrate coatings restore charge interactions

between the C3S grains, thereby restoring the tendencies for flocculation

(Ramachandran, 2002). Redosing of the superplasticizer neutralizes these charges and

restores the fluidity, but continued hydration will lead eventually to slump loss.

In a superplasticized concrete, generally no segregation or bleeding is expected to

occur at the saturation dosage of the superplasticizer. When redosing is used to

control the loss of slump, however, bleeding and segregation may occur if the second

41
dosage is too high; overdosage of superplasticizer generally promotes segregation and

formation of a white deposit containing lime, calcium sulphate and calcium carbonate

on the top surface when hardened (Ramachandran, 2002).

When superplasticizers are added to concrete at the appropriate dosage, the 28-day

compressive strength of superplasticized concrete is often equal or greater than the

corresponding strength of the reference concrete (Rixom and Mailvaganam, 1999;

Ramachandran, 2002). This is attributed to the reduction in w/c as well as the higher

workability and better compaction achieved due to the incorporation of

superplasticizers Dhir and Yap (1983) also reports that even though there could be

some initial retardation, superplasticizers produce compressive strengths of the same

order as the normal concrete at 3, 7 and 28-days. This is because, normally, the C3A

and C4AF phases rapidly adsorb large amounts of SNF and SMF, so that only small

amounts of SMF and SNF are available with the aqueous phase to retard the hydration

of C3S.

Nevertheless, there are some reports of strength reduction while using naphthalene

based superplasticizers and melamine based superplasticizers, especially with cements

having low C3A content and high C3S and C2S contents (Ramezanianpour et al.,

1995). This may be either due to higher adsorption on the C3S and C2S phases or

gypsum crystallization and consequent false set in the absence of adequate C3A, as

reported earlier (Roberts, 1995; Ramachandran, 2002)

The creep and shrinkage of superplasticized concrete is comparable to or is less than

that of reference concrete (e.g., Seung-Bum, 1999), although there are several

exceptions. For the same amount of moisture loss, the superplasticized concrete could

exhibit larger shrinkage which may be related to the larger dispersion of cement and

42
cement hydrates (Ramachandran, 2002). Dhir and Yap (1983), and Gettu et al. (2002)

have reported that drying shrinkage is higher for superplasticized concrete with high

workability; with SNF based superplasticizers leading to higher total shrinkage than

PCE based products whereas SMF based shows lower shrinkage than PCE based

superplasticizers. This could be attributed to the refinement of pore structure to more

number of fine pores (Roncero, 2000; Gettu et al., 2002; Roncero et al., 2002).

2.3.6 Admixture – Admixture Interaction

2.3.6.1 Cement-Superplasticizer Interaction in the Presence of Mineral


Admixtures

Supplementary cementitious materials (or mineral admixtures) are generally

incorporated in concrete as a replacement for cement to reduce the heat of hydration

and to enhance strength and durability. It is usually reported that the addition of

mineral admixtures reduces the workability due to the higher surface area and

consequently increases the water and superplasticizer demand (e.g., Agullo et al.,

1999). Nevertheless, there could be some workability enhancement due to the addition

of fly ash and silica fume, in low volumes, which is generally attributed to the

spherical shape of the particles and consequent reduction in the interparticle friction

(Ferraris et al., 2001).

The incorporation of fly ash could also affect the hydration process of cement due to

the additional aluminates (of the fly ash) that suppress the silicate reaction. This

retarding effect is worsened due to the lack of sulphates in the presence of

superplasticizers (Roberts, 1995).

43
2.3.6.2 Influence of Other Chemical Admixtures on Cement-Superplasticizer
Interaction

The other chemical admixtures generally used in concrete include viscosity modifying

admixtures (VMA), air entraining agents, damp proofing compounds and shrinkage

reducing admixtures. A viscosity modifying admixture is usually incorporated with

superplasticizers to produce flowing/self compacting concrete. The interaction of the

superplasticizer with the VMA depends on the mode of action by which the latter

provides higher viscosity in the concrete. Most VMAs tend to increase the viscosity

by forming gel/cement slurries rather than increasing the viscosity of water in the mix.

The VMA molecules could compete with the superplasticizer for active sites on the

cement particles, resulting in lower superplasticizer adsorption (Mailvaganam, 1999;

Bedard and Mailvaganam, 2006).

The addition of a superplasticizer modifies the relationship between the air content

and the spacing factor (air bubble spacing), and sometimes significantly destabilizes

the air-void system (Bedard and Mailvaganam, 2005) caused by the incorporation of

an air-entraining agent. For example, when a lignosulphonate based water reducer is

added to concrete containing an air entraining agent, a substantial increase in air

content occurs and the specific surface of air bubbles are reduced.

The combination of a shrinkage reducing admixture (SRA) with a superplasticizer

reduces the efficiency of the SRA in reducing shrinkage. Also, the combination

increases the retardation in setting more than the superplasticizer alone (Gopinath and

Gettu, 2008).

2.4 Need for the present study

The current level of knowledge worldwide about the factors affecting the cement-

superplasticizer interaction is described in this chapter. Specific issues of

44
incompatibility have also been addressed, namely, loss of workability, alteration of

setting behaviour, reduced rates of strength gain and change in the microstructure.

There are contradictions regarding the influence of molecular weight and the side

chain lengths of PCE on the adsorption behaviour, which have also been brought out.

In the Indian context, cement standards in our country are not very stringent, and

enable manufacturers to adjust their product in many different ways. For example,

while the minimum fineness is specified for different grades of cement, there is no

control on the maximum values. Similarly, wide ranges in the chemical composition

are acceptable. This could result in significant variability in the cement properties.

Also, the alkali content, which has a major influence on cement-superplasticizer

interaction, is not considered as a characteristic of the chemical composition for a

particular grade of cement.

From the viewpoint of the use of superplasticizers, there is insufficient knowledge

among users regarding the limitations of different types of chemicals. Users, who are

unaware of compatibility issues, often suffer when the supply of cement and/or

admixture has to be changed during the course of a project. Problems arising from

compatibility issues are often mistaken for problems with the concrete mixture design

because of the lack of information about the subject amongst practicing engineers.

Admixture manufacturers try to overcome the problem by formulating project-specific

chemicals, which is obviously, only a short term solution.

Cement-superplasticizer interaction in concrete is due to a complex blend of chemical

and physical mechanisms that are interdependent. The complicated nature of the

problem prevents the development of simple solutions to address field related issues

of the application of superplasticizers. Hence, a detailed study on the cement-

45
admixture interaction is required for optimizing different types of admixtures in a

cementitious system.

Studies on cement-superplasticizer interactions in India have been limited to the

workability evaluation of concretes containing these chemicals. There have not been

any investigations to understand the physico-chemical nature of this interaction. Thus,

the results from such studies are not broad-based, i.e. they apply to a small group of

cements and/or chemical admixtures. There is a distinct need for a study of the

interaction of Indian cements and admixtures, as well as to select an appropriate

dosage of the superplasticizer. Moreover, the wide range of cements used, varying

transportation durations and climatic conditions necessitate fundamental studies that

can explain the mechanisms of interaction and help establish methods for identifying

incompatibilities in practical situations.

Simple methodologies are required to identify these incompatibilities due to such

interactions and to further understand the fundamental nature of admixture behaviour

in cement-based systems. The understanding of these interactions should be both at

the applications scale (for example, studying flowability and retention of workability

in pastes and concretes, retardation in setting), as well as at the micromolecular scale,

where some insight can be obtained into the physico-chemical interactions between

cement particles and superplasticizer molecules.

46
CHAPTER 3

STUDY OF THE FLOW BEHAVIOUR OF


SUPERPLASTICIZED CEMENT PASTE

3.1 GENERAL

A brief description about the cement hydration reactions and the influence of

superplasticizers on the hydration reactions, morphology of hydrated products and the

properties of superplasticized concrete has been presented in the previous chapter. In

addition to providing good workability, the superplasticizer also influences the slump

retention and setting process. The interaction between the cement and superplasticizer

governs the flow behaviour of concrete. To study this interaction, it is proposed to

analyse the flow of cement paste with simple methods like Marsh cone and mini-slump

tests, and more fundamental tests like those with a viscometer.

The basic objective of this chapter is to formulate the methodology for studying the

fluidity of superplasticized cement pastes that will be used in the more detailed studies in

the rest of the thesis. The specific objectives are

• To select a suitable mixing method for preparing the superplasticized cement

paste that is representative of the paste phase in concrete.

• To validate the suitability of simple test methods, such as the Marsh cone and

mini slump tests, for studying the flow behaviour

• To understand the significance of results from simple methods with regard to

rheological parameters.
In order to meet the above objectives, the influence of the mixing method on the fluidity

of superplasticized cement paste is studied through the Marsh cone and mini-slump tests.

Subsequently, the influence of type and dosage of superplasticizer on the flow behaviour

of the paste is studied. Finally, the correlations between the results obtained with the

simple test methods and those obtained from viscometric studies are presented which

helps to select an appropriate methodology to study the fluidity of cement paste and to

understand the results from empirical tests better.

3.2 MATERIALS USED IN THE STUDY

3.2.1 Cement

Ordinary 53 grade portland cement (ASTM Type 1 Cement) (denoted as Cement 1)

satisfying the requirements of IS 12269 (2004) was used in the work presented in this

chapter. The chemical properties are given in Table 3.1, as determined by the National

Test House, Chennai, for the sample used.

The compounds present in the cement can be quantified using the results of the chemical

analysis of the clinker corresponding to the cement and the Bogue equation. Accordingly,

the clinker was obtained from the cement manufacturer and tested in Holcim Roodeport

Laboratory in South Africa; the results of the analysis are also given in Table 3.1. These

data were used to determine the proportions of the compounds present in the cement (see

Table 3.2).

From the test results (Tables 3.1 and 3.2), it is confirmed that the cement used can be

considered as a low alkali cement since the (Na2O)eq is less than 0.6% (Hewlett, 2004);

that the SO3 content is within the normal range (i.e., 2–3.5%) for ordinary Portland 53

grade cement and, therefore, is not expected to affect the superplasticizer adsorption

48
significantly (Chandra and Bjornstrom, 2002a); and that the C3A content is in the normal

range (i.e., 4-8%).

Table 3.1 Chemical properties of the cement and the corresponding clinker

Characteristics Quantity
(% by mass)
Cement Clinker
Calcium oxide (CaO) 60.81 63.10
Silica (SiO2) 19.50 20.80
Alumina (Al2O3) 4.12 5.80
Iron oxide (Fe2O3) 6.06 5.30
Magnesia (MgO) 1.52 0.80
Sulphur anhydrite (SO32-) 2.48 0.68
Insoluble residue 1.51 -
Total loss on ignition 3.41 0.40
Total chloride content (Cl-) 0.01 -
Alkali content: Na2O 0.05 -
K2O 0.28 -
(Na2O)eq 0.23 0.20
Lime saturation factor 0.93 0.92
Ratio of alumina to iron oxide 0.68 1.09

Table 3.2 Bogue composition of the clinker

Compound Composition
(%)
C3S 51.7
C2S 20.8
C3A 6.5
C4AF 16.1

In addition to the chemical properties, it is necessary to determine certain physical

properties of the cement used for characterizing its behaviour. In this work, the Blaine

specific surface area, water demand for a paste of standard consistency, and the initial

49
and final setting times were determined experimentally, as detailed in the following

sections.

3.2.1.1 Determination of the Water Demand for Standard Consistency of Cement

The consistency of the cement paste is determined as per ASTM C 187 (2004)and IS

4031: Part 4: (2005), where the quantity of water found that produces a cement paste of

standard consistency, defined as that which permits the Vicat plunger to penetrate upto 5

to 7 mm from the bottom of the Vicat mould. Cement pastes with varying w/c are

prepared (by hand mixing) and the penetration is checked for each of them. The amount

of water in the paste that satisfies the above mentioned condition is expressed as a

percentage by mass of dry cement, and taken as the water demand of the cement. This

can be considered as the minimum w/c needed for the paste to be homogeneously, and

uniformly mixed and moulded. The water demand, obtained as an average result of 3

trials was 31% (see Table 3.3).

Table 3.3. Physical properties of cement 1 (C1)

Property Value (mean and standard deviation)

Water demand for normal consistency of


cement paste 31% (± 1.3%)
Initial setting time 106 minutes (± 7 minutes)
Final setting time 190 minutes(± 6 minutes)
Blaine specific surface area 332 m2/kg (± 10 m2/kg)

3.2.1.2 Determination of Initial and Final Setting Times

The initial and final setting times are determined as per IS 4031: Part 5 (2005). The

cement paste is prepared with 0.85 times the water required to give a paste of standard

consistency (see previous section). It is filled in the Vicat mould and the corresponding

needle is allowed to penetrate through the paste. The period elapsing between the time

50
when water is first added to the cement and the time at which the penetration of the

needle is 35.0 ± 0.5 mm is taken as the initial setting time. Note that this is similar to

procedure of ASTM C-191 (2001a), except that the paste used (i.e., same w/c) is of

standard/normal consistency, and the penetration for the initial setting time is 25 mm.

The test is continued with a standard needle with an annular attachment for determining

the final setting time. The period elapsing between the time when water is added to the

cement and the time at which the needle makes an impression on the surface of the test

block while the attachment fails to do so is taken as the final setting time.

The values obtained for the setting times are given in Table 3.3 as averages from three

trials. As the initial setting time is more than 30 minutes and the final setting time is less

than 10 hours, the IS code requirements for the setting times are satisfied. It should be

noted, however, that the difference between the initial and final setting times is only

about 90 minutes, which is relatively low.

3.2.1.3 Determination of the Fineness of Cement

The fineness of cement has been determined by the Blaine air permeability method as per

IS 4031: Part 2 (2005) and ASTM C 204 (2007). The specific surface area obtained

(332 m2/kg) is within the limits given for a 53 grade OPC.

3.2.2 Superplasticizers

Superplasticizers for this study were selected from the four different families of products

available (i.e., lignosulphonates, naphthalenes, melamines and polycarboxylates) in order

to obtain representative results that would help understand the effect of the different types

on the flow behaviour of cement paste. Table 3.4 gives the details of the commercial

products used; the products are denoted as the abbreviation of the type followed by a

51
number. The values of the solid content, pH, density were obtained in the laboratory as

per IS 9103 (2004), and cross-checked with data from the corresponding supplier. It can

be seen that the active part of superplasticizer varies from 32% to 44 % for different

families as well as for the same family. The recommended dosage of the superplasticizers

is also given in the Table 3.4. The superplasticizer dosages given in this work are all in

terms of the solid content; the water contained in the superplasticizer has been accounted

for in the water content of the mixes.

Table 3.4 Properties of superplasticizers

Density Solid Recommended


Chemical type
Designation content pH dosages
kg/litre (%) (sp/c %)
LS-C1 Lignosulphonate 1.18 38 7.5 0.18-0.27

Polynaphthalene
SNF-S1 1.09 37 6.7 0.2-0.8
sulphonate
Polynaphthalene
SNF-D1 1.16 32 6.8 0.22-0.74
sulphonate
Polynaphthalene
SNF-D2 1.25 44 7.8 0.44-0.66
sulphonate
Polynaphthalene
SNF-C1 1.20 40 7.4 0.2-0.3
sulphonate
SMF-S1 Polymelamine
1.30 40 8.5 0.2-0.68
sulphonate
Polymelamine
SMF-C1 1.28 38 7.0 0.19-0.3
sulphonate
Polycarboxylate
PCE-S1 1.09 38 7.5 0.08-0.33
ether
Polycarboxylate 33
PCE-D1 1.10 7.4 0.07-0.3
ether

3.3 TEST METHODS USED FOR STUDYING THE FLOW BEHAVIOUR OF


THE CEMENT PASTE

Due to the important functions of superplasticizers in concrete, the selection and dosage

criteria are fundamental for determining the optimum composition of concrete. From the

52
practical point of view, it is better to make this choice using the actual concrete at the

construction site under the local conditions. However, tests of concrete imply significant

labour, material and time. Besides, such tests do not reflect the fundamental action of the

superplasticizer with regard to physical and chemical interactions. Therefore, several

methods based on tests of paste and mortar have been developed for determining the

optimum superplasticizer dosage in concrete.

The Marsh cone and mini slump tests are simple empirical methods to characterize the

flow behaviour of cement paste whereas rheological tests with a viscometer provide more

fundamental information about the material but require more complicated equipment and

skilled operators. The fundamentals of these three methodologies, and the procedures

adopted in this work are detailed in the following sections.

3.3.1 Marsh Cone Test

The Marsh cone (Figure 3.1) test has been used previously to evaluate the relative fluidity

and the saturation superplasticizer dosage in cement pastes and mortars (de Larrard et al.,

1998; Aitcin, 1998; Khayat and Yahia,1998; Agullo et al., 1999; Roussel and Roy, 2005;

Royy and Roussel, 2005). In the present study, a metal cone (as per the guidelines of

European standards EN 445, French standards P 18-358, which is similar to ASTM C 939

(1987), with a nozzle of diameter 8 mm was employed. An initial volume of 1000 ml of

paste was poured into the cone and the time required for 500 ml of it to flow out was

measured. The test gives the fluidity of the paste in terms of the flow time; higher the

flow time, lower is the fluidity of the paste. The saturation point is the dosage beyond

which further addition of superplasticizer does not increase fluidity significantly but can

53
produce segregation (Agullo et al., 1999); the saturation dosage can be taken as the

optimum superplasticizer dosage for a given cement paste.

R, H - Radius and length of nozzle


R1, H1 - Radius of the free surface and height of fluid
in the cone at the initial moment
Rt, Ht - Radius of the free surface and height of fluid
in the cone at any time t
R2, H2 - Radius of the free surface and height of fluid
in the cone at the final moment
φ – Angle between the axis and the generator of the
cone

Fig. 3.1 Geometry and dimensions (in millimetres) of the Marsh cone

Gomes et al. (2001) proposed a method for the objective determination of the saturation

dosage based on the Marsh cone flow time curve of pastes, which is illustrated in

Figure 3.2. In this method, the internal angle (α) corresponding to each data point is

calculated and the superplasticizer dosage corresponding to an internal angle of 140°+10°

is taken as the saturation dosage. Interpolation is used to determine the dosage when there

are no data points corresponding to that range of angles. This criterion was proposed

based on about 200 tests on superplasticized cement pastes (Gomes, 2002).

54
1.8 SMF-S1

1.7

log (flow time , sec)


1.6
1.5

1.4 saturation point corresponding to 140


0

0 0
1.3 140 +10

1.2

1.1

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %

Fig. 3.2. Flow time curve showing the determination


of optimum superplasticizer dosage
(according to the method of Gomes et al., 2001)

The Marsh cone test (see Figure 3.3) can also be used to study the loss of fluidity with

time, the fluidity of cement mortar, optimization of cement paste with mineral admixtures

and incompatability between cement and superplasticizer (Agullo et al., 1999).

Fig. 3.3 Marsh cone test showing continuous and discontinuous flow

55
3.3.2 Mini-slump Test

The mini-slump test developed by Kantro (1980) has also been used to study the flow

behavior of superplasticized cement paste (Aitcin, 1998; Khayat and Yahia,1998; Gomes,

2002; Roussel, 2006; Tregger et al., 2008). In this test, a mould in the shape of a

truncated cone, with dimensions proportional to the Abram’s cone, is filled with cement

paste and the spread diameter is measured after lifting the mould. Additionally, the time

taken for the paste to reach a diameter of 115 mm was also determined here. Also, visual

examination helps evaluate the bleeding and segregation of the paste.

3.2 19.0 6.4

15.9
22.2

57.0
38.1

3.0
38.1
All dimensions are in millimetres

Fig. 3.4 Mini-slump cone

(a) Filling of the mini-slump (b) Spread of the Paste

Fig. 3.5 Mini-slump test

56
3.3.3 Viscometer Test

Viscometer tests on cement paste can be used to study the effect of the changes in

cement, type and dosage of admixtures, such as superplasticizers, on paste characteristics

like yield stress and plastic viscosity (Asaga and Roy,1980; Banfill, 1981; Masood and

Agarwal, 1994; Beaupre and Mindess, 1998; Claisse et al., 1999; Cyr et al., 2000; Park et

al., 2005). The basic principle is to apply different shear rates to the paste in a viscometer

and measure the corresponding shear stresses. Generally, a loading-unloading cycle is

applied to the paste, preceded by some pre-shearing, and the response during unloading is

used to determine the rheological parameters (Goisis et al., 2003; Papo and Piani, 2004).

Cement paste normally exhibits shear-thinning behaviour, where the slope of the shear

stress versus shear strain rate curve decreases with an increase in the shear rate. Such

shear-thinning response is attributed to the structural breakdown and rebuilding that takes

place in the paste (Tattersall and Banfill, 1983) with the former mechanism

predominating (Lapasin et al., 1983). There are, however, notable exceptions; e.g., dilute

superplasticized pastes may exhibit Newtonian behaviour while an overdosage of

superplasticizers may result in shear thickening (strain-hardening) that could be attributed

to the agglomeration of excess superplasticizer in the aqueous solution of the paste or

segregation(Cyr et al., 2000; Papo and Piani, 2004; Lootens et al., 2004).

In the present study, a Brookfield HA DV II +Pro viscometer was used with a coaxial

cylinder setup (inner cylinder radius = 16.77 mm, gap width = 1.14 mm) as shown in

Figure 3.6 and Rheocalc software for programming the test. The details of the coaxial

cylinder set up are shown in Figure 3.7. The rheological behaviour depends on the type of

spindle, shear rate, the gap between the cylinders and environmental conditions. The

57
basic principle of the experiment is to apply a known shear rate to the fluid through the

spindle and measure the corresponding shear stress produced.

Fig. 3.6 Brookfield HA DV II +Pro viscometer (Asphalt Laboratory, IITM)

(a) Co-axial cylinder (b) SC4 21 - Spindle (c) HT-2- Sample chamber

Fig. 3.7 Details of co-axial cylinder setup

Three loading-unloading cycles were imposed by increasing, and later decreasing, the

shear rate from 23 to 163 s-1 in seven steps. The shear history of rheological testing for

one hysteresis cycle is shown in Figure 3.8.

58
180
160
140

Shear rate (1/sec)


120
100
80
60
40
20
0
0 20 40 60 80 100 120 140
Time (sec)

Fig. 3.8 Shear history of rheological testing

The typical response is shown in Figure 3.9, where the first two cycles are considered as

pre-shearing and the downward curve of the third cycle is used for determining the

rheological parameters. The pre-shearing is needed for the structural breakdown of

cement paste sample and to create uniform conditions before testing.

35
SNF-S1
Dosage-0.4%
30
Shear stress (Pa)

25

20

First cycle
15 Second cycle
Third cycle

10
20 40 60 80 100 120 140 160 180
Shear rate (1/sec)

Fig. 3.9. Typical graph showing hysteresis cycles for a superplasticized cement paste

59
The shear stress-strain behavior of cement paste has been represented using several

models (Atzeni et al., 1985; Nehdi and Rahman, 2004), the simplest of which is the

Bingham model:


τ =τ0 + μp γ (3.1)


where τ is the shear stress in (Pa), γ is the shear strain rate (1/sec), μ p is the plastic

viscosity (Pa s) and τ 0 is the yield shear stress (Pa). The Bingham model is generally

fitted to the experimental data (during unloading), as illustrated in Figures 3.10(a-c), to

obtain the yield shear stress and the plastic viscosity of the cement paste. This appears to

give a satisfactory representation of the rheological behaviour in most cases. Even slight

shear thickening response can lead to a negative yield stress when the Bingham model is

fitted to the experimental data. When the τ 0 -value is negative, as in Figure 3.10c, the

data are normally re-fitted by forcing the condition of τ 0 =0 (i.e., Newtonian model). For

these reasons, a more versatile model, namely the Herschel-Bulkley model, has been used

by several researchers (de Larrard et al., 1998):


τ =τ0 + kγ n
(3.2)


where τ is the shear stress in (Pa), γ is the shear strain rate (1/sec), n is the power index,

k is the consistency (Pa sn), and τ 0 is the yield stress (Pa). Since the consistency and

power index are interdependent, there is no single term that represents the plastic

viscosity of the material.

60
18
Bingham model
Herschel Bulkley model
16

14 SNF-C1 0.1%
Shear stress (Pa)
w/c 0.35

12

10 Bingham model- τ0= 6.724Pa


2
μ=0.057 Pa s, R = 0.99
8 Herschel Bulkley model-τ0=5.122
n 2
k=0.269 Pa s , n= 0.722, R = 0.99
6
0 40 80 120 160 200
Shear rate (1/sec)

(a) Model parameters at less than saturation dosage

22
Bingham model
20 Herschel Bulkley model

18
SMF-S1 0.2%
Shear stress (Pa)

16 w/c 0.35
14

12
Bingham model- τ0=6.895 Pa
10 2
μ=0.0759 Pa s, R =0.99
8 Herschel Bulkley model-τ0=6.427 Pa
n 2
k= 0.117Pa s , n= 0.92, R =0.99
6
0 40 80 120 160 200
Shear rate (1/sec)

(b) Model parameters at saturation dosage

61
7
Bingham model
Herschel Bulkley model
6

5 PCE-S1 1%
w/c 0.35

Shear stress (Pa)


4

2 Bingham model- τ0=-0.264 Pa


2
μ=0.0358 Pa s, R = 0.99
1 Herschel Bulkley model- τ0= 0.388 Pa
n 2
k= 0.0057 Pa s , n= 1.344, R = 0.99
0
0 40 80 120 160 200
Shear rate( 1/sec)

(c) Model parameters at dosage much more than saturation

Figure 3.10 Fits of experimental data with the Bingham and Herschel-Bulkley
models at (a) less than saturation (b) at saturation
(c) dosage much more than saturation

3.4 SELECTION OF MIXING METHOD

The purpose of concrete mixing is to achieve a uniform mixture of all material

components and facilitate hydration. The type of mixer used has an influence on the

mixing efficiency (i.e., degree of dispersion of cement particles) depending on whether it

is intensive or extensive (Yang and Jennings, 1995). Mixing is especially important for

high performance concrete (HPC) of lower binder content. The type and efficiency of

mixer will influence the mixing time needed, the superplasticizer dosage required, and

the quality of HPC produced (Chang and Peng, 2001). Previous studies of Yang and

Jennings (1995), and Diamond (2005) have concentrated on the microstructure and

rheology of pastes mixed in different ways. They have shown that intensive mixing leads

to more homogeneity and better dispersion of cement particles. Also, the viscosity of the

62
paste decreases with intensive mixing, reflecting the effectiveness of such methods

(Williams et al., 1999; Ferraris, 2001).

Studies conducted by Diamond (2005) show that prolonged mixing can induce higher

stiffness in the mix. Therefore, it is better to have an intensive mixing procedure of short

but adequate duration instead of a prolonged mixing time.

Hence, it is important to understand the effect of different mixing methods on the fluidity

of superplasticized cement paste as well as to select the type of mixing method which

leads to a paste that is representative of concrete. In the present work, two simple

methods, the Marsh cone and mini-slump tests, are used for studying the fluidity of

cement paste made with different mixing techniques.

3.4.1 Materials Used in the Selection of Mixing Method

In the present study, tests were conducted on cement paste with water/cement ratio of

0.35. Birla Super 53 grade ordinary portland cement (OPC) and a polycarboxylic ether

based superplasticizer (PCE-D1) was used. The recommended dosage of superplasticizer

is 0.5-1.6 litres per 100 kg of cement. Distilled water was used for mixing the cement

paste. Superplasticizer dosage, denoted as sp/c, is expressed as the ratio of solid content

of superplasticizer to cement content by weight. All the test materials were kept in the

environmental chamber at a temperature of 27 ºC for 24 hours prior to testing.

3.4.2 Experimental Details

The first paste was mixed with a Hobart mixer (1/6 hp) with a B-flat beater (Figure 3.11a)

with low speed (with shaft speed of 139 rpm and planetary speed of 61 rpm), the second

paste with a Hobart mixer with a D-wire whip (Figure 3.11 b) at medium speed (with

shaft speed of 285 rpm and planetary speed of 125 rpm) and third with a high shear rate

63
kitchen blender (550 watts, 16000-18000 rpm). The fourth paste was made by grinder

(1/4 hp, 300 rpm) mixing and fifth by hand mixing. The mixing time of 5 minutes was

selected to minimize any effects due to incomplete mixing with respect to the capability

of the machine. After determining the proportions of constituents, the mixing was

performed in the following sequence for all the cases. The cement and 70 % of the water

required were mixed together in the mixer for one minute, and after that superplasticizer

and the remaining water were added to the cement paste. This sequence of addition is

preferred because the delayed addition of superplasticizer reduces the amount of

superplasticizer required (see Chapter 2). In this case, the water content of the

superplasticizer was deducted from the water added. The paste was mixed for 2 minutes

at the same speed. The sides of mixer bowl were scraped for 15-30 seconds. Again the

paste was mixed for 2 minutes.

Fig. 3.11 (a ) Hobart mixer with B-flat beater (b) Hobart mixer with D-wire whip

3.4.3 Ball Milling

The study for selecting the mixing method is conducted on cement paste instead of

concrete because of the reasons explained in sections 3.3 and 3.4. However to ensure that

the paste is representative of the paste within the concrete, “ball milling” was done for

64
simulating the concrete mixing through the use of steel balls or glass marbles of 27.5 mm,

17.5 mm and 12.5 mm placed in a drum along with the paste (Figure 3.12).

(a) Los Angeles abrasion testing machine (b) Glass marbles

(c) Mixing of paste with glass marbles (d) Sieving of cement paste

Fig. 3.12 Test set up for ball milling

The steel/glass balls were selected instead of stone aggregates because they do not absorb

water or paste and therefore it would be possible to extract the paste for the flow tests.

Glass marbles were selected because the density (2575 kg/m3) is almost same as that of

stone aggregates. Of course, there will be some differences between the effects of the

balls and the aggregates, as the size, shape and surface texture of these are different.

Cement paste and steel balls/glass marbles were used in the ratio of 1:1 by volume.

Different sizes of steel and glass marbles were selected based on the percentage fraction

of aggregates (Mehta and Monteiro, 2005); 30% were of 25mm size, 40% were of

65
17.5mm and 30%, 12.5mm. The paste was mixed by hand for 1 minute before putting it

into the drum and then the steel balls /glass marbles and superplasticizer were added and

the paste mixed for 4 minutes. The mixing was done by rotating the drum at 32 rpm. The

paste was screened through 4.75 mm and 600 micron sieves before the Marsh cone test.

3.4.4. Results and Discussions

Figure 3.13 shows the Marsh cone flow times for the different mixing methods and

dosages of superplasticizers. In all cases, there is an increase in fluidity with an increase

in the dosage of superplasticizer up to a certain dosage, which is taken as the saturation

dosage (Agullo et al., 1999). The saturation superplasticizer dosages in the case of Hobart

mixer with B-flat beater and ball milling with steel balls and glass marbles are

comparable as seen in Table 3.5. The unit weights of paste prepared are almost same, as

shown in Table 3.5, which implies that air entrainment does not vary for different mixing

methods. The blender mixing gives a higher fluidity but the saturation dosage is also

higher. In addition, the paste prepared with the blender had a temperature that was about

10-12 ºC higher than that prepared with the other methods. The saturation dosages

obtained through grinder mixing and Hobart with D wire whip is comparable. Hand

mixing gave the lowest fluidity. The results of three trials for selection of mixing method

are given in Table A.1 of Appendix A.

66
2.2
Hobart with B flat beater
2.0 Hobart with D wire whip
Blender mixing
1.8 Grinder mixing
log (flow time, sec)

Hand mixing
Ball milling with steel balls
1.6
Ball milling with glass marbles

1.4

1.2

1.0

0.0 0.1 0.2 0.3 0.4 0.5 0.6


sp/c%

Fig. 3.13 Typical flow time curves obtained with


Marsh cone for different mixing methods

Table 3.5 Optimum superplasticizer dosage and range of


unit weight for different mixing methods

Type of mixing Saturation Unit weight of


superplasticizer paste
dosage % (kg/ litre )

Hobart with B flat beater 0.09 1.97-2.03


Hobart with D wire whip 0.16 1.97-1.99
Blender mixing 0.16 1.94-2.02
Grinder mixing 0.16 1.97-2.05
Hand mixing 0.20 1.93-1.95
Ball mill effect (steel balls) 0.11 1.96-2.00
Ball mill effect (glass marbles) 0.14 1.94-1.98

The mini-slump spread generally increases with an increase in the dosage of

superplasticizer, as shown in Figure 3.14. The points that have been omitted in the figure

correspond to higher dosages where there is bleeding. Generally, it is seen that bleeding

67
occurs when the dosage is well above the saturation dosage and the tendency for it to

occur increases with the intensity of the mixing.

mini-slump spread (mm) 180

160

140
Hobart with B flat beater
Hobart with D wire whip
120 Blender mixing
Grinder mixing
Hand mixing
Ball milling with steel balls
100 Ball milling with glass marbles

0.0 0.1 0.2 0.3 0.4 0.5


sp/c %

Fig.3.14. Typical flow curves obtained with mini-slump for different mixing methods

Considering the different types of mixing studied here, it can be stated that the blender,

Hobart mixer with the flat beater and ball mill provide intensive mixing, with the blender

being the most intensive of the three while the hand mixing, grinder and whip mixing are

not intensive. In the case of intensive mixing, the mixing is expected to yield better

dispersion of the particles and the breaking up of the initial hydration products (e.g., first

growth of ettringite). The higher mixing intensity also appears to decrease the

superplasticizer demand, in terms of the saturation dosage, with the exception of the

blender, which results in a higher saturation point probably due to the higher surface area

of the particles due to the crushing of the hydration products. The effect of a high dosage

of superplasticizer is also more evident in the case of intensive mixing with bleeding

occurring beyond the saturation point. The values of the unit weight (Table 3.5) of the

paste from different mixing methods show that the variation for different dosages is

68
small. This shows that there is not much segregation of paste for different mixing

methods.

In a concrete mixer, it is expected that the cement paste is mixed in an intensive manner.

This was simulated, to a large extent, by the ball mill. Since the behaviour of the paste

mixed with the Hobart mixer with B-flat beater and the ball mill is similar in terms of

saturation dosage and fluidity, it can be concluded that the Hobart mixer leads to a paste

that is representative of that in concrete.

3.4.5 CONCLUSIONS

• The Marsh cone and mini-slump tests are suitable methods for the objective

comparison of the effects of different mixing methods on the fluidity of

superplasticized cement pastes. Both are simple methods to determine the

comparative fluid behaviour and saturation superplasticizer dosage.

• The mixing with Hobart mixer with a flat beater and a ball mill with glass/metal

can be categorized as intensive as in concrete mixing.

• In the case of blender mixing, the mixing is more intensive, leading to a higher

saturation dosage of superplasticizer and more fluidity. The temperature rise of

the paste prepared with this mixing method is 10-12 ºC higher than that with other

methods and causes rapid hydration leading to change in microstructure.

• The results presented indicate that the paste prepared with the Hobart mixer with

flat beater can be considered as representative of the paste of concrete since the

properties are similar to that of paste mixed in a ball mill, with steel/glass balls, in

terms of saturation dosage as well as flowability. Hence, pastes will be prepared

using the Hobart mixer with flat beater for the rest of work in this thesis.

69
3.5. INFLUENCE OF TYPE AND DOSAGE OF SUPERPLASTICIZER ON THE
FLOW BEHAVIOUR OF CEMENT PASTE

In the previous section, Marsh cone and mini-slump tests were used to study the fluidity

of cement paste and for selecting a suitable mixing method for the preparation of cement

paste. These are simple empirical methods to characterize flow behaviour of cement paste

whereas rheological tests are fundamental but more complicated and require skill for

conducting the tests. Hence, the establishment of correlations between these tests helps to

select an appropriate test to study the fluidity of cement paste and to understand the

results from empirical tests better. This study is more relevant considering

incompatibility problems associated with superplasticized cement paste (see Chapter 2),

which necessitate the use of simple tests for studying different combinations of cement

and admixture. As the admixtures mainly affect the flow behaviour of the cement paste

without significant interaction with the aggregate, it seems reasonable to select

admixtures and evaluate their performance through tests on the cement paste.

3.5.1 Testing Details

In the previous section, it was concluded that cement paste prepared with the Hobart

mixer and B flat beater represents that of concrete satisfactorily. Hence, the same mixing

method, with the exception that the superplasticizer addition was delayed further to two

minutes after the mixing of cement and water to get better efficiency. All the material

components were kept in a humidity chamber at a temperature of 27ºC for one day before

the tests in order to minimize the effect of ambient temperature on the flow behaviour.

Marsh cone, mini-slump and viscometer tests were done as explained in Section 3.3

70
3.5.2 Results and discussions

The Marsh cone flow time curves for different types and dosages of superplasticizers are

shown in Figure 3.15, along with the saturation dosages corresponding to the internal

angle of 140°± 10° for each case (following the method of Gomes et al., 2001); the

individual data are given in Table 3.6. The results show that fluidity increases with an

increase in the dosage of superplasticizer up to the saturation point, after which the curves

remain relatively flat.

1.8
1.7 SP saturation dosage
(sp/c%)
1.6 LS-C1 0.25
SNF-S1 0.20
log (flow time, sec)

SNF-D1 0.24
1.5 SNF-D2 0.23
SNF-C1 0.16
1.4 SMF-S1 0.23
SMF-C1 0.20
1.3 PCE-S1 0.20
PCE-D1 0.07
1.2
1.1
1.0

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %
Fig. 3.15 Marsh cone flow time for different superplasticizers

The lignosulphonate-based product LS-C1, though not expected to be as effective as the

other products, yields a response that is similar to those of the naphthalenes and

melamines. Most of the admixtures based on naphthalenes and melamines exhibit similar

flow time curves, except for SNF-C1, which yields lower flow times and a comparatively

low saturation superplasticizer dosage of 0.16%. In comparison, the PCE-based

superplasticizers are more effective, with lower saturation dosages than the other

products; PCE-D1 gives the lowest saturation dosage of 0.07%. These trends are in

71
accordance with the basic actions of the different families of superplasticizers (see

Chapter 2).

The mini-slump test results (see Figure 3.16) show that the spread increases with the

increase in dosage of superplasticizer. However, bleeding was observed in the paste

whenever the saturation superplasticizer dosage was exceeded. The mini-slump spread

(i.e., flow) is least for SNF-D2 and highest for PCE-D1. The time taken for a spread of

115 mm at saturation dosage is less than 1 sec.

200 LS-C1
SNF-S1
180 SNF-D1
Mini-slump spread (mm)

SNF-D2
160 SNF-C1
SMF-S1
SMF-C1
140
PCE-S1
PCE-D1
120

100

80
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 3.16 Mini-slump spread for different superplasticizers

In Figure 3.17 Marsh cone and mini-slump spread are plotted together, showing that as

the flow time increases, the mini-slump spread decreases. A better correlation is observed

for pastes with high fluidity.

72
200

180

Mini-slump spread (mm)


160

140

120

100

80

5 10 15 20 25 30 35 40 45 50
Marsh cone flow time (sec)

Fig. 3.17 Correlation between Marsh cone flow time and mini-slump spread

A typical set of shear stress-strain rate curves obtained from the rheological tests are

shown in Figure 3.18 (a-c) for pastes with the superplasticizer PCE-D1. At the low

dosage of sp/c = 0.05%, the area within the loading-unloading loop of first cycle is much

higher than the subsequent cycles, reflecting the energy needed for the structural

breakdown of the paste (see Figure 3.18 a). Moreover, it can be seen that the curves are

nonlinear reflecting the shear-thinning response. At the sp/c dosage of 0.1%, which is just

above the saturation dosage, the shear stress levels are remarkably lower, as seen in

Figure 3.18 b, and the stress-strain rate response is practically linear, especially in the

second and third cycles. At both dosages (see Figures 3.18 a&b), it is observed that the

areas within the loading-unloading loop decrease, along with the stress levels, in

subsequent cycles. At the higher superplasticizer dosage of 0.4% (see Figure 3.18 c), the

curves are similar to that of sp/c = 0.1% but with two significant differences; i.e., the

response of the third cycle is above that of the second cycle and the response in the third

73
cycle shows slight shear-thickening behaviour, indicated by a slight increase in the slope

of the stress-strain rate curve. These two aspects can be attributed to the settlement of the

cement particles due to the over dosage of the superplasticizer leading to the stiffening of

the lower part of the paste, as observed visually in this work and in previous works

(Bhatty and Banfill, 1982; Billberg, 1997; Cyr et al., 2000; Papo and Piani, 2004). At

higher strain rates, more of the settled flocs of cement particles are forced into the

suspension, which would increase the stress resulting in the strain-hardening effect.

Responses similar to those presented in Figures 3.17 a-c were observed for all the

admixtures, except that no shear thickening could be seen for LS-C1 and SNF-C1 even at

the dosage of 1%.

90
SP-PCE-D2
80 Dosage-0.05%

70
Shear stress (Pa)

60

50

40 First cycle
Second cycle
30 Third cycle

20
40 80 120 160
Shear rate (1/sec)

(a) Hysteresis cycles at low dosage

74
8

7 SP-PCE-D2
Dosage-0.1%
6
Shear stress (Pa) 5

3
First cycle
2 Second cycle
Third cycle
1

0
40 80 120 160
Shear rate (1/sec)

(b) Hysteresis cycles at optimum dosage

8 SP-PCE-D2
7 Dosage- 0.4%

6
Shear stress (Pa)

5
4
3
First cycle
2 Second cycle
Third cycle
1
0
40 80 120 160
Shear rate (1/sec)

(b) Hysteresis cycles at excess dosage

Fig. 3.18 Typical graphs showing three hysteresis cycles


for a PCE based superplasticized cement paste at
(a) low (b) optimum and (c) excess dosages

75
The shear stress-strain data were fitted in each case with the Bingham and Herschel-

Bulkley models, and the parameters obtained are given in Table 3.6. It can be observed

that, in general, an increase in the dosage of superplasticizer reduces the yield stress and

plastic viscosity of the Bingham model, which is attributed to the better dispersion of

cement particles (Rixom and Mailvaganam, 1999). The Bingham yield stress tends to

reduce to zero, as the superplasticizer dosage increases; the drop in yield stress is sharpest

in the case of the PCE-based products, confirming the results of previous works (Puertas

et al., 2005). The plastic viscosity decreases initially with an increase in the dosage of

superplasticizer but remains practically constant after a certain point. In the case of the

Herschel-Bulkley model, the yield stress ( τ 0 ) has the same trend as that of the Bingham

model though the values of the former are lower at smaller superplasticizer dosages and

vice versa at higher dosages. These differences are attributed to the change in the

rheological nature of the paste from shear-thinning to slight shear-thickening, as

discussed earlier. The consistency and power index (n) of the Herschel-Bulkley model are

interdependent and hence well defined trends could not be identified for the individual

parameters. Note that for shear-thinning materials, n < 1, and that for shear-thickening

materials, n > 1.

76
Table 3.6 Rheological characteristics and Marsh cone flow times

Type sp/c Bingham Herschel Bulkley Parameters Flow time


of sp % Parameters predicted by the
Nguyen et al.
approach

flow time (sec)


Experimental
τ0 μp τ0 K n
(Pa Sn)

Bing-ham
(Pa) (Pa S) (Pa)

Herschel
Bulk-ley
material
(sec)

(sec)
0.1 20 12.2 0.122 11.0 0.319 0.817 18 16
0.2 14 5.8 0.089 5.0 0.169 0.880 14 13
0.3 13 2.0 0.059 2.0 0.053 1.017 12 12
LS-C1 0.4 13 1.0 0.050 0.7 0.077 0.920 11 10
1 12 0.5 0.068 0 0.135 0.876 12 11
0.2 13 16.2 0.143 12.29 0.668 0.723 20 16
SNF- 0.3 11 1.8 0.055 1.66 0.066 0.965 11 11
S1 0.4 11 0.0 0.033 0.08 0.030 1.020 10 10
1 11 0.1 0.031 0.12 0.029 1.010 10 9
0.1 45 18.31 0.197 14.34 0.657 0.783 27 22
0.2 20 12.13 0.139 11.69 0.265 0.872 20 17
SNF- 0.3 13 2.88 0.074 3.09 0.059 1.041 13 13
D1 0.4 14 0.08 0.049 0.18 0.043 1.027 11 11
0.5 14 0.11 0.043 0.47 0.021 1.132 11 11
1 14 0.09 0.033 0.45 0.013 1.177 10 11
0.1 29 16.99 0.197 20.46 0.034 1.328 27 30
0.2 17 6.94 0.088 7.56 0.100 0.959 14 14
SNF- 0.4 14 0.55 0.042 0.98 0.017 1.168 11 12
D2 0.6 13 0.37 0.041 0.67 0.022 1.113 11 11
1 13 0.13 0.045 0.29 0.034 1.049 11 11
0.1 11 6.72 0.057 5.12 0.269 0.722 12 10
0.2 10 1.25 0.042 1.95 0.008 1.300 10 12
SNF- 0.3 10 0.72 0.029 1.20 0.006 1.303 9 11
C1 1 10 0.24 0.023 0 0.050 0.876 9 8
0.1 62 20.39 0.177 3.24 4.914 0.43 26 16
SMF- 0.2 18 6.89 0.076 6.33 0.103 0.948 14 14
S1 0.25 16 2.36 0.098 1.99 0.128 0.950 15 14
0.3 15 1.54 0.078 1.34 0.109 0.950 13 13
1 15 0.77 0.048 0.98 0.033 1.06 11 11
0.1 >120 33.14 0.199 20.53 2.514 0.559 36 24
0.2 16 13.44 0.128 6.36 1.07 0.638 19 14
SMF- 0.3 13 2.88 0.074 3.09 0.059 1.040 13 13
C1 0.5 12 0 0.039 0.08 0.025 1.090 10 11
1 12 0 0.037 0.13 0.020 1.115 10 11
0.05 26 17.55 0.155 15.93 0.310 0.873 23 20
0.1 16 11.07 0.089 16.01 2.399 0.558 15 20
PCE- 0.2 12 0.51 0.073 0.68 0.010 1.260 13 12
S1 0.3 11 0 0.034 0.44 0.007 1.270 10 11
1 12 0 0.034 0.39 0.006 1.344 10 12
0.05 18 24.45 0.214 1.65 6.910 0.410 32 19
0.1 12 0.01 0.035 0.37 0.014 1.168 10 11
PCE- 0.2 12 0 0.035 0.48 0.006 1.304 10 11
D1 0.4 12 0 0.036 0.53 0.006 1.310 10 11
1 14 0 0.047 0.07 0.020 1.158 10 12

77
3.5.3 Correlations between the Results from Empirical Tests and the Viscometer

In order to study the relationships between the results from empirical tests (i.e., the Marsh

cone and mini-slump tests) and the rheological parameters of cement pastes, the data

presented above are compared in this section. It can be seen in Table 3.6 that the Marsh

cone flow time increases as the yield stress and plastic viscosity increase. More

importantly, the variation of the Marsh cone flow time with the superplasticizer dosage

has the same trend as those exhibited by the yield stress and plastic viscosity, as also

illustrated in the typical plots of Figure 3.19 (shown for SMF-C1). It can be seen that all

three parameters decrease with an increase in superplasticizer dosage until the saturation

dosage and remain practically constant after that. Figure 3.19 also shows the trend of the

mini-slump spread, where it is seen that the spread increases with the superplasticizer

dosage until the saturation dosage, after which it is practically constant. These trends can

be generalized to all the paste systems studied in this work, and are in agreement with the

results of previous works (Gettu et al., 1997; Monte and Figueiredo, 2003;

Schwartzentruber et al., 2006). The correlation between Marsh cone flow time with

rheological parameters for other superplasticized pastes are given in Figures A.1 and A.2

of Appendix A.

78
log (flow time, sec)
mini slump spread (mm)
SMF-C1 SMF-C1 200
Yield stress
Plastic viscosity

160

Plastic viscosity (cP)


1.24 200 40

mini slump spread (mm)


120
1.2 180
log (flow time, sec)

30

Yield stress (Pa)


80
1.16 160
20
40
1.12 140

10 0
1.08 120

1.04 100 0
0 0.2 0.4 0.6 0.8 1
sp/c%

Fig. 3.19 Comparison of data from different tests

The Roussel-Coussot model (Roussel and Coussot, 2005) can be used for estimating the

yield stress from the measured conical slump. The model is applicable for different flow

regimes based on the height (H) and radius (R) of the slump cone. According to this

model, the spread distance (R) can be expressed as a function of the yield stress (τc) and

material volume (Ω) and density of material (ρ) as expressed in the equation 3.3.

225 ρ g Ω 2
τc = (3.3)
128 π 2 R 5

The prediction of yield stress from mini-slump spread using the Roussel-Coussot model

is given in Figure 3.20. The result show that even though the graph shows a good

correlation, individual experimental data and predicted values are not comparable.

79
40
Experimental
35 Predicted
30

Yield stress (Pa)


25
20
15
10
5
0
80 100 120 140 160 180 200
Mini-slump spread (mm)

Fig. 3.20 Prediction of yield stress from mini-slump spread using Roussel-Coussot model

It is clear that the saturation dosage estimated with the procedure of Gomes et al. (2001)

from the Marsh cone test gives a reliable approximation of the superplasticizer dosage

beyond which the plastic viscosity and yield shear stress remain constant. On the other

hand, the mini-slump test on the tested pastes does not seem to give as reliable an

estimate of the saturation dosage due to bleeding that may occur at higher dosages. From

the data obtained, it does not seem possible to identify unique relations between the

Marsh cone flow time and mini-slump spread with either the plastic viscosity or yield

shear stress, as previously concluded (Ferraris et al., 2001). Nevertheless, it was shown

that the use of flow times from two Marsh cones of slightly different geometries can be

used to determine the rheological characteristics of a paste (Roussel and Roy, 2005).

Conversely, the Marsh cone flow time can be obtained by adequately modeling the flow

and material behaviour, as in the semi-analytical approach, where the Marsh cone flow

80
time is obtained from fluid mechanics principles and finite volume method (Nguyen et

al., 2006).

Nguyen et al. (2006) suggested a semi analytical approach by for relating rheological

parameters to Marsh cone flow time, where it is given that the flow characteristics of

cement paste can be described by Herschel-Bulkley model by three parameters τ 0 , k and

n which relates the shear stress to shear rate. As per this method, the flow time is related

to rheological parameters through the equation 3.4.

H2
( R + H t Tanφ ) 2
t= ∫
H1 Vβ R 2
dH t (3.4)

where R is the radius of the nozzle, Ht is height of fluid in the Marsh cone at any time,

φ is the angle of the Marsh cone, V β is the corrected velocity at the exit of the nozzle.

t is the time taken to lower the level of paste in the Marsh cone from H1 to H2.

The flow times for the different pastes studied here have been predicted using this

approach considering both the Bingham and Herschel-Bulkley models for the paste and

given in Table 3.6. Note that the density of paste was taken as 1.9 kg/litre in the

predictions. A Matlab programme was used for solving the equations and for predicting

the flow time; the programme code is given in Appendix B.

The results in Table 3.6 show that the predictions of the Nguyen et al. approach are, in

general, good for both the rheological models. This confirms that the Marsh cone flow

time can be determined through appropriate modeling of the material response and the

flow behaviour.

81
3.5.4 Influence of Type and Dosage of the Superplasticizer on the Non-Newtonian
Characteristics of Cement Paste

The non-Newtonian characteristics of the superplasticized cement paste depend on the

type of superplasticizer and its dosage and generally rheological model parameters are

used to represent the same, as explained in the previous Sections. However, they do not

represent the transition in the fluid nature based on the dosage of the superplasticizer.

Therefore, steady shear, creep and recovery, and stress relaxation tests were also

conducted in order to develop an understanding of the transition of the behaviour of these

mixtures from viscoelastic to non-Newtonian to linearly viscous, as the superplasticizer

dosage is increased. The test results show that as the superplasticizer dosage increases, in

general, there is a transition in the behaviour from viscoelastic to viscous fluid behaviour.

It was found that the dosages determined based on the Marsh cone tests were close to the

dosages at the transition to the linearly viscous behaviour. The experimental details and

the test results are given in detail in Appendix C.

3.5.5 Change of Flow behaviour with Time

The flow properties of cement paste, especially that of superplasticized cement paste, can

change significantly after mixing. This is attributed to hydration, intercalation of

superplasticizers, chemical degradation of the polymer due to the high pH of the aqueous

solution and physical relaxation or change in molecular structure due to adsorption

(Tattersall and Banfill, 1983; Justnes et al., 2003). In general, both yield stress and plastic

viscosity tend to increase with time (Roy and Asaga, 1980). Consequently, the Marsh

cone flow time tends to increase with time (Agullo et al., 1999) whereas the mini-slump

spread decreases (Kantro, 1980; Lim et al., 1999).

82
The change in paste flow behaviour has been studied here by comparing the test results

obtained for the same materials just after mixing and 60 minutes later. During the interim

period the paste was maintained in an air-tight container at 27ºC. The paste was taken out

of the container and mixed for 15 seconds before performing the 60-minute tests. The

variation of the flow parameters with time using Bingham model can be seen from the

results given in Table 3.7.

Table 3.7 Test data for pastes with different superplasticizers at 0 and 60-minutes

Flow time Mini-slump spread Yield stress Plastic viscosity


Type of (sec) (mm) (Pa) (Pa s)
sp sp/c %
0 60 0 60 0 60 0 60
min min min min min min min min
0.1 20 21 93 90 12.20 13.50 0.122 0.132
0.2 14 16 153 123 5.81 7.59 0.089 0.097
LS-C1 0.3 13 13 162 143 2.0 3.15 0.059 0.075
0.4 13 14 163 163 1.0 0.35 0.050 0.064
1 12 13 164 162 0.5 0.23 0.068 0.049
0.2 13 26 143 90 16.21 24.61 0.143 0.185
SNF-S1 0.3 11 11 151 150 1.80 5.91 0.055 0.066
0.4 11 11 165 163 0.0 0.21 0.033 0.054
1 11 11 170 178 0.09 0.10 0.031 0.025
0.1 45 * 82 * 18.31 20.23 0.197 0.241
0.2 20 70 96 * 12.13 16.61 0.139 0.178
SNF-D1 0.3 13 13 127 115 2.88 5.98 0.074 0.089
0.4 14 17 bleeding 141 0.08 0.09 0.049 0.083
0.5 14 18 bleeding 164 0.11 0 0.043 0.045
1 14 13 bleeding 163 0.09 0.12 0.033 0.035

SNF- D2 0.1 29 * 87 * 16.99 25.52 0.197 0.194


0.2 17 39 106 87 6.94 12.10 0.088 0.149
0.4 14 14 bleeding 148 0.55 0 0.042 0.043
0.6 13 13 bleeding 176 0.37 0.25 0.041 0.039
1 13 12 bleeding 160 0.13 0.07 0.045 0.038

SNF-C1 0.1 11 14 126 95 6.72 7.24 0.057 0.046


0.2 10 11 152 128 1.25 6.69 0.042 0.064
0.3 10 10 168 168 0.72 0.77 0.029 0.046
1 10 9 173 178 0.24 0.17 0.023 0.035
0.1 62 * * 20.39 * 0.177 *
SMF-S1 0.2 18 44 102 90 6.89 2.39 0.076 0.092
0.25 16 22 115 109 2.36 3.77 0.098 0.101
0.3 15 18 122 113 1.54 8.97 0.078 0.065
1 15 18 167 161 0.77 0 0.048 0.043

83
Table 3.7 (Contd..)
0.1 >120 * * * 33.14 * 0.199 *
SMF-C1 0.2 16 26 108 103 13.44 30.92 0.128 0.139
0.3 13 16 132 112 2.88 7.71 0.074 0.100
0.5 12 12 181 169 0 0 0.039 0.042
1 12 11 177 172 0 0 0.037 0.041

0.05 26 * 86.3 * 17.55 * 0.155 *


PCE-S1 0.1 16 * 110 * 11.07 23.67 0.089 0.162
0.2 12 14 bleeding 157 0.51 0.36 0.073 0.068
0.3 11 13 bleeding 179 0 0.08 0.034
0.5 11 14 bleeding 192 0 0 * 0.035
1 12 12 bleeding bleeding 0 0.034 0.034

0.05 18 * 97 * 24.45 * 0.214 *


PCE-D1 0.1 12 16 166 147 0.01 2.37 0.035 0.079
0.2 12 14 170 bleeding 0 0.09 0.035 0.043
0.4 12 14 199 bleeding 0 0 0.036 0.036
1 14 16 193 bleeding 0 0 0.047 0.054
* could not be determined

Generally, in all the compositions tested, the rheological parameters increase with time

after mixing, with the increase in yield stress being more pronounced than that of the

plastic viscosity. Also, the rate of increase is more significant at superplasticizer dosages

less than the saturation point. Similarly, when the dosages are low, the Marsh cone flow

time increases with age as seen when the 60-minute data are compared with those

obtained immediately after mixing (See Figure 3.21).

1.8 SMF-S1
flow time (0-minutes)
1.7 flow time 60-minutes

1.6
log (flow time, sec)

1.5

1.4

1.3 Saturation dosage


1.2

1.1

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %
Fig. 3.21 Flow time curve showing the determination of optimum
superplasticizer dosage according to Aïtcin, 1998

84
Note that when the flow behaviour at 60 minutes is more significant (for a particular

application) than the immediate behaviour, the saturation point could be defined based on

the flow time curves at 60 minutes, or any other critical time (See Table A.2 of Appendix

A). Nevertheless, Aïtcin (1998) suggests that the flow time and saturation dosage at 60

minutes should not differ significantly from those obtained immediately for a cement-

superplasticizer combination to be considered compatible.

3.5.6 Summary

The following conclusions can be made from the study performed on the flow-related

properties of superplasticized cement pastes.

• The flow time determined from Marsh cone test is a good indication of the

relative fluidity of superplasticized paste and the saturation dosage obtained from

the same test is appropriate for the selection of the superplasticizer.

• Both the Herschel-Bulkley and Bingham models fit the experimental data from

the viscometer study satisfactorily. It is observed that the nature of flow in

superplasticized paste varies slightly with the dosage of superplasticizer, as

follows. At lower dosages, nonlinear shear thinning is generally observed; around

the saturation dosage the response follows the Bingham model; and at higher

dosages, there may be some shear thickening. The yield stress values obtained

with both the models have the same trend with respect to superplasticizer dosage,

though the values from the Bingham model tend to be higher at smaller

superplasticizer dosages. Even though shear-thickening nature of paste is better

represented through the Herschel Bulkley model, the Bingham model represents

the behaviour of normal pastes well. The Marsh cone flow times could be

85
predicted satisfactorily through the Nguyen et al. approach with the parameters of

both models.

• An increase in superplasticizer dosage leads to a decrease in the yield stress,

plastic viscosity and Marsh cone flow time, and an increase in mini-slump spread,

as long as the dosages are below the saturation point. Beyond the saturation

dosage, these parameters are practically constant.

• All the nine superplasticizers tested resulted in pastes with good fluidity over 60

minutes when the corresponding saturation dosages were used. It is concluded

that the cement-superplasticizer combinations studied here are all compatible as

far as the flow behaviour is concerned.

86
CHAPTER 4

CORRELATIONS BETWEEN THE FRESH AND


HARDENED PROPERTIES OF SUPERPLASTICIZED
PASTE, MORTAR AND CONCRETE

4.1 GENERAL

Optimization and characterisation of the superplasticized paste is considered as a first

step in the design of high performance concrete, as the paste phase is mainly

responsible for workability, mechanical properties and durability of concrete. The

incorporation of superplasticizers results in the production of high performance

concrete due to several beneficial effects associated with it in the fresh and hardened

state of concrete. However, the characteristics of cement paste and type and dosage of

superplasticizers could sometimes compromise the benefits of incorporating the

superplasticizer because of compatibility problems, leading to slump loss, retardation

in setting time or excessive heat of hydration (Rixom and Mailvaganam, 1999).

Therefore, for the optimization of the composition of high performance concretes

(HPC), the optimization of the paste and mortar phases are essential and should

include the choice of superplasticizer type and dosage in order to obtain the desired

fresh and hardened properties of concrete. The methodology for the determination of

the saturation dosage of superplasticizer through the Marsh cone test and its influence

on the loss of fluidity has been explained in the third chapter.

Tests involving the paste are simpler and require much less effort than tests of

concrete. Nevertheless, results obtained from paste tests need to be correlated to the

behaviour of concrete to take the possible influence of aggregates and the aggregate-

paste interfaces into account (Ferraris and Gaidis, 1992; Toralles-Carbonari et al.,

1996; Roncero et al., 2000; Tang et al., 2001; Giaccio and Zerbino, 2002; Banfill,
2003; Laskar and Talukdar, 2007). For a particular particle size distribution and

volumetric fraction of aggregate, the flow properties of concrete are related with

rheology of paste, which is generally studied through simple flow or viscometer tests

as explained in Chapter 3 (Ferraris and Gaidis, 1992; Toralles-Carbonari et al., 1996;

Roncero et al., 2000; Lachemi et al., 2007).

The work presented in this chapter deals with the comparison of the results of tests on

superplasticized cement pastes with the behaviour of mortars and concretes. The

fluidity of mortar is studied through Marsh cone and flow table spread. Finally, an

attempt is made to correlate the cement paste flow behaviour with that of concrete as

characterized through the slump, setting time and compressive strength.

4.2 EXPERIMENTAL DETAILS

4.2.1 Materials

The physical and chemical characteristics of cement and four types of

superplasticizers used for the correlation study are same as given in Section 3.2. River

sand (of 0-4.75 mm grain size range) and crushed granite coarse aggregate (of 4.75-10

mm and 10-20 mm grain size ranges), satisfying the requirements of IS 2386 (2007),

are employed in the preparation of concrete.

The sieve analysis was done according to the Indian Standard IS 2386 (2007) and the

test results on fine and coarse aggregates are given in Figures 4.1 and 4.2. The sand

has a fineness modulus of 2.29 and conforms to grading zone II as per IS 383 (2002).

The properties of aggregates are given in Table 4.1. The specific gravity, bulk density

and coefficient of water absorption were determined for each aggregate according to

IS 2386 (2007). The absorption coefficients for the 10-20 mm, 4.75-10 mm and 0-

4.75 mm aggregates are 0.4%, 0.5% and 1.5%, respectively.

88
100

Percentage passing
80

60

40

20

0
0.1 1 10
Sieve size (mm)

Fig. 4.1 Particle size distribution of fine aggregates (0-4.75 mm)

10 mm down
20 mm down

100
Percentage passing

80

60

40

20

0
1 10 100
Sieve size (mm)

Fig. 4.2 Particle size distribution of coarse aggregates (4.75-20 mm)

Table 4.1 Summary of properties of aggregates

Properties Sand Coarse aggregates

4.75-10 mm 10-20 mm
Specific gravity 2.62 2.70 2.72
Water absorption (%) 1.5 0.5 0.4
Fineness modulus 2.29 5.71 7.51
Bulk density (kg/ m3) 1670 1568 1618
Flakiness index (%) - - 16.1
Elongation index (%) - - 29
Angularity number - - 8
Crushing Value (%) - - 28
Impact Value (%) - - 30

89
4.2.2 Test Procedures

4.2.2.1 Tests on Paste

The Marsh cone, mini-slump and viscometer tests, as explained in Chapter 3, were

conducted for understanding the flow behaviour of the paste. The Marsh cone test

results for four superplasticized pastes, already presented in Chapter 3, are

summarized in Section 4.3.1.1 for making the comparison of flow behaviours of

paste, mortar and concrete possible.

4.2.2.2 Tests on Mortar

Marsh cone and flow table tests (as per IS 1199 (2004) and ASTM C109 (1998)) were

used for understanding the fluidity of mortar. A Marsh cone with an aperture diameter

of 12.5 mm was used for the mortars. In addition, the consistency of mortars has been

determined using a flow table. Here, a hollow truncated metallic cone with a base

diameter of 100 mm, top diameter of 70 mm and a height of 50 mm was used. It was

placed on the table and filled with mortar, the cone was removed and the mortar was

jolted 25 times in 25 seconds causing it to spread. The mean final base diameter was

taken as the flow of consistency. In order to account for the higher fluidity of

superplasticized cement mortar, the flow table was modified by increasing its

diameter using a polycarbonate plate, as shown Figure 4.3. This is in accordance with

the test procedure of Domone (2006) for studying the flow behaviour of

superplasticized mortars.

Fig. 4.3 Flow table test for mortar


90
4.2.2.3 Tests on Concrete
In order to correlate the results obtained from paste tests to the behaviour of concrete,

tests were conducted on concrete with four superplasticizers studied for paste and

mortar. The fresh behaviour of concrete has been characterized using the slump test

and the flow table test, as per IS 1199 (2004), as shown in Figure 4.4. For concrete,

the following mix proportions were used per m3: 450 kg cement, 675 kg sand, 454 kg

of 20 mm coarse aggregate and 680 kg of 10 mm coarse aggregate with a w/c of 0.35

(mix design as per IS 10262 (2004)). Note that the sand-cement ratio was maintained

as 1.5 in both the mortar and concrete. In all the cases, the aggregates were assumed

to be in a saturated surface condition and, therefore, the water needed for saturating

them was added according to the absorption coefficients given in Table 4.1. The

humidity of the aggregates was compensated for, after its determination using the

ASTM microwave method (ASTM D 4643 (1987)). The fresh density of concrete

with different superplasticizers was also measured to account for the probable air

entrainment due to superplasticizers. In addition to the slump and flow table tests, for

all superplasticizer dosages, the loss of slump and setting time of concrete at the

saturation dosage of superplasticizer were determined. Also, the compressive

strengths at 3, 7 and 28 days were evaluated for different dosages of superplasticizers.

(a) Slump test

(b) Flow table test on concrete


Fig. 4.4 Tests for workability of concrete

91
4.2.3 Mixing Methods for Paste, Mortar and Concrete

The pastes and mortars were prepared in the Hobart mixer (1/6 hp) with the B flat

beater at low speed (i.e., shaft speed of 139 rpm and planetary speed of 61 rpm),

based on the study explained in Chapter 3. For mixing of the paste, the cement and

70% of the water required were mixed for 2 minutes; after that, the superplasticizer,

along with the remaining water (to ensure better dispersion of superplasticizer), was

added to the cement paste; the mixing was stopped and the sides of mixer bowl were

scraped (for 15-30 seconds); and the paste was again mixed for 3 minutes.

In the case of mortar mixing, oven-dried sand was mixed with the water necessary for

its saturation for 30 seconds. Then the cement was added and mixed with the wetted

sand for 15 seconds to ensure the homogeneity of solid particles before the

incorporation of water. The same sequence as in cement paste was followed for the

addition of water and superplasticizer.

For the concrete, the materials were dry mixed and the subsequent mixing sequence

for the addition of water and superplasticizer was the same as that for paste; the total

mixing time was 5 minutes.

A water-cement ratio (w/c) of 0.35 was employed throughout this study. Also, note

that all materials, including distilled water, that were used in the paste and mortar

mixes, were kept at a temperature of 27° C in an environmental chamber for at least

24 hours before mixing. This was done to limit the influence of the ambient

temperature on the results.

92
4.3 RESULTS AND DISCUSSIONS

4.3.1 Comparison of the Flow Behaviour of Paste, Mortar and Concrete

4.3.1.1 Flow Behaviour of Paste

Figure 4.5 shows the flow curves for different superplasticized pastes obtained with

the Marsh cone immediately after mixing (0 minutes), along with the corresponding

saturation dosages; when the flow time is higher, fluidity is less. The saturation

dosages and the corresponding flow times are given in Table 4.2. The PCE based

superplasticizer performs better than the other superplasticizers in terms of fluidity as

well as the saturation dosage (0.07%), as expected. The pastes with low dosages of

SMF-S1 exhibit high flow times indicating poorer dispersion of the cement particles

at such dosages. Also, at saturation, superplasticizers have different flow times

indicating their different effectiveness.

1.8
Type of SP saturation dosage
1.7 (sp/c%)
SNF-S1 0.20
1.6
log (flow time, sec)

SNF-D2 0.23
1.5 SMF-S1 0.23
PCE-D1 0.07
1.4
1.3
1.2
1.1
1.0
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 4.5 Marsh cone flow time curves for superplasticized cement paste

93
Table 4.2 Summary of Marsh cone data and concrete cost

Type Marsh cone test results *Dosage Unit *Cost of


of sp of sp cost/kg materials
Paste Mortar needed (Rs./kg) for
Saturation Flow Saturation Flow (kg/m3) concrete
dosage of time at dosage of time at (Rs./m3)
sp saturation sp saturation
(sp/c%) (sec) (sp/c%) (sec)
SNF- 0.20 13 0.22 13 0.23 35 2538
S1
SNF- 0.23 17 0.23 22 0.25 50 2570
D2
SMF- 0.23 16 0.30 18 0.37 60 2692
S1
PCE- 0.07 12 0.07 11 0.08 160 2616
D1
*for a mix with a slump of 100 mm

4.3.1.2 Flow Behaviour of Mortar

The results obtained from Marsh cone test of mortar are shown in Figure 4.6 and in

Table 4.2. The fluidity of the PCE based superplasticized cement mortar at low

dosages is higher than that with the SNF and SMF based superplasticizers. The PCE

yields the highest fluidity with a relatively low saturation dosage whereas the SMF

gives the least fluidity and the highest saturation dosage compared to the other

products. This is because, even at low dosages, better dispersion is produced due to

steric repulsion in the case of the former. Also, it has been observed that high dosages

(i.e., above the saturation point) of the SNF with retarder (i.e., SNF-D2) lead to

bleeding due to the loss of cohesion in the paste. The flow table test results are given

in Figure 4.7, where it can be seen that the spread generally increases with an increase

in the dosage of superplasticizer up to the saturation dosage and remains practically

constant beyond that; all the mortars exhibited spreads of 250-300 mm at saturation.

94
1.5 SNF-S1
SNF-D2
SMF-S1
1.4

log (flow time, sec)


PCE-D1

1.3

1.2

1.1

0.0 0.2 0.4 0.6 0.8


sp/c %

Fig. 4.6 Marsh cone flow times of cement mortar

300
flow table spread (mm)

SNF-S1
SNF-D2
250 SMF-S1
PCE-D1

200

150

100
0.0 0.2 0.4 0.6 0.8
sp/c %

Fig. 4.7 Flow table spreads for superplasticized mortars

Considering that both the flow table spread and the inverse of the Marsh cone flow

time are measures of fluidity, it can be seen that the trends obtained in the results with

increasing dosage are similar for all the mortars studied. Moreover, the saturation

dosages obtained in the mortars are comparable with those obtained from the test

results of paste.

95
4.3.1.3 Flow Behaviour of Concrete

In order to correlate the behaviour of cement paste and mortar to that of concrete and

to validate the test of cement paste as a tool for the mix design of high performance

concrete, tests were also conducted on concrete with four superplasticizers. The flow

behaviour of concrete, with the same paste and mortar compositions as those

discussed in the previous sections, has been characterized using the slump and flow

table tests as per IS 1199 (2004). The flow table test, which is similar in principle to

the flow table test for mortar, was used to measure the flowability of concrete and to

observe segregation. Concrete was moulded within the 120 mm high hollow frustum

of a cone with bottom and top diameters of 250 mm and 170 mm, respectively. After

removing the mould, the spread diameter was measured in four perpendicular

directions after jolting for 15 times in 15 seconds. The results of the two tests on fresh

concrete are shown in Figures 4.8 and 4.9. The data for all the concretes reflect an

increase in slump (see Fig. 4.8) as the superplasticizer dosage increases, as expected.

In the case of the naphthalenes, it can be seen that SNF-S1 gives a relatively low

slump at its saturation dosage (0.2%) but the slump increases rapidly just beyond this

dosage, whereas SNF-D2 gives a steady increase in slump with an increase in dosage

though the workability achieved at higher dosages is not as much as with SNF-S1. In

comparison, the slump values obtained with SMF-S1 are lower for similar dosages of

the other superplasticizers. On the other hand, a fluid concrete with a slump of more

than 100 mm is obtained for PCE-D1 even at the dosage of 0.1%.

It can be observed in Figure 4.8 that in order to obtain a slump of 100 mm, the

required dosage of superplasticizer is 0.08% for the PCE, about 0.25% for the

naphthalenes and 0.37% for the SMF based superplasticizer. The material costs for

concretes possessing this slump value have been estimated and are presented in

96
Table 4.2. The comparison of the costs shows that a low-cost superplasticizer need not

always result in cost-effective concrete as the dosage of the superplasticizer for

optimum fluidity is also an important factor (Gettu et al., 2006). The results of the

flow table spread tests are generally in accordance with those of the slump tests,

exhibiting an increase in spread with an increase in the dosage of superplasticizer (see

Figure 4.9). Again, the PCE based product performs best and the SMF based product

is the least effective.

200
Concrete slump (mm)

150

100

SNF-S1
SNF-D2
50
SMF-S1
PCE-D1

0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
sp/c %
Fig. 4.8 Variation of concrete slump with dosage of superplasticizer

420

400
Flow table spread (mm)

380

360

340
SNF-S1
SNF-D2
320 SMF-S1
PCE-D1
300
0.0 0.1 0.2 0.3 0.4 0.5 0.6
sp/c %
Fig. 4.9 Variation of flow table spread with dosage of superplasticizer

97
Note that the data from the slump and flow table tests have not been plotted for cases

where segregation or bleeding is observed. It is seen that there is a high tendency for

segregation and/or bleeding when the superplasticizer dosage is much higher than the

saturation dosage, and almost certain to occur when the dosage is close to twice that

of saturation. For example, the concrete with a 0.3% dosage of the PCE showed

severe bleeding, segregation and stiffening within 5 minutes after mixing. The

stiffening seen when the superplasticiser is overdosed could be attributed to shear

thickening caused by the excess of superplasticizer in the aqueous phase that could

interlock leading to the stiffening (Papo and Piani, 2004). This has also been observed

in cement pastes in the present study, especially at high dosages.

4.3.2 Correlation between the Fluidity of Paste, Mortar and Concrete

In Figures 4.10-4.13, the results from the Marsh cone tests of pastes and mortars are

plotted along with the slump values of concrete as a function of the superplasticizer

dosage. It can be seen that the Marsh cone flow time curves of paste and mortar are

generally parallel and the saturation dosages are comparable, as previously observed

by other researchers (Roncero et al., 1999; Banfill, 2003; Lachemi et al., 2007). The

test results of concrete with SNF-S1 and SMF-S1 indicate a low workability at the

saturation dosage of paste obtained from the Marsh cone tests whereas in the cases of

SNF-D2 and PCE-D1 the saturation dosages yield medium workability in the

concrete. In all the cases, a dosage slightly higher than saturation provides good

workability. This can be attributed to the absorption of part of the superplasticizer by

the fines in the crushed coarse aggregate This is also confirmed through the

adsorption studies; the absorbance of the superplasticized paste with aggregate was

higher than that of superplasticized paste (the details of the test is given in Chapter 6).

Consequently, the amount of superplasticizer available in the cement paste decreases

and maximum dispersion of cement particles is not achieved in the concrete at the

98
same dosage as in the paste tests. Therefore, further addition of superplasticizer is

required for high slump values in the concrete. Nevertheless, the saturation dosage

obtained from paste studies can be used as a guideline for the selection of

superplasticizers and the dosage needed in concrete, especially when the gravel or

sand does not have large amounts of fines, where it can be expected that the saturation

dosages obtained from the paste tests will yield good workability in the concrete, as

reported by Roncero et al. (1999), and Hidalgo et al. (2008).

1.40
SNF-S1
200
1.35

1.30
log (flow time, sec)

Paste flow time 150

Slump (mm)
Mortar flow time
1.25 Concrete slump
100
1.20

1.15 50

1.10
0
0.1 0.2 0.3 0.4
sp/c %
Fig.4.10 Correlation between paste, mortar and concrete
in terms of flow time and slump for SNF-S1
1.50 180
SNF-D2
1.45 160

1.40 140
log (flow time, sec)

1.35 Paste flow time 120


Mortar flow time
Slump (mm)

1.30 concrete slump 100


80
1.25
60
1.20
40
1.15
20
1.10
0
0.0 0.2 0.4 0.6 0.8
sp/c%
Fig. 4.11 Correlation between paste, mortar and concrete
in terms of flow time and slump for SNF-D2

99
1.8 SMF-S1
200
1.7

1.6

log (flow time, sec)


150
Paste flow time

Slump (mm)
1.5 Mortar flow time
Concrete slump 100
1.4

1.3
50
1.2

1.1
0
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %
Fig. 4.12 Correlation between paste, mortar and concrete
in terms of flow time and slump for SMF-S1

200
1.25 PCE-D1 180
Paste flow time
Mortar flow time 160
Concrete slump
1.20 140
log (flow time, sec)

120

Slump (mm)
1.15 100
80
1.10 60
40
1.05 20
0
0.0 0.1 0.2 0.3 0.4 0.5 0.6
sp/c %

Fig. 4.13 Correlation between paste, mortar and concrete


in terms of flow time and slump for PCE-D1

Figure 4.14 shows that no clear correlation exists between the yield stress and plastic

viscosity of paste with the slump of concrete. Nevertheless, some observations can be

made regarding the data obtained. It appears that a yield stress of less than 5 Pa and

plastic viscosity of less than 0.05 Pa s is required to achieve a slump of more than

100 mm. This is in accordance with the reported results of Lachemi et al. (2007), and

100
Hidalgo et al. (2008). It also appears that the concrete can be categorized as low

(<25mm) medium (25-50mm) and high slump (>100 mm) where the values of the

yield stress and plastic viscosity of paste have different ranges as indicated in

Figure 4.14.

24 Slump vs Yield stress


Slump vs Plastic viscosity 0.20

Plastic viscosity (Pa s)


20
0.16
Yield stress (Pa)

16
0.12
12
0.08
8

4 0.04

0 0.00
0 50 100 150 200
Concrete slump (mm)

Fig. 4.14 Correlation between slump of concrete, and the


yield stress and plastic viscosity of paste

4.4. LOSS OF FLUIDITY

4.4.1. Loss of Fluidity of Paste

The change in the flow behaviour of paste with time has been studied by comparing

the results obtained just after mixing and 60 minutes later. During the interim period,

the paste was maintained in an air-tight container at 27ºC, after which it is taken out

of the container and re-mixed for 15 seconds before performing the 60-minute tests.

The Marsh cone flow-time curves obtained after 60 minutes are given in Figure 4.15

(along with the data obtained immediately after mixing plotted with dashed lines for

reference). It can be seen that the curves for the 60-minute tests have the same trends

as the curves obtained immediately after mixing (denoted hereafter as 0-minute tests),

with a decrease in flow time with increase in the superplasticizer dosage upto the

101
saturation dosage, beyond which the flow time does not change much or may even

increase slightly (as in the cases of the SMF and PCE based superplasticizers) due to

settling of the cement particles in the paste. As before (see Section 3.5.2), it can be

observed that the saturation dosages differ with the type of superplasticizer (see

values in the inset of the figure).

More importantly, when the 60-minute plots are compared with those obtained at 0

minutes, it can be observed that there is a shift to the right for all the combinations

due to higher flow times at 60 minutes, indicating loss of fluidity over this period. The

saturation dosage of superplasticizers also increases with time with saturation at

60 minutes occurring at dosages significantly higher than at 0 minutes. This can be

attributed to the increase in surface area of the hydrated cement particles and the

covering up of the superplasticizer by the newly formed hydrates (see Chapter 2 for

further explanations).

1.8
Type of Sp (time) Saturation dosage
SNF-S1 (0) 0.20
1.6 SNF-S1 (60) 0.30
log (flow time, sec)

SNF-D2 (0) 0.23


SNF-D2 (60) 0.40
SMF-S1 (0) 0.23
SMF-S1 (60) 0.30
1.4 PCE-D1 (0) 0.07
PCE-D1 (60) 0.17

1.2

1.0

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %
Fig.4.15 Loss of fluidity of paste

102
The plots in Figure 4.15 also show that the difference in flow time between the 0- and

60-minute curves is higher at lower superplasticizer dosages, with comparable levels

of fluidity being achieved at 60 minutes only with initial overdosage of the

superplasticizer (i.e., dosage higher than the 0-minute saturation point), except in the

case of the PCE, where the 60-minute curve is always above that at 0 minutes. The

high fluidity loss at low dosages can be attributed to more rapid flocculation due to

inadequate dispersion in pastes with less than saturation dosage. In the context of

readymix concrete, the results suggest that when the fluidity of the paste after 60

minutes is more important than the immediate (0-minute) fluidity, the saturation

dosage corresponding to 60 minutes should be used in the mix design rather than

arbitrarily overdosing the superplasticizer, as is often done in practice.

4.4.2 Loss of Fluidity for Mortar

To evaluate the change in the flow behaviour of mortar with time, tests were

conducted with the SMF and PCE based superplasticizers that led to higher loss of

fluidity in the paste tests than the others; Marsh cone flow time and flow table spread

were measured immediately after mixing and after 60 minutes, and the test results are

presented in Table 4.3. It is observed that the flow time increases and the flow table

spread decreases with time for both the superplasticizers, especially at lower dosages.

The loss of fluidity is higher for the SMF than the PCE, as expected (Chandra and

Bjornstrom, 2002b). Note that in some cases the mortar did not pass through the

Marsh cone or there were clear signs of bleeding indicating the segregation of the

mortar.

103
Table 4.3 Loss of fluidity of mortar

Type of sp sp/c % Marsh cone flow time (sec) Flow table spread (mm)
0 60 0 60
0.2 22 no flow 242 193
0.3 21 no flow 265 198
0.4 13 29 bleeding 260
SMF-S1 0.5 11 15 bleeding 305
0.8 11 14 bleeding bleeding
0.05 13 no flow 310 203
0.1 11 19 bleeding 281
PCE-D1 0.15 11 15 bleeding 290
0.2 11 12 bleeding 313
0.3 11 12 bleeding 315

In the case of SMF based mortar, low dosages of superplasticizers are not sufficient

even to initiate the flow after 60 minutes. Moreover, loss of fluidity occurs in the case

of the SMF even at the saturation dosage corresponding to the 60-minute tests of

pastes whereas for the PCE, there is no appreciable loss at the 60-minute saturation

dosage. These results imply that the PCE is more effective in maintaining the fluidity

than the SMF, especially at high dosages. However, in both cases the loss of fluidity

in the mortar is more than that of the paste.

The results clearly show that when overdosing of the superplasticizer is done to obtain

flowable mixes after one hour or so after mixing, the initial mix may segregate. This

has to be taken into account during mixing, discharging of the material from the mixer

into the truck and transportation. Another important factor is that any material

delivered ahead of the estimated delivery time could be segregated and unusable.

The comparison of the paste and mortar test results show that the incorporation of

sand does not affect the fluidity or saturation superplasticizer dosage in the 0-minute

tests but it seems to increase the loss of fluidity due to the absorption of water and

superplasticizer.

104
4.4.3 Loss of Fluidity of Concrete

To understand the change in flow behaviour of concrete with time, four mixes with

the corresponding saturation dosages of superplasticizer (corresponding to 0-minute

tests on pastes; see Section 3.5.2) were prepared. The slump was observed at intervals

of 30 minutes and the concrete was kept at rest in the pan mixer between

measurements; just before each slump test, the concrete was mixed for 15 seconds.

The results are shown in Figure 4.16, where it can be observed that the slump

decreases with time in all the combinations.

The loss of slump for SNF-S1 is 17% over 30 minutes and 54% over 60 minutes. The

loss of slump for SNF-D2 is the least compared to other superplasticizers, as seen in

Figure 4.16; it yields the lowest loss in slump of 18% and 26% over 30 and 60

minutes, respectively, reflecting its retarding nature. Slump losses of 54% over 30

minutes and 87% over 60 minutes are obtained in the case of the SMF based

superplasticizer, and slump losses of 57 % over 30 minutes and 82% over 60 minutes

for the PCE based superplasticizer. The SMF gives a higher loss of slump than the

PCE, as already seen in the case of mortar. This loss of slump with time in all cases

can be either due to cement hydration or drying due to water evaporation or

absorption of superplasticizer and water by the aggregates. The fines in the aggregates

also seem to increase the superplasticizer demand over the period of 60 minutes. This

indicates that in such cases, the saturation dosage determined at 0 minutes in the paste

will not be able to maintain the slump over a long period of time, say 60 minutes, in

concrete. When this occurs, it is recommended that the dosage of superplasticizer in

the concrete be taken as the saturation dosage obtained in the paste at 60 minutes.

105
180
SNF-S1
160 SNF-D2
SMF-S1
140 PCE-D1
120

Slump (mm)
100
80
60
40
20
0
0 20 40 60 80 100 120
Time (minutes)

Fig.4.16 Loss in slump of concrete with time

4.5 EFFECT OF SUPERPLASTICIZER ON SETTING OF CEMENT


PASTE AND CONCRETE

As already discussed, the superplasticizer slows down the hydration and dissolution of

ions, setting of superplasticized cement paste can be retarded (Torrents et al., 1998;

Roncero et al., 1999; Brooks et al., 2000). In this study, the Vicat apparatus was used

to determine the setting time of superplasticized cement paste. The cement paste was

prepared using a Hobart mixer, with a w/c of 0.35 and the saturation dosage of

superplasticizer (sp/c%) of each case. In Figure 4.17, the beginning and end of setting

(defined as the first incomplete Vicat penetration reading and the first zero penetration

reading are shown for each superplasticizer at saturation dosage). The graph shows

that retardation produced due to SNF-S1 and PCE-D1 is more than that of the SMF

based superplasticizer. This is because the higher molecular weight SMF will be

adsorbed more than the lower molecular weight SNF whereas, more of the SNF

remains in the solution and retards more than that of SMF. In the case of PCE-D1, the

retardation is because of the higher concentration of ionic functional groups

(sulphonic and carbonic) in the aqueous phase or due to shorter mean backbone chain

106
lengths in the PCE, in accordance with the reported results of Yamada et al. (2000).

The paste with the SNF with the retarder sets more slowly, as expected.

It can be observed that the time difference between the beginning and end of setting

varies approximately from 1 to 3 hrs. This shows that for the tested cement paste

systems, the superplasticizers lead to a delay in the beginning of setting due to delay

in the hydration of alite. However, it does not prolong the setting process because the

hydration is not affected further by the superplasticizer, after the beginning of alite

hydration (Yamada et al., 2000).

Control paste
40 SNF-S1
SNF-D2
Initial setting
Depth of penetration (mm)

SMF-S1
PCE-D1
30

20

10

0
200 400 600 800 1000 1200
Time (minutes)
Fig. 4.17 Vicat penetration measurements for pastes with different superplasticizers

The setting time of concrete was determined using a standard penetrometer as per IS

8142 (2002) as shown in Figure 4.18. The concrete was sieved through a 4.75 mm

mesh. The passing through mortar was tested, and the results obtained are shown in

Figure 4.19. The times at the penetration resistances of 3.43 N/mm2 and 26.97 N/mm2

are taken as the initial and final setting times of the concrete, respectively; the setting

times of paste and concrete at the saturation dosage of superplasticizer are shown in

Table 4.4, along with the paste flow time and slump values. The initial setting time

(IST) for the PCE based concrete was 8 hrs and final setting time was 10 hrs. The

107
SNF-S1 shows a similar trend whereas the SMF sets much earlier and the SNF with

retarder (SNF-D2) sets much later. However, except in the case of PCE-D1, the

difference between initial and final setting times (FST) is approximately 2 hours in

both the paste and concrete. It should be noted that both the Vicat penetration test on

paste and the penetration resistance test on mortar measure setting through the

determination of the degree of stiffening. Since the initial fluidity of the paste or

mortar also affects the stiffness and its evolution, the comparison of the different

results is not straightforward, especially since each of the pastes has a different initial

consistency. However, it appears that the superplasticizer that causes more retardation

of setting in the paste also tends to retard the setting of the concrete having the same

paste proportions.

Fig.4.18 Penetration test on concrete; (a) Pocket penetrometer,


(b) Penetration test on specimen (c) Impressions
made on the surface during measurements

30 SNF-S1
Final setting SNF-D2
25 SMF-S1
PCE-D1
Pressure (N/mm )
2

20

15

10

5 Initial setting

0
300 350 400 450 500 550 600 650
Time (minutes)
Fig. 4.19 Setting of concrete at saturation dosage of superplasticizer
108
Table 4.4 Initial and final setting time of paste and concrete

Type of Paste Concrete


superplasticizer Flow time IST FST Slump IST FST
(sec) (hrs) (hrs) (mm) (hrs) (hrs)
Control mix * 3.6 5.2 0 * *
SNF-S1 13 9.5 11.9 40 7.9 9.6
SNF-D2 17 16.3 18.7 90 8.2 10.5
SMF-S1 16 6.6 8.3 45 6.7 8.8
PCE-D1 12 10.5 14 73 8.0 9.9

4.6 ANALYSIS OF SETTING BEHAVIOUR OF SUPERPLASTICIZED


CEMENT PASTE USING ELECTRICAL RESISTIVITY METHOD

4.6.1 Background

The setting of cement paste and concrete is usually determined through penetration

resistance, as seen in the previous section. Electrical property measurements are being

proposed as an alternative to the penetration measurements for studying the setting

and hardening processes within the portland cement pastes over the initial 24 hours of

hydration through continuous real time monitoring (McCarter and Curran, 1984;

McCarter and Tran, 1996; McCarter et al., 1999, 2003). Calleja (1952a, 1952b) has

used electrical resistance measurements for estimating the initial and final setting time

of hydraulic materials. Torrents et al. (1998) and Roncero (2000) have used AC

impedance spectroscopy over a frequency range of 1 kHz to1 MHz for understanding

the influence of superplasticizers on the electrical resistivity of cement paste. More

recently, Xiao et al. (2007) have used electrical resistance methods to compare the

effect of superplasticizers at the saturation dosage and have correlated resistivity to

strength gain and setting time. The previous studies of Srinath (2007) on electrical

conductivity of cement mortar at the frequencies of 1, 10 and 100 kHz show that a

lower conductivity is obtained at 1 kHz whereas the values are similar at 10 and 100

kHz, indicating that frequencies of 10 kHz and above are adequate for the

109
conductivity measurements on fresh pastes since electrode polarization effects are

reduced to negligible proportions (McCarter and Tran, 1996).

The objective of this study is to understand the influence of superplasticizers on the

different stages of setting of cement paste based on the electrical conductivity of

paste. Tests were conducted with pure cement paste and with 4 types of

superplasticizers. The results are compared with standard Vicat test measurements.

4.6.2 Experimental Procedure

Aplab MT 4080A high accuracy LCR meter (hand held AC/DC impedance

measurement instrument), which can measure the inductance, capacitance and

resistance with imposed currents with frequencies of up to 100 kHz, was used in the

present study. A mould with internal dimensions of 50 × 50 × 50 mm was made with

polycarbonate plates of 2 mm thickness. This sample size was considered large

enough to obtain a representative bulk electrical measurement on paste (following

McCarter and Tran, 1996). Stainless steel plates of 50×50 mm were placed on

opposite walls of the mould. The cement paste with the saturation dosage of

superplasticizer was prepared, as explained in Section 3.4. The sample (cement paste)

was placed inside the mould and the steel plates were connected to the H and L ports

of the LCR meter to serve as the electrodes (Figure 4.20). Impedance values were

determined at the frequency of 10 kHz and the voltage of 1 Vrms over a period of 24

hours. The resistance values were converted to electrical resistivity using equation

(4.1).

RA
ρ= ohm metres (4.1)
d
where d is the length of the specimen, A is the cross sectional area of specimen, ρ is

the electrical resistivity. Due to sample preparation and setting-up procedures, test

110
measurements were taken 15 minutes after mixing. So the zero minute reading

presented in the graph in this section corresponds actually to 15 minutes after casting.

The resistance values were taken subsequently every 15 minutes over a test period of

24 hours. The top surface of the sample was covered with a polycarbonate plate to

avoid evaporation. All tests were carried out in ambient conditions.

LCR meter

Fig. 4.20 Comparative study on setting behaviour of


paste with Vicat apparatus and LCR meter

4.6.3 Results and Discussions

The conductivity and its derivative for pure cement paste and the superplasticized

pastes are plotted as a function of time in Figure 4.21. The variation in conductivity

can be correlated with the different stages of hydration of cement paste. Stage 1

corresponds to the initial stage in which the conductivity increases due to the

dissolution of ions, in all the cases. However, the increase in conductivity is more for

the pure cement paste than that of superplasticized pastes. This confirms that the

presence of superplasticizer prevents further dissolution of ions, which slows down

the increase in conductivity in the initial stage. Stage 2 corresponds to the dormant

stage in which the conductivity remains almost constant. In the case of pure cement

111
paste, the dormant stage is caused due to the supersaturation of solution and

consequent inhibition of dissolution of ions. The superplasticizer appears to mainly

extend this dormant stage and it could be attributed to; (1) the lowering of the rate of

chemical reactions further by inhibiting the dissolution of ions due to the presence of

superplasticizers in solution; and (2) the formation of an adsorption layer on the

cement particle and the formation of complexes with Ca2+. In stage 3, the conductivity

decreases as the paste sets and the resistance increases due to the decrease in the

mobility of Ca2+, SO42- and OH- ions. The study confirms that the addition of

superplasticizer generally extends the dormant stage of hydration process, retarding

the onset of acceleration stage, in accordance with the results of earlier studies.

(Torrents et al., 1998; Roncero, 2000; Xiao et al., 2007). However, the evolution of

the derivative of conductivity clearly reflects the rate of chemical reactions only in the

case of pure cement paste.

1.8 0.015

Conductivity
1.6 0.010
derivative of conductivity
Derivative of conductivity
Conductivity (siemens)

0.005
1.4
0.000
1.2 2
1 3
-0.005
1.0
-0.010

0.8 -0.015

0 50 100 150 200 250 300 350


Time (minutes)

(a) Pure cement paste with w/c = 0.35

112
2.0
Conductivity 0.003
1 derivative of conductivity

Derivative of conductivity
1.6
0.002
Conductivity (siemens)
2
1.2 0.001
3

0.8 0.000

-0.001
0.4
-0.002
0.0
0 500 1000 1500 2000
Time (minutes)

(b) Cement paste with LS-C1

2.0
1 2
0.002
Conductivity

Derivative of conductivity
1.5 Derivative of conductivity 0.001
Conductivity (siemens)

3
0.000
1.0
-0.001

0.5 -0.002

-0.003
0.0
0 500 1000 1500 2000
Time (minutes)
(c) Cement paste with SNF-S1

113
2.0
1 2 0.0010
Conductivity
Derivative of conductivity
0.0005

Derivative of Conductivity
1.5

Conductivity (Siemens)
3 0.0000

1.0 -0.0005

-0.0010
0.5
-0.0015

-0.0020
0.0
0 500 1000 1500 2000
Time (minutes)

(d) Cement paste with SMF-S1

0.8 Conductivity 0.0015


Derivative of conductivity
1 0.0010

Derivative of conductivity
Conductivity (siemens)

2
0.6 3
0.0005

0.4 0.0000

-0.0005
0.2
-0.0010

-0.0015
0 400 800 1200
Time (minutes)

(e) Cement paste with PCE-D1

Fig. 4.21 Evolution of conductivity for different superplasticizers

The conductivity curves for the different pastes, along with the beginning and end of

setting (IST and FST) obtained from Vicat penetration tests, are shown in Figure 4.22.

Compared to pure cement paste, the initial conductivity is much lower for the PCE

and similar for the other superplasticizers. As discussed earlier, the results confirm

that the dormant stage lasts longer for the superplasticized pastes.

114
The comparison of the evolution of the conductivity with the Vicat test results shows

that for pure cement paste, as well as paste with LS, SNF and PCE based

superplasticizers, both the IST and FST fall in the acceleration stage. For the SMF-

based superplasticizer, setting begins in the dormant stage whereas the final setting

time falls in the acceleration stage. This is in accordance with the penetration test

results of both paste and concrete where the retardation of setting was seen to be the

least for the SMF-based superplasticizer. It is observed that the superplasticizer that

extends the dormant stage of hydration most, yields the highest IST and FST in

concrete. Hence, the test results obtained from electrical resistivity method, which is

closely related to the chemical activity in the paste, indicate that the setting time is in

the following order: Cement paste < Paste with SMF< Paste with SNF< Paste with

PCE< Paste with LS.

2.0
Control paste
Conductivity (siemens)

LS-C1
1.6 SNF-S1
SMF-S1
IST PCE-D1
1.2

FST
0.8

0.4

0.0
0 500 1000 1500 2000
Time (minutes)
Figure 4.22 Electrical conductivity graphs for different pastes
along with beginning and end of setting defined by
Vicat penetration measurements

Another point to be noted is that even though the tests are conducted at the saturation

dosage for different superplasticizers, the maximum conductivity values varies with

115
the type of superplasticizer. This indicates that the concentration of ions in the

aqueous solution is also modified by the type of superplasticizer, especially in the

case of PCE-based superplasticizer. This further confirms that the influence of the

superplasticizer on setting, as well as on fluidity, varies with its mechanism of action,

as explained in Chapter 2. It can be observed from Figure 4.22 that the time difference

between the beginning and end of setting is approximately the same in all cases.

Hence, it can be concluded that for the present cement paste system, the

superplasticizers considered delay the beginning of setting but do not prolong the

setting process.

4.7 EFFECT OF SUPERPLASTICIZER ON COMPRESSIVE STRENGTH


OF CONCRETE

The compressive strengths of concretes incorporating different types of

superplasticizers are shown in Table 4.5. It can be observed that the strength of

superplasticized concrete at 28 days is generally more than that of the reference

concrete, as also observed by other researchers such as Dhir and Yap (1983), Seung-

Bum (1999), Ramachandran (2002) and Kapalko (2006). It can be observed that the

concretes with the saturation dosages of the superplasticizer exhibit slightly higher

strengths than the reference concrete, probably due to better compaction.

Nevertheless, superplasticizer dosages much above saturation lead to bleeding and

immediate stiffening, which cause strength reduction.

116
Table 4.5 Compressive strength of concrete for different combinations

Type of Dosage of Saturation Slump Compressive strength


superplasticizer superplasticizer dosage (mm) 3 days 7 days 28 days
(sp/c %) of sp N/mm2 N/mm2 N/mm2
(sp/c %)
0 (Reference) 0 39 45 56
0.1 0 38 47 56
SNF-S1 0.2 0.20 25 39 49 58
0.3 195 34 54 69
0.4 bleeding 31 40 50
0.1 10 47 57 62
0.2 67 46 57 64
SNF-D2 0.25 0.23 100 40 55 63
0.3 130 43 59 65
0.4 165 37 56 67
0.1 20 39 - 57
0.2 32 45 55 69
SMF-S1 0.3 0.23 70 40 53 67
0.5 158 41 52 73
0.6 bleeding 40 52 60
1 bleeding 45 - 53
0.05 35 36 46 58
0.1 120 42 47 54
PCE-D1 0.2 0.07 180 46 55 67
0.3 bleeding 34 41 45

4.8 SUMMARY

An extensive study on paste, mortar and concrete aimed at establishing correlation

between their behaviour has led to the following conclusions:

• The saturation dosages obtained for paste and mortars are comparable whereas

a dosage slightly higher than the saturation dosage of paste was required for

adequate workability in concrete due to the adsorption of some

superplasticizer by the fines present in the coarse aggregate used here. The

comparison of the cost of superplasticized concretes shows that a low-cost

superplasticizer need not always result in cost-effective concrete, as the dosage

of the superplasticizer for the optimum fluidity is also an important factor.

117
• The study of the flow behaviour of pastes at 60 minutes after mixing shows

that loss of fluidity occurs for all the superplasticizers and dosages considered.

If the fluidity of superplasticized paste at 60 minutes or at any other duration is

more important than the immediate fluidity, the saturation dosage

corresponding to that duration should, therefore be used in the mix design of

concrete rather than arbitrarily overdosing the superplasticizer. In the case of

mortar, the addition of sand results in higher loss of fluidity than in the paste,

probably because of the absorption of superplasticizers by the fine aggregates.

Generally, it was found that the SMF-S1 and PCE-D1 led to more slump loss

than the SNF-S1 based concrete as observed in the case of corresponding

paste. Also, as expected, the SNF with retarder was effective in limiting the

slump loss of concrete.

• The Vicat penetration test results shows that the setting behaviour depends on

the type of superplasticizer, and that the setting time is higher for the

lignosulphonates and polycarboxylate than the sulphonated naphthalene-and

melamine-based superplasticizers. A comparison of the penetration tests on

paste and concrete indicates that the trend in the setting with different types of

superplasticizers is almost same in both cases, and hence the test of paste can

be used as a guideline in the selection for the superplasticizers for concrete in

terms of setting time. Also, it is confirmed through the study that

superplasticizers tend to retard the beginning of setting and do not necessarily

prolong the setting process, i.e., the difference between the initial and final

setting times are not affected.

• Electrical conductivity measurements on the superplasticized pastes have been

used to characterize the setting behaviour, along with Vicat penetration

measurements. It has been observed that the superplasticizer delays the initial
118
setting without prolonging the setting process, which is mainly manifested as

an increase in the duration of the dormant stage.

• The study shows that the compressive strength can increase slightly due to the

addition of superplasticizers, probably due to better compaction, especially at

the saturation dosage of superplasticizer.

Generally, it can be concluded that the selection of the type and dosage of

superplasticizer in concrete for high performance can be based on the test results from

paste. This could lead to substantial savings in material and effort within the mix

optimization process.

119
CHAPTER 5

EFFECT OF SUPERPLASTICIZER ON
PROPERTIES OF HARDENED CEMENT PASTE

5.1 GENERAL

The study of the flow behaviour of superplasticized cements paste and its correlation

with concrete shows that the fluidity and flow retention depend on the type and

dosage of superplasticizers. The purpose of adding superplasticizers is to improve the

fresh state properties of concrete without adversely affecting the hardened state

properties. However, the literature shows that cement-superplasticizer incompatibility

results in an adverse effect on the rate of hydration and hydrated products (Jolicoeur

and Simard, 1998; Hanehara and Yamada, 2008; Agarwal et al., 2000; Prince et al.,

2002; Nkinamubanzi and Aitcin, 2004; Bedard and Mailvaganam, 2005).

The objective of this chapter is to have an understanding on the influence of type of

superplasticizers on the hydration behaviour and on the properties of hardened

concrete. Indirect techniques like differential thermal analysis/ thermogravimetric/

differential thermogravimetric (DTA/TG/DTG) and X-ray diffraction study (XRD)

give information about the average characteristics of the microstructure whereas

direct methods like scanning electron microscopy (SEM) provide information about

the way in which component phases are arranged in the microstructure (Scrivener,

1989; Ramachandran and Beaudoin, 2001).

The influence of superplasticizers on the hardened cement paste and more precisely,

on the hydration processes and the microstructural development is explained in this

Chapter. The cement paste phases are analyzed using XRD, micrographs, thermal

analysis and nuclear magnetic resonance (Si NMR) for pure cement paste as well as

for superplasticized pastes at saturation dosages. The complementary use of these


approaches is intended to provide a complete picture of the hydration processes, with

respect to the formation of new phases, the consumption of the anhydrous phases as

well as the formation of C-S-H and C-H.

5.2 X-RAY DIFFRACTION (XRD) STUDY

5.2.1 Background

X-ray diffraction study helps to understand the influence of the incorporation of the

superplasticizer on the hydration process of cement paste with respect to the

consumption of crystalline anhydrous phases, as well as through the identification of

crystalline products of hydration. The basic principle of XRD is that every crystalline

substance gives a specific pattern; the same substance always gives the same pattern

and in a mixture of substances each produces its pattern independently of others.

When a beam of X-rays falls on a crystalline material, it is diffracted by the incidence

angle given by Bragg’s law:

2 d sinθ = n λ (5.1)

where d is the distance between the crystallographic planes, λ is the wavelength of the

X-ray, n is an integer and θ is the diffraction angle.

XRD is used to identify the polycrystalline phases of hardened cement paste and

superplasticized cement paste through the recognition of the X-ray patterns that are

unique for each of the crystalline phases. So the technique allows the detection of

ettringite (AFt) and portlandite (C-H) along with the consumption of anhydrous

phases of the cement (gypsum, C3S, C2S, C3A, C4AF, SiO2 and CaCO3).

Superplasticizers generally affect the hydration by the reduced formation of Ca(OH)2

and ettringite in the initial ages of curing, which can be identified through XRD

121
(Singh et al., 1992; Mollah et al., 1995; Roncero, 2000; Prince et al., 2002; Roncero et

al., 2002; Prince et al., 2003).

5.2.2 Materials Used and Sample Preparation

Cement C1 of 53 grade and 6 types of superplasticizers with properties as given in

Section 3.2 along with a reference paste at several ages like half setting

(corresponding to 20 mm Vicat penetration), 3, 7 and 28 days were tested. These ages

are selected as reference ages so as to compare the effect of superplasticizers on the

rate of hydration. The pastes were prepared with the saturation dosage of

superplasticizers (determined from the Marsh cone test) and the mixing method

adopted was the same as selected from Section 3.4. The samples were cured under

water and the cement paste samples of 50mm×50mm×10mm were cut and kept in

acetone to prevent the hydration after the required age. Just before each test, the

samples were dried in an oven at 60°C and the surface was polished to remove

undulations. To account for the variations in the preparation of sample and the

instrument intensity, a standard sample of aluminium foil was placed over the

specimen and the normalized intensity was used for comparison purposes.

5.2.3 Experimental Procedure

The XRD analysis was performed with a Bruker D8 Discover powder diffractometer.

In this apparatus, the Cu Kα (λ = 1.5418Å) radiation is generated in a Cu tube at

35kV, 25 mA. The tests were performed over a Bragg angle (2θ) range of 5-70° with

a scan speed of 1 sec per step on the samples. The ratio of the intensity of sample to

the intensity of aluminium at a particular Bragg angle is taken for the comparison of

crystalline phases of superplasticized pastes.

122
5.2.4 Results and discussions

The XRD pattern of cement paste and superplasticized pastes for half setting, 3, 7 and

28 days are shown in Figure 5.1 (a-g). The relative evolution with age of the identified

crystalline phases like C3S, gypsum, ettringite and portlandite can be compared with

the intensity. In the case of pure cement paste, amount of unhydrated phases decrease

with an increase in the age of curing and C-H increases with the age of curing, as

shown in the X-ray diffractograms. Ettringite is present at half setting and disappears

after 3 days as expected. The peak of calcite (CaCO3) indicates the occurrence of

carbonation in all the samples.

The XRD pattern of superplasticized paste shows some qualitative differences in the

hydration rate due to the incorporation of superplasticizers which is supported by the

reported results (Mollah et al., 1995, Roncero, 2000; Prince et al., 2003). This

qualitative difference varies with different superplasticizers. The delay of C3A

hydration in the presence of superplasticizers is obvious through the absence of

ettringite at half setting in the XRD pattern of superplasticized pastes. It affects the

setting and hardening due to the formation of electrical double layer along with “the diffuse

ion swarm” of superplasticizers on Ca2+ ions preventing the hydration (Mollah et al., 1995).

11 C-H
C-H
C-H C-H Al
10 C3S 28 days
9
8
C-H Calcite C-H Al
7 C3S C-H C-H
ISample/IAl

6 7 days
5
C-H Calcite C-H Al
4 C-H
3 days
3
2 C3S
C-H Al
C-H AFCalcite C-H C S Half setting
1 t 3

0
10 20 30 40 50 60 70 80
2θ (degrees)
(a) Cement paste with w/c 0.35
123
C-H
10 Al 28 days

8 C-H
7 days

Al
ISample/ I 6
C-H C-H 3 days
C3S Al
4
Gypsum
Calcite Calcite C3S
2 Half setting

0
10 20 30 40 50 60 70 80
2θ (degrees)

(b) LS-C1 based paste

C-H C-H Al
10 C-H
28 days

8
C-H Calcite C-H Al
C-H
7 days
ISample/IAl

Calcite
C-H C-H Al
4
3 days
C3S
Calcite C-H
2 Gypsum Al C3S
Al
Half setting
0
10 20 30 40 50 60 70 80
2θ (degrees)

(c) SNF-S1 based paste

124
10 C-H C-H
28 days
9
8
ISample/ IAl 7 7 days
6
5 C-H
4 Al
AFt
3 3 days
CS
2 Gypsum Calcite 3 C-H
C-H C-H C3S Half setting
1
0
10 20 30 40 50 60 70 80
2θ (degrees)

(d) SNF-D2 based paste

C-H C-H
10

28 days
8
ISample/IAl

6 C-H
Calcite
7 days
4
C-H SiO C-H
2 C3S
Al 3 days
2 AFt C-H C-H
Gypsum
C3S Half setting
0
10 20 30 40 50 60 70 80
2θ(degrees)

(e) SNF-C2 based paste

125
10 C-H C-H 28 days
9
8
C-H C-H
7 7 days
6
ISample/IAl
5
SiO C-H
4 AFt 2 3 days
3
2 Half setting
CS Al
1 Calcite 3 C-H C-H C3S
0
10 20 30 40 50 60 70 80
2θ(degrees)

(f) SMF-S1 based paste

C-H Al
10 28 days

8
C-H Al
7 days
ISample/IAl

C-H C-H Al
4 3 days
CalciteC3S
C-H
Gypsum Calcite C3S
2
Half setting

0
10 20 30 40 50 60 70 80

(g) PCE-D1 based paste

Fig. 5.1 (a-g) X-ray diffractogram of cement paste and superplasticized cement paste

The trends of the evolution of the crystalline phases of pure cement paste and

superplasticized cement paste at different ages are summarized in Table 5.1. The

arrows indicate the relative change in the amount of the crystalline phases observed.

The amount of unhydrated compounds like C3S/C2S decreases with an increase in the
126
age of curing. Even though the time of half setting is different for the studied pastes,

major qualitative difference could not be observed between the superplasticized

pastes. The presence of portlandite in the superplasticized cement pastes at half setting

indicates that the formation of portlandite is not affected by the superplasticizers.

Generally, the amount of portlandite increases with age of curing, as expected.

However, anomaly is observed with some types of superplasticizers.

Table 5.1 Relative evolution of the crystalline phases with age

Type of C3S,
Age SiO2 Gypsum Ettringite Portlandite
paste C2S
Half setting P P A P P
3 days P P A A P
Reference 7 days P P A A P
28 days P P A A P
Half setting P P P A P
3 days P P A P P
LS-C1 7 days P P P A P
28 days P P A A P
Half setting P P P A P
3 days P P A P P
SNF-S1 7 days P P A P P
28 days P P A P P
Half setting P A P A P
3 days P P A P P
SNF-D2 7 days P P A P P
28 days P P A P P
Half setting P P P A P
3 days P P A P P
SMF-S1 7 days P P A P P
28 days P P A P P
Half setting P A P A P
3 days P A A P P
SNF-C2 7 days P A A A P
28 days P P A P P
Half setting P A P A P
3 days P P A P P
PCE-D1 7 days P P P P P
28 days P P A P P
P - present; A – absent
Increase or decrease of the phase, respectively, with reference to the previous
age

127
The comparison of the diffractograms of pure cement paste with superplasticized

pastes at half setting indicates that superplasticizer mainly affects the hydration of

C3A and consequently the formation of ettringite; it can be observed that no gypsum

is present in the pure cement paste at half setting due to its conversion to ettringite by

reacting with C3A. However, in the superplasticized pastes, no traces of ettringite

were observed at half setting; instead, gypsum is present. This clearly shows that the

setting is retarded due to the addition of superplasticizers because it inhibits the

hydration reaction. The possible reasons for set retardation have been already

explained in Chapter 2. The results also support the data from the setting time tests

reported in Chapter 4.

Also, traces of ettringite are present upto 28 days in all the superplasticized pastes

except for the paste with LS-C1, in which it is present upto 3 days only. However, the

peak intensity of ettringite decreases between 3 days and 7 days and again increases

after 7 days in the case of SMF-S1 and PCE-D1. This is in accordance with the

reported results (Prince et al., 2003) that the addition of superplasticizer affects the

crystallization of primary ettringite and when no more superplasticizer is available to

block the natural growth of ettringite, it reappears after 28 days. The diffraction

pattern does not show any traces of AFm or secondary gypsum phases which indicate

that phase reversal (see Chapter 2) has not occurred. In short, from the diffractograms

observed in the particular cementitious system under study, it is seen that there is very

strong interaction between the superplasticizer and ettringite at early ages.

5.3 SCANNING ELECTRON MICROSCOPY (SEM)

5.3.1 Background

Backscattering electron (BSE) imaging is generally used to identify the different

phases of hardened cement paste and to understand the development of the


128
microstructure of cement paste. Scrivener (1989) has used the technique to distinguish

between the dense and porous phases in the microstructure of same paste. The basic

principle is that the electrons of the incident beam, rebound from the surface of the

specimen depending on the composition of the phases. The intensity of the BSE signal

is a function of the average atomic number of the sample and the contrast observed is

based on the atomic number of phases (Famy et al., 2002).

The unhydrated particles have higher molecular weight with higher back scattering

intensity than hydrated particles (Zhao and Darwin, 1992; Puertas et al., 2005). In

such studies it has been shown that C4AF has the highest intensity, C-H is brighter

than C-S-H and that the pores appear black.

A comparative study of the microstructure of different superplasticized cement pastes

and reference pastes is conducted here to understand the influence of different types

of superplasticizers on the hydration process and consequently on the development of

microstructure.

5.3.2 Sample Preparation

The quality of the BSE image depends on the sample surface. A highly polished

surface is required for BSE microscopy (Scrivener, 1989). The sample preparation

included the preparation of cement paste samples, curing for the prescribed ages,

cutting, grinding and polishing. The superplasticized cement paste cubes with four

families of superplasticizers along with the reference paste cubes were prepared as in

the case of the XRD study. The cubes were cured under water for 3, 7 and 28 days.

After the period of curing, the samples were put in acetone to remove the pore water

by the solvent replacement method to prevent further hydration as in Roncero (2000).

Samples of 10mm×10mm×10mm were cut and carefully ground by hand at moderate

129
pressure on a middle-speed lap wheel with p800, p1000, p1200, p1600 and p2000

sand papers. Afterwards, lap wheel with 6-12, 3-6 and ½ -1 μm diamond paste was

used to get a fine polished surface.

5.3.3 Comparison of the Microstructure of Superplasticized Paste and


Reference Paste

The microstructural study was performed on pastes with no admixture (i.e., reference

paste) and those containing saturation dosages of superplasticizers. The BSE

micrographs of reference paste and paste with LS-C1, SMF-S1, PCE-D1 and SNF-S1

based superplasticizer are shown in Figure 5.2 (a-i). The microstructure of cement

paste shows the bright grains of unhydrated cement, the intermediate gray of C-S-H

and slightly brighter gray (shown as red for easy identification) of calcium hydroxide,

along with dark pores. As the age of curing increases up to 28 days (see Figure 5.2 a

and b), the amount of unhydrated particles decreases and a uniform distribution of

hydrated particles is identified in the picture.

The BSE morphological analysis of superplasticized pastes showed no substantial

difference in either the form or the texture of the different hydration products in pastes

with and without admixtures. However, superplasticized pastes show more amount of

unhydrated particles at 3 days than reference paste confirming the retardation of

hydration in early ages due to the incorporation of superplasticizers as already

discussed in Chapters 2 and 3. Even though, the micrographs are obtained at the

saturation dosage of superplasticizers, the amount of unhydrated particles is different

for different pastes. In the case of LS based superplasticizer, more amount of C-S-H

can be observed. The amount of unhydrated products is more for the paste with the

SMF at 3 days compared to that with the PCE. More unhydrated particles than in the

pastes with SMF and PCE can be observed in the paste with SNF at 28 days. Also, the

130
pastes with the SMF micrograph shows more C-H at 28 days, compared to those of

the other superplasticizers.

A comparison of the microstructure of the pastes with SMF and PCE shows that a

more uniform microstructure is produced in the latter case. The number of pores in the

superplasticized paste is more than in the reference paste and it is higher in the paste

with PCE. The pores are smaller when PCE is added as reported by Sakai et al.

(2006). Generally, the increase in strength due to the addition of superplasticizer can

be attributed to making the microstructure more dense, uniform and compact

(Uchikawa, et al., 1995). The micrographs shown in Fig. 5.2 also support the

modification of the microstructure with uniform distribution of C-S-H and C-H

C-S-H

Unhydrated
C-H

(a) Reference paste, 3 days (b) Reference paste, 28 days

131
(c) LS-C1 based paste, 3 days (d) LS-C1 based paste, 28 days

(e) SMF-S1 based paste, 3 days (f) SMF-S1 based paste, 28 days

(g) PCE-D1 based paste, 3 days (h) PCE-D1 based paste, 28 days

132
(i) SNF-S1 based paste, 28 days

Fig. 5.2 (a-i) BSE (250X) micrograph of pure cement paste and
superplasticized cement paste

Energy dispersive X-ray analysis (EDAX) was also done for the selected samples to

find out the elemental composition of different phases. It is a technique used for

identifying the elemental composition of the specimen, or an area of interest thereof.

Each of the peaks in an EDAX spectrum is unique to an atom, and therefore

corresponds to a single element. The higher a peak in a spectrum, the more

concentrated the element is in the specimen. An EDAX spectrum not only identifies

the element corresponding to each of its peaks, but the type of X-ray to which it

corresponds as K-alpha peak or K-beta peak.

Figure 5.3 (a-c) shows an example of the BSEM micrograph along with the EDAX

spectrum of a C-S-H particle for the paste with the lignosulphonate. The elemental

composition of Ca and Si in the EDAX confirms the presence of C-S-H in the

micrograph. The Ca/Si ratio can also be determined for validating the identified

phases through the microstructure. A higher value of Ca/Si ratio for C-S-H in some

cases of superplasticized pastes shows that it is more amorphous than the pure cement

133
paste (Taylor, 1997). However, the summary of the results is not reported in this study

as the numbers of studied samples are limited.

Element Wt% At%


OK 40.83 61.13
AlK 01.76 01.56
SiK 12.88 10.99
KK 00.84 00.51
CaK 41.86 25.02
FeK 01.83 00.78

(a) EDAX spectrum of different elements (b) Elemental composition


in the selected spot corresponding to (Kα) radiation

(c ) Micrograph of C-S-H particle

Fig. 5.3 (a-c) BSE micrograph along with EDAX for a dark grey
(C-S-H) particle in the LS based paste

134
5.4 THERMAL (DTA/TG/ DTG) ANALYSIS

5.4.1 Background

Differential thermal analysis (DTA) along with thermogravimetric (TG) and

differential thermogravimetric analysis (DTG) can be used to understand the

hydration reactions of superplasticized cement paste (Prince et al., 2002). The basic

principle is that the material is subjected to continuous heating at a uniform rate and

the loss in mass is studied. In DTA, the difference in temperature (ΔT) between the

sample and reference material such as α-Al2O3 is recorded while both are subjected to

the same heating program (Ramachandran and Beaudoin, 2001).

TGA measures the loss in weight as the temperature of the substance is raised at a

uniform rate. The reactions occurring at the heating process are responsible for the

changes in weight. Knowing that certain reactions occur at specific temperatures

enables the identification of the constituents of the sample. TGA curves are usually

used to measure the weight loss corresponding to thermal decomposition whereas

DTG analysis helps locate the temperature peak due to the decomposition of different

phases in cement pastes.

5.4.2 Result and Discussions

The analyses were conducted with a NETZSCH STA 409 C/CD, in nitrogen

atmosphere, within a temperature range of 20-1400 ºC, at a uniform rate of 10 K/min.

The samples were obtained by powdering the prepared paste fragments and sieving

through 70 μm sieves.

Figure 5.4 (a-c) shows the DTA analysis of superplasticized paste and reference paste

after 3 days of curing. The small endothermic peak at 94.5ºC for pure cement paste

corresponds to dehydration of the non-evaporable water from the C-S-H gel (Prince et

135
al., 2002, 2003). In the case of pastes with SMF and PCE, an endothermic peak at

about 160ºC shows the decomposition of ettringite, gypsum and C-S-H gel (Prince et

al., 2003). The increase in the intensity of this effect with time is indicative of

increased formation of C-S-H with time. The endothermic peak in the range of 425-

550ºC shows the decomposition of C-H, in all the three cases (Tzouvalas et al., 2005;

Zhang, 2007).

o
1 peak at 94.5 C
exo 0
o
peak at 459.3 C
-1
DTA mw/mg

-2
-3
-4
-5
-6
-7
0 200 400 600 800 1000 1200
o
Temperature ( C)
(a) Reference paste, 3 days

0
0
Peak:164.1 C
-1 Peak:437.3 C
0
DTA (mw/mg)

-2

-3

-4

-5

-6
0 200 400 600 800 1000 1200
o
Temperature ( C)
(b) Paste with SMF-S1, 3 days

136
0.0 peak at 161.1
exo-0.5
-1.0 peak at 442

DTA mw/mg
-1.5
-2.0
-2.5
-3.0
-3.5
-4.0
0 200 400 600 800 1000 1200
o
temperature ( C)

(c) Paste with PCE-D1, 3 days


Fig.5.4 DTA analysis of cement paste and superplasticized paste at 3 days

Figure 5.5 (a-d) shows the TG curves of the different pastes at 3 days. The total loss in

mass represents the degree of hydration of the pastes. The quantity of calcium

hydroxide is calculated from the weight loss at the temperature around 450-550ºC.

The loss in mass for superplasticized paste is less than that of pure cement paste at 3

days showing that the formation of C-H is slightly retarded at early ages due to the

presence of superplasticizers. This is in accordance with the reported results of

Ramachandran et al. (1995) and Puertas et al. (2005). The mass loss is small and

within the ranges of the reported results (Prince et al., 2002; Prince et al., 2003;

Puertas et al., 2005).

137
100
o
Inflection 78.8 C

98
Mass change -7.5%

mass %
96

Mass change -2.1%


94
o
Inflection 453.8 C
o
Inflection 646.4 C
92

0 200 400 600 800


o
Temperature ( C)

(a) Reference paste, 3 days

100

98 Inflection at 88.8 C
o

Mass change -7%


96
mass %

Total Mass change -10%


94

o Mass change -1.53%


92 Inflection 434.5 C

Mass change -1.4%


90
o
Inflection 645 C

88
0 200 400 600 800 1000 1200
o
Tempeature ( C)

(b) Paste with SMF-S1, 3 days

138
100
98 Inflection 81.9 C
o

96 Mass change -7.7%


Total mass change -14%
94

mass % 92
90 o
Mass change -2%
Inflection 438.3 C
88 Mass change -1%

86 o
Inflection 642.4 C

84
0 200 400 600 800 1000 1200 1400
o
Temperature ( C)
(c) Paste with PCE-D1, 3 days

100
o
Inflection 100.7 C
98
Mass change -5.3%

96 Total Mass change -9.4%


mass %

94 o
Mass change -1.39%
Inflection 445.8 C

Inflection 685.2 C Mass change -2.1%


o

92
Mass change -0.057%

o
90 Inflection 1077.1 C

0 200 400 600 800 1000 1200 1400


o
Temperature ( C)

(d) Paste with SNF-C2, 3 days

Fig. 5.5 (a-d) TG and DTG analysis of pastes at 3 days

The results of the TG/DTG analysis for different pastes at 28 days are shown in

Figure 5.6 (a-d). The total mass loss increases in all cases with age of curing. The

mass loss corresponding to 600-800 ºC for the pastes at 3 and 28 days, is generally

139
interpreted as partly due to carbonation and partly to the final stages of dehydration of

C-S-H and the hydrated aluminate phases (Taylor, 1997). However, the XRD results

confirmed the presence of calcite; hence, this loss is mainly due to the dehydration of

C-S-H, hydrated aluminate phases and calcites.

100
o
Inflection 82.3 C

Mass change -8.7%


95

Total mass change -18%


mass %

Mass change -2%


90 o
Inflection 444.9 C

Mass change -6.4%


85 Inflection 725.5 C
o

80
0 200 400 600 800 1000 1200 1400
o
Temperature ( C)
(a) Reference paste, 28 days
100
o
98 Inflection 101.6 C

96 Mass change -9.3%


94
mass %

Total Mass change -17%


92
90
88 o
Inflection 454.6 C Mass change -1.93%

86 o
Mass change -2%
Inflection 695.3 C
Mass change -0.8%
84 o
Inflection 909.8 C
82
0 200 400 600 800 1000 1200 1400
o
Temperature ( C)
(b) Paste with SMF-S1, 28 days

140
100
o
98 Inflection 98.8 C

96
Mass change -10%
94
Total Mass change -17%

mass %
92
90
88 o Mass change -2.2%
Inflection 453.6 C
86 Mass change -2%
o
Inflection 691.6 C
Mass change -0.85%
84
o
Inflection 909.8 C
82
0 200 400 600 800 1000 1200 1400
o
Temperature ( C)

(c) Paste with PCE-D1, 28 days

100
98 o
Inflection 97.6 C
96 Mass change -10%

94
mass %

92 Total Mass change -15%

90
o
Inflection 451.9 C Mass change -2.1%
88
Mass change -1.1%
86 o Mass change -0.58%
Inflection 681.1 C
o
Inflection 879.7 C
84

0 200 400 600 800 1000 1200 1400


o
Temperature ( C)

(d) Paste with SNF-C2, 28 days

Fig. 5.6 TG and DTG analysis of pastes at 28 days

Table 5.3 gives the summary of total mass loss, as well as the mass loss corresponding

to the decomposition of C-H at the temperature of around 450-550 ºC. It is observed

that the total mass loss increases with the age in all the cases whereas the mass loss

corresponding to C-H decomposition is comparable at 3 days and 28 days. This shows


141
that the superplasticizers retards the hydration mainly in the initial stage. However,

the percentage mass loss corresponding to the decomposition of C-H at 3 days and 28

days for different combinations seems to be too small for a detailed interpretation.

Table 5.2 Summary of mass loss in different pastes

Type of mix Mass loss corresponding to Total mass loss


the decomposition of C-H (%)
(%)
3 days 28 days 3 days 28 days
Pure Cement paste 2.1 2 * 18
Paste with SMF-S1 1.5 1.9 10 17
Paste with PCE-D1 2.0 2.2 14 17
Paste with SNF-C2 1.4 2.1 9.4 15
*Could not be determined

5.5 NUCLEAR MAGNETIC RESONANCE SPECTROSCOPY (Si NMR)

5.5.1 Background

Nuclear magnetic resonance spectroscopy is a technique used to characterize the state

of polymerization of silicates in the cement paste. NMR determines the energy

(resonance absorption) needed to invert the nuclear spin in Si atoms due to a powerful

alternating magnetic field. Magic angle spinning (MAS) NMR is a technique used on

solid phases, which consists of spinning the sample at high frequencies at the magic

angle of 54°44’ to the magnetic field (Roncero, 2000; Roncero et al., 2002; Puertas et

al., 2005).

The MAS NMR of siliceous materials provides useful information about the state of

the SiO44- tetrahedrons. The Si MAS NMR spectra exhibit peaks at 5 different

chemical shifts that correspond to the different electronic environments of the Si

atom, which are affected by the length of the Si-O bond, the Si-O-Si angle and the

nature of neighbouring atoms. The chemical shifts depend on the number of oxygen

atoms shared by the tetrahedral, known as the degree of condensation of the

142
tetrahedra. Each type of tetrahedron connectivity is denoted as Qn, where n denotes

the number of shared oxygen atoms. Thus, the five chemical shifts are Q0, Q1, Q2, Q3,

Q4 in which the Si tetrahedra share 0, 1, 2, 3 and 4 oxygen atoms, respectively, with

other Si atoms.

Table 5.4 shows the chemical shifts corresponding to the five electron environments

described above. The anhydrous silicate phases of the cement (C3S, C2S) correspond

to the monomer state (Q0), which undergoes polymerization as the hydration

proceeds. The hydrated cement pastes manifest the presence of Q1 and Q2

tetrahedrons. The Q1 and Q2 peaks correspond to the formation of the C-S-H gel, the

structure of which is formed mainly by 3n-1 long chains of tetrahedrons.

Table 5.3 Ranges of 29 Si chemical shifts of the Qn tetrahedrons

Ranges of chemical shifts (ppm)


0 1
Q Q Q2 Q3 Q4
-66 to -74 -75 to -82 -85 to -89 -95 to -100 -103 to -115

5.5.2 Sample Preparation

The preparation of the sample was the same as that in the case of the DTA/TG

analysis. Tests were conducted on selected pastes at half setting, 3 days and 28 days.

The powdered sample was kept in air tight containers till the date of testing.

5.5.3 Results and Discussions


29
A Bruker BZH 300 solid state high resolution spectrometer was used for the Si

NMR of the pastes. Tetramethylsilane (TMS) was used as the reference material, and

the chemical shifts were determined with reference to that of the TMS. The details of

the Si NMR spectrum of reference paste, along with superplasticized pastes are given

in Figure 5.7 (a-g). The chemical shifts are represented along the horizontal axis and

the magnitude or intensity of the NMR resonance signal is displayed along the vertical

axis of the spectrum and it is proportional to the molar concentration of the sample. It

is observed that at half-setting, the peak corresponding to Q0 is identified in all the

143
pastes showing the anhydrous silicate phases of the cement (C3S, C2S) corresponding

to the monomer state. Two peaks corresponding to Q0 and Q1 can be identified in the

studied pastes for 3 days and 28 days. The peaks shows that the amount of dimer

gradually increases with a reduction in monomeric silicates for 3 to 28 days. Q2 units

were detected in the case of paste with SMF at 28 days compared to the paste with PCE.

This indicates that the retardation in hydration due to SMF is less than that with PCE.

Intensity of NMR signal

Chemical shifts

(a) Cement paste, half setting (b) Cement paste, 3 days (c) Cement paste, 28 days

144
-73.191
-82.091
-73.232
-84.669

-73.247
-86.299

-68.084
-76.155
-50 -100 -50 -100 -50 -100 -50 -100

(d) and (e) SMF based paste, 3 and 28days (f) and (g) PCE based paste, 3 and 28days

Fig. 5.7 Details of Si NMR spectrum of different pastes for different ages

The relative concentration of the Qn units can be determined by the integration of their

corresponding curves. Since the Q0 and Q1 peaks in the NMR spectra overlap, the

areas corresponding to the Q0 and Q1 peaks, denoted as A0 and A1, respectively, are

obtained for each spectrum as in Figure 5.8.

Q0

Q1

A0
A1

-60 -80 -100

Fig. 5.8 Determination of area of peaks of the Si NMR spectrum

145
The relative area of the Q1 peak (A1/A0) ratios for each paste is shown in Figure 5.9. It

can be observed that the superplasticizer has a significant effect on the polymerization

of the silicates during the hydration process. Its incorporation leads to the formation

of a lower amount of dimers at 3 days as well as at 28 days, reflected by the lower

relative areas of Q1 peak in all the superplasticized pastes compared to reference

paste.

2.5 2.436
Half setting
3 days
28 days
2.0 1.828

1.510
1.5
A1/A0

1.172

1.0
0.699 0.638

0.5

0.0
Reference SMF PCE
Type of mix

Fig. 5.9 Comparison of area of the Q1 peak (A1) with reference to the
area of the Q0 peak (A0) at different ages

It is observed that the silicate polymerization is more in the paste with SMF than in

the paste with PCE i.e., (A1/A0) is more for SMF). The results obtained from setting

tests also supports this result as the setting time for the paste with PCE was more than

that for the paste with SMF. Further to this, the change in the relative area of peaks

indicates a change in the C-S-H morphology in the presence of superplasticizers.

5.6 SUMMARY

A multiple technique based approach, instead of a single test method, is used for

studying the effect of the superplasticizer type on the hydration and the

146
microstructural development for a better understanding of the influence of the

incorporation of superplasticizers on the hardened properties. The comparison of X-

ray diffractograms of the reference paste and superplasticized paste shows that the

superplasticizers mainly influence the hydration of aluminates and consequently the

formation of ettringite rather than the formation of portlandites. However, the

quantitative analysis based on the thermal techniques shows that the amount of C-H

formed for superplasticized pastes is less than that of reference paste at 3 days

Also, the retardation in hydration in the presence of superplasticizers can be observed

through the BSE images, in which the amount of unhydrated particles in the

superplasticized paste is more than that in the reference paste. Also, a lower

polymerization of the silicates has been observed through the NMR study, in the

presence of the superplasticizer, which is reflected by lower proportion of dimers at

the age of 3 and 28 days. In addition, the silicate polymerization is also affected by

the type of superplasticizer; the paste with PCE has a lower amount of dimers

compared to the paste with the SMF. This has been observed in setting tests in which

PCE retarded the setting more than the SMF.

Even though the polymerization of C-S-H is lower in the superplasticized pastes, a

slight increase in compressive strength due the incorporation of superplasticizer was

observed (see Section 4.7); it appears that the mechanical behaviour is dependent

more on the dense and uniform microstructure developed in the superplasticized

pastes. Hence, it is confirmed through these studies that the superplasticizers not only

affect the fresh state properties but also the hardened state properties of the cement

paste.

147
CHAPTER 6

CEMENT-SUPERPLASTICIZER COMPATIBILITY

6.1 GENERAL

The influence of the type and dosage of superplasticizers on the flow behaviour of

paste and concrete, as well as on the microstructure of the cement paste, has been

discussed in the previous Chapters. In addition, the chemical composition of cement

also plays an important role in the cement-superplasticizer interaction and the

incompatibilities that can arise. An incompatibility can be defined as any undesirable

effect that may be produced in the properties of concrete due to the particular

combination of cement and superplasticizer. It is obvious that with a variety of

cements and variability of their chemical compounds incompatible situations could

arise. The literature reveals that cement-superplasticizer compatibility is affected by

the following parameters related to the cement: chemical and phase composition,

especially the C3A content, alkali content, amount and type of calcium sulphate

(gypsum, hemihydrates or anhydrite), cement fineness and free lime content

(Hanehara and Yamada., 1999; Jiang et al., 1999; Yamada et al., 2001; Tagnit-Hamou

et al., 2003).

The objective of this chapter is to understand the influence of the chemical

composition of the cement on the cement-superplasticizer interaction and to suggest

some guidelines for selecting the best combinations. Compatibility in terms of

rheology and setting time is investigated for different types of cements with different

superplasticizers. Also, an attempt is made through an adsorption study based on UV

absorbance to understand the reason for the loss of fluidity and to explain why a

higher dosage is required for some types of superplasticizers.


6.2 MATERIALS SELECTION AND CHARACTERISATION

6.2.1 Types of cements

To consider a reasonable number of material combinations, three commercial brands

of cements were identified. Cement that had created some incompatibility problems in

the laboratory trials was selected as the fourth one. The chemical analyses of these

cements were done in National Test House, Chennai, and the physical properties were

determined in the Construction Materials Laboratory of IITM. The chemical

properties of selected cements are given in Table 6.1, along with its Bogue

composition determined as per ASTM C 150 (2007a). The physical properties are

given in Table 6.2. The important aspects of the four cements are given below.

Cement 1 (C1)

This cement has C3S, C2S and oxide contents in the normal range and a low C3A

content. The alkali and SO3 contents are also within the allowable limits of 0.2-1.2%

and 2-3.5%, respectively, of ordinary portland cement.

Cement 2 (C2)

It contains high C2S and high insoluble residue but a low C3S content. It has the

highest SiO2 content of the cements studied whereas the other oxides are within

normal ranges. This cement has a low alkali content as the (Na2O)eq is 0.086%. Also,

its fineness is higher compared to the other cements.

Cement 3 (C3)

CaO content and alumina/iron oxide ratio is higher for this cement compared to other

cements. Other compounds are within the permissible limits.

Cement 4 (C4)

CaO and SO3 are least for this cement whereas Al2O3 and MgO are highest, compared

to the other cement studied. Loss on ignition is least for this cement whereas the

149
insoluble residue is high. It has the highest C3A content and also a high alkali content

compared to other cements.

Table 6.1. Chemical and Bogue compositions of different cements

Analyte Percentage of mass


C1 C2 C3 C4
CaO 60.81 61.14 62.13 59.44
SiO2 19.5 22.95 20.84 21.04
Al2O3 4.12 4.62 4.88 5.67
Fe2O3 4.72 4.1 4.29 4.99
MgO 1.52 0.72 0.85 1.78
SO3 2.48 2.48 1.69 1.39
L. O. I 3.41 3.61 3.47 1.97
Insoluble residue 1.51 2.98 1.41 3.3
Cl 0.01 0.02 0.03 0.03
Na2O 0.05 0.04 0.18 0.22
K2 O 0.28 0.07 0.52 0.54
(Na2O)eq 0.234 0.086 0.52 0.575
Lime saturation factor 0.93 0.82 0.91 0.86
Alumina/iron oxide ratio 0.68 1.13 1.14 1.136
C3S 53.99 28.21 48.72 30.78
C2S 15.18 44.51 22.99 37.10
C3A 2.94 5.31 5.673 6.58
C4AF 14.36 12.48 13.05 15.18

Table 6.2. Physical properties of different types of cements

Physical property Quantity


C1 C2 C3 C4
Water demand for normal consistency of
cement (%) 33 30 31 31
Initial setting time (minutes) 99 100 170 85
Final setting time (minutes) 184 220 290 215
Blaine specific surface area (m2/kg ) 316 371 355 301

6.2.2 Superplasticizers

Superplasticizers used in the present compatibility study with cements C2, C3 and C4

were selected from the work presented in Chapter 3 where only cement C1 was used.

Four superplasticizers (a lignosulphonate, a melamine, a naphthalene and a

150
polycarboxylate) all belonging to ASTM C 494 (2008a) type F and the fifth one

belonging to Type G were selected; see Table 3.4 for the properties. Another

naphthalene based superplasticizer (SNF-C2), which had created problems of slow

setting, was also used along with C3 and C4; the properties of this admixture are:

solid content of 40%, density of 1.2 kg/litre and pH of 7.5.

6.3 Experimental Procedure

The mixing procedure for the preparation of cement paste was the same as given in

Section 3.4. The pastes were tested at ambient temperature, in the range of 30-32°C.

The tests conducted to understand the flow behaviour were the Marsh cone, mini-

slump and viscometric tests. Marsh cone test was conducted to determine the

saturation dosage of superplasticizer in each case, which is supported by mini-slump

test results. The Bingham model was used to derive the yield stress and plastic

viscosity of different compositions from viscometric test results. The setting of

different combinations was studied with the Vicat apparatus.

6.4 Results and discussions

6.4.1 Influence of Cement Composition on the Flow Behaviour of


Superplasticized Pastes

The flow behaviour of cement C1 with different types of superplasticizers has been

discussed in detail already in Chapter 3; the Marsh cone flow time curves are

reproduced in Figure 6.1(a) to provide comparisons with the behaviour obtained with

other three cements that will be discussed here. The salient features of the behaviour

of the pastes with the cement C1 are: (1) the saturation dosage of LS-C1 is slightly

higher than those of the SNF and SMF based superplasticizers whereas the flow

behaviour is comparable; (2) at very low dosages, SMF-S1 seems to give a high flow

time, indicating the poorer dispersion of cement particles and lower fluidity; (3) the

lower saturation dosage of 0.07% for PCE-D1 indicates that the combined
151
mechanisms of electrostatic repulsion and steric hindrance makes it effective even at

very low dosages; and (4) the flow time curves corresponding to 60 minutes are

similar to those at 0-minutes for the saturation dosage and above (except for the case

of the SMF-S1), indicating a negligible loss in fluidity beyond the saturation dosage,

which reiterates the importance of using an appropriate dosage of superplasticizer for

maintaining the fluidity for the required period.

The Marsh cone flow time curves of 0 and 60 minutes for the other three cements C2,

C3 and C4 are given in Figure 6.1 (b-d). In the case of cement C2, the saturation

dosages are generally higher than those obtained with cement C1 and the loss of

fluidity is high only in the paste with LS-C1. With cement C3, though the initial

fluidity is in the same range as in the other cements for the studied combinations,

there is generally a higher loss of fluidity, especially at dosages less than the

saturation point. The loss of fluidity with cement C4 is even higher for all the studied

combinations and severe at low superplasticizer dosage. This can be attributed to a

high C3A content in cement C4 (compared to cement C1).

LS-C1 0
LS-C1 60
70 Cement C1 SNF-S1 0
SNF-S1 60
Marsh cone flow time (sec)

60 SNF-D1 0
SNF-D1 60
SNF-D2 0
50 SNF-D2 60
SNF-C1 0
40 SNF-C1 60
SMF-S1 0
SMF-S1 60
30
SMF-C1 0
SMF-C1 60
20 PCE-S1 0
PCE-S1 60
10 PCE-D1 0
PCE-D1 60
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

(a) Marsh cone flow time curves for Cement C1

152
60 Cement C2
LS-C1 0
LS-C1 60

Marsh cone flow time (sec)


50 SNF-S1 0
SNF-S1 60
SNF-D2 0
40 SNF-D2 60
SMF-S1 0
SMF-S1 60
30 PCE-D1 0
PCE-D1 60

20

10
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

(b) Marsh cone flow time curves for Cement C2

60
Cement C3
Marsh cone flow time (sec)

LS-C1 0
50 LS-C1 60
SNF-S1 0
SNF-S1 60
40 SNF-D2 0
SNF-D2 60
SNF-C2 0
30 SNF-C2 60
SMF-S1 0
SMF-S1 60
PCE-D1 0
20
PCE-D1 60

10

0.0 0.2 0.4 0.6 0.8


sp/c %

(c) Marsh cone flow time curves for Cement C3

153
70
LS-C1 0
Cement C4 LS-C1 60
60 SNF-S1 0
SNF-S1 60

Marsh cone flow time


50 SNF-D2 0
SNF-D2 60
SNF-C2 0
40 SNF-C2 60
SMF-S1 0
30 SMF-S1 60
PCE-D1 0
PCE-D1 60
20

10

0.0 0.1 0.2 0.3 0.4 0.5 0.6


sp/c %
(d) Marsh cone flow time curves for Cement C4

Fig. 6.1 Marsh cone flow times of different cement-superplasticizer combinations

The data from the Marsh cone, mini-slump and viscometer tests for the pastes with

different superplasticizers at 0 and 60-minutes and all the four cements are given in

Tables 6.3-6.6, respectively. The results show that the trends obtained are generally

the same; i.e., the fluidity of the paste increases with an increase in the dosage of

superplasticizer upto saturation dosage and remains practically constant after that.

The mini-slump spreads reflect the loss of fluidity at low superplasticizer dosages, as

already discussed in the context of the Marsh cone flow times. Beyond the saturation

dosage, the spread value at 60 minutes is almost the same as that obtained

immediately after mixing. It was seen that pastes with mini-slump spreads of more

than 165 mm exhibited bleeding. Also, some pastes that exhibited bleeding at 0

minutes improved in this aspect after 60 minutes due to the reduction in fluidity.

154
Table 6.3 Test data for pastes with different superplasticizers for cement C1

Flow time Mini slump spread Yield stress Plastic viscosity


Type of (sec) (mm) (Pa) (Pa s)
sp/c
sp
%
0 60 0 60 0 60 0 60
min min min min min min min min

0.1 20 21 93 90 12.20 13.50 0.122 0.132


0.2 14 16 153 123 5.81 7.59 0.089 0.097
LS-C1 0.3 13 13 162 143 2.0 3.15 0.059 0.075
0.4 13 14 163 163 1.0 0.35 0.050 0.064
1 12 13 164 162 0.5 0.23 0.068 0.049
0.2 13 26 143 90 16.21 24.61 0.143 0.185
SNF-S1 0.3 11 11 151 150 1.80 5.91 0.055 0.066
0.4 11 11 165 163 0.0 0.21 0.033 0.054
1 11 11 170 178 0.09 0.10 0.031 0.025
0.1 45 * 82 * 18.31 20.23 0.197 0.241
0.2 20 70 96 * 12.13 16.61 0.139 0.178
SNF-D1 0.3 13 13 127 115 2.88 5.98 0.074 0.089
0.4 14 17 bleeding 141 0.08 0.09 0.049 0.083
0.5 14 18 bleeding 164 0.11 0 0.043 0.045
1 14 13 bleeding 163 0.09 0.12 0.033 0.035

SNF- D2 0.1 29 * 87 * 16.99 25.52 0.197 0.194


0.2 17 39 106 87 6.94 12.10 0.088 0.149
0.4 14 14 bleeding 148 0.55 0 0.042 0.043
0.6 13 13 bleeding 176 0.37 0.25 0.041 0.039
1 13 12 bleeding 160 0.13 0.07 0.045 0.038

SNF-C1 0.1 11 14 126 95 6.72 7.24 0.057 0.046


0.2 10 11 152 128 1.25 6.69 0.042 0.064
0.3 10 10 168 168 0.72 0.77 0.029 0.046
1 10 9 173 178 0.24 0.17 0.023 0.035
0.1 62 * * 20.39 * 0.177 *
SMF-S1 0.2 18 44 102 90 6.89 2.39 0.076 0.092
0.25 16 22 115 109 2.36 3.77 0.098 0.101
0.3 15 18 122 113 1.54 8.97 0.078 0.065
1 15 18 167 161 0.77 0 0.048 0.043

0.1 >120 * * * 33.14 * 0.199 *


SMF-C1 0.2 16 26 108 103 13.44 30.92 0.128 0.139
0.3 13 16 132 112 2.88 7.71 0.074 0.100
0.5 12 12 181 169 0 0 0.039 0.042
1 12 11 177 172 0 0 0.037 0.041

0.05 26 * 86.3 * 17.55 * 0.155 *


PCE-S1 0.1 16 * 110 * 11.07 23.67 0.089 0.162
0.2 12 14 bleeding 157 0.51 0.36 0.073 0.068
0.3 11 13 bleeding 179 0 0.08 0.034
0.5 11 14 bleeding 192 0 0 * 0.035
1 12 12 bleeding bleeding 0 0.034 0.034

0.05 18 * 97 * 24.45 * 0.214 *


PCE-D1 0.1 12 16 166 147 0.01 2.37 0.035 0.079
0.2 12 14 170 bleeding 0 0.09 0.035 0.043
0.4 12 14 199 bleeding 0 0 0.036 0.036
1 14 16 193 bleeding 0 0 0.047 0.054

155
From the data of the viscometer tests, it can be seen that in general the yield stress

decreases significantly with an increase in the dosage of superplasticizer down to a

value of practically zero beyond the saturation dosage. The plastic viscosity also

reduces with an increase in the superplasticizer dosage. It is observed that the yield

stress and plastic viscosity developed vary from 0-43 Pa and 0.034-0.188 Pa.s,

respectively. Both the rheological parameters increase with time as seen in the

comparison of the 0 and 60 minute data.

Table 6.4 Test data for pastes with different superplasticizers for cement C2

Type Flow time Mini slump spread Yield stress Plastic viscosity
of sp/c (sec) (mm) (Pa) (Pa s)
sp % 0 60 0 60 0 60 0 60
min min min min min min min min
0.1 36 * 85 * 33.45 * 0.185 *
0.2 13 59 123 90 15.33 24.34 0.095 0.142
LS-C1 0.3 13 45 140 113 7.72 12.99 0.086 0.113
0.5 11 14 bleeding 143 0.26 3.61 0.047 0.090
1 14 23 bleeding 120 0.24 1.28 0.068 0.130
0.1 26 * 84 * 38.16 * 0.162 *
0.2 13 28 125 89 11.64 1.83 0.095 0.076
SNF- 0.3 11 15 bleeding 138 1.2 3.04 0.073 0.073
S1 0.5 12 11 bleeding bleeding 0.19 0.32 0.040 0.034
1 11 10 bleeding bleeding 0.31 0.13 0.043 0.043
0.1 31 * 85 * 22.56 * 0.170 *
0.2 12 19 130 113 6.78 26.27 0.074 0.181
SNF- 0.3 11 13 153 143 0.44 3.85 0.057 0.073
D2 0.5 10 10 bleeding bleeding 0 0.014 0.043 0.039
0.8 10 10 bleeding bleeding 0 0 0.038 0.046
0.1 * * * * * * * *
0.2 20 * 85 * 20.73 34.26 0.123 0.188
0.3 15 16 148 145 13.13 12.32 0.090 0.053
SMF- 0.4 15 15 149 150 1.65 3.72 0.083 0.078
S1 0.5 14 15 153 148 0.70 1.15 0.084 0.082
1 13 13 bleeding bleeding 0.04 0 0.052 0.052
0.05 47 * 75 * 42.97 * 0.165 *
0.1 18 40 115 85 13.39 23.98 0.110 0.143
PCE- 0.2 13 16 145 153 0.51 0.80 0.087 0.102
D1 0.3 13 14 155 163 0.42 0.36 0.074 0.074
1 16 18 158 148 0.32 0.56 0.086 0.092

156
Table 6.5 Test data for pastes with different superplasticizers for cement C3

Type Flow time Mini slump spread Yield stress Plastic viscosity
of sp/c (sec) (mm) (Pa) (Pa s)
sp % 0 60 0 60 0 60 0 60
min min min min min min min min
0.05 22 * 83 * 14.99 * 0.163 *
0.1 15 * 102 74 13.98 17.33 0.119 0.112
0.2 13 * 128 77 11.39 21.82 0.099 0.147
LS-
C1 0.3 12 34 145 91 4.07 15.74 0.072 0.122
0.4 12 23 bleeding 108 0.29 12.70 0.056 0.100
0.6 12 21 bleeding 148 0.067 2.11 0.019 0.103
0.025 29 * 97 75 16.10 21.78 0.152 0.197
0.05 21 * 100 85 15.84 25.34 0.169 0.145
0.1 15 * 107 85 14.26 19.04 0.109 0.182
SNF-
S1 0.2 11 15 139 118 3.21 12.16 0.061 0.109
0.3 10 11 bleeding 154 0.59 0.53 0.043 0.054
0.6 10 11 bleeding 0 0.62 0.029 0.084
0.025 21 * 85 76 3.189 22.91 0.086 0.168
0.05 14 40 108 85 13.272 15.93 0.113 0.135
0.1 12 21 125 108 7.573 14.95 0.058 0.124
SNF-
D2 0.2 11 11 140 148 1.023 0.76 0.036 0.065
0.3 12 12 bleeding 165 0 1.61 0.048 0.289
0.6 10 11 bleeding 170 0 0 0.033 0.021
0.025 26 * 93 * 16.02 * 0.138 *
0.05 19 * - 78 15.68 18.39 0.145 0.182
0.1 16 44 - 79 9.61 18.65 0.079 0.176
SNF-
C2 0.2 15 27 108 83 7.75 16.05 0.073 0.137
0.3 15 15 144 108 3.22 5.02 0.048 0.065
0.4 12 12 bleeding 190 0 0 0.019 0.042
0.6 11 12 bleeding bleeding 0 0 0.016 0.012
0.025 23 * 98 * 17.59 21.75 0.17 0.217
0.05 18 * 93 * 14.60 20.42 0.126 0.179
0.1 16 * 103 73 14.31 * 0.122 *
SMF-
S1 0.2 11 16 128 103 5.30 9.63 0.071 0.069
0.3 11 13 141 115 1.53 5.46 0.056 0.073
0.5 11 12 bleeding bleeding 0.72 0.15 0.043 0.039
0.8 11 11 bleeding bleeding 0 0 0.03 0.031
0.025 12 * 130 * 6.78 * 0.077 *
0.05 11 55 135 85 9.45 17.02 0.056 0.116
0.075 10 13 bleeding 130 1.44 3.87 0.039 0.071
PCE-
D1 0.1 10 14 bleeding 120 1.12 4.38 0.035 0.094
0.2 10 10 * 168 0 - 0.030 -
0.3 10 9 bleeding 175 0 0.12 0.032 0.0096

157
Table 6.6 Test data for pastes with different superplasticizers for cement C4

Flow time Mini slump spread Yield stress Plastic viscosity


Type (sec) (mm) (Pa) (Pa s)
sp/c
of sp
% 0 60 0 60 0 60 0 60
min min min min min min min min
0.1 36 * 78 * * * * *
0.2 14 * 105 * 22.83 * 0.144 *
LS-C1 0.3 13 * 125 * 10.36 * 0.088 *
0.4 12 18 165 100 2.59 * 0.073 *
0.6 14 16 bleeding bleeding 0.30 0.1 0.047 0.054
0.05 37 * 88 * 24.81 * 0.221 *
0.1 25 * 86 * 22.26 * 0.188 *
SNF- 0.2 13 23 130 106 5.42 17.40 0.072 0.137
S1 0.3 11 12 bleeding 146 0.15 1.07 0.038 0.044
0.5 11 10 bleeding 168 0 0 0.026 0.026
0.05 65 86 * 29.59 * 0.264 *
0.1 14 >84 118 83 10.95 22.60 0.095 0.238
SNF- 0.2 12 14 158 124 0.37 2.71 0.045 0.071
D2 0.3 11 11 178 160 0 0.0029 0.034 0.055
0.5 12 11 bleeding bleeding 0 0 0.031 0.026
0.1 56 * 81 * 33.21 * 0.209 *
0.2 16 * 101 83 21.80 37.50 0.159 0.261
0.3 14 28 114 85 10.31 22.73 0.082 0.156
SNF- 0.4 11 15 bleeding 123 1.129 9.21 0.063 0.071
C2 0.5 11 12 bleeding bleeding 0 0.16 0.051 0.037
0.6 11 11 bleeding 165 0 0 0.043 0.037
0.1 40 * 87 * 35.41 * 0.304 *
0.2 16 * 107 * 20.69 * 0.185 *
0.3 12 21 139 91 3.99 22.03 0.072 0.182
SMF- 0.4 12 17 145 105 4.04 9.29 0.078 0.087
S1 0.5 10 11 bleeding 168 0.24 0.09 0.052 0.040
0.05 31 * 83 * * * * *
0.075 14 * 112 * 28.25 * 0.132 *
PCE- 0.1 13 * 123 * 13.33 * 0.079 *
D1 0.15 10 16 bleeding 111 0.21 7.48 0.044 0.102
0.2 11 13 bleeding bleeding 0.09 0 0.034 0.06
0.3 11 11 178 bleeding 0.03 0 0.025 0.035
0.6 10 10 188 184 0 0 0.026 -
* could not be determined

Table 6.7 gives a comparison of the saturation dosages for the four cements obtained

from Marsh cone flow curves at 0 minutes, using the same method as given in

Chapter 3. It is observed that the saturation dosages of the LS, SMF and PCE based

superplasticizers are generally the highest for the cement C2, mainly because it has

comparatively higher surface area and hence the amount of superplasticizer needed is

higher, which is in accordance with the reported results (Erdogdu, 2000). Also, it is a

158
low alkali cement and therefore is expected to adsorb more superplasticizer (Dodson

and Hayden, 1989; Jiang et al., 1999)

Table 6.7 Comparison of the saturation dosages at 0-minutes for different cements

Type of Saturation dosages%


superplasticizer C1 C2 C3 C4
LS-C1 0.25 0.40 0.25 0.4
SNF-S1 0.20 0.25 0.20 0.25
SNF-D1 0.24 * * *
SNF-D2 0.23 0.3 0.2 0.2
SNF-C1 0.16 * * *
SNF-C2 * * 0.2 0.4
SMF-S1 0.23 0.4 0.2 0.3
SMF-C1 0.20 * * *
PCE-S1 0.2 * * *
PCE-D1 0.07 0.2 0.08 0.15

Cement C3 gives good fluidity even at very low dosages and the saturation dosages

obtained are comparatively lower for the various superplasticizers. The yield stress

and plastic viscosity values are lower compared to those of pastes with other cements.

However, some pastes with cement C3 do show a loss of fluidity within 60 minutes.

All combinations with cement C4 had rapid slump losses within 60 minutes and the

saturation dosages of SNF-C2, PCE-D1 and LS-C1 were high. This may be due to the

high C3A content, which causes some of the superplasticizer to be absorbed by the

initial hydration products of C3A and hence less would be available in the solution for

further adsorption. Also, cement C4 has a low SO3 content and hence the SO42- ions in

the superplasticizer will compete with the sulphate ions of the cement, causing more

initial adsorption, resulting in the increase of the superplasticizer demand.

159
The results reported by Kumar (2006) on the compatibility of concrete using cement

C1 and C2 and different superplasticizers used here generally support the results of

the present study; he found that the superplasticizers SMF-S1, SNF-S1 and PCE-D1

in combination with cement C1 gave better slump retention in concrete, even at low

superplasticizer dosages, compared to the concretes with cement C2. As an extension

of the previous study, Das (2007) conducted tests to study the compatibility of

cements C3 and C4 and the same superplasticizers. His results also show that

concretes with cement C4 had more slump loss than those with cement C3, as

reported in the present study.

6.4.2 Influence of Cement Composition on the Setting of Superplasticized Pastes

The setting time is affected by both the chemical composition of cement, as well as by

the type of superplasticizer, as seen in Figure 6.2 (a-c). It is seen that the trend of

setting varies among the cements studied. It can be observed that for cement C1 and

C3, the lowest retardation is seen for SMF-S1 but not in C4. For cement C4, LS-C1

and SNF-D2 superplasticizers gave the least retardation compared to other

superplasticizers with the same cement. This is in sharp contrast to the behaviour of

LS-C1 and SNF-D2 in the other cements, where it leads to high retardation. The mix

with cement C4 and admixture SNF-C2 does not set even after 24 hours whereas the

paste with cement C3 and the same admixture sets in about 10 hours.

It can be seen that the setting behaviour can vary significantly depending on the

cement composition and the superplasticizer used. Some of the trends cannot be

explained on the basis of the literature or the data obtained in this work, and call for

further research to provide the understanding needed.

160
Cement C1 Control paste
40 LS-C1
SNF-S1
Initial setting SNF-D2

Depth of penetration (mm)


SMF-S1
30 PCE-D1

20

10

0
200 400 600 800 1000 1200
Time (minutes)

(a) Vicat penetration test results of Cement C1

45 Control paste
Cement C3 LS-C1
40
SNF-S1
Initial setting SNF-D2
Depth of penetration (mm)

35
SNF-C2
30 SMF-S1
PCE-D1
25
20
15
10
5
0
200 400 600 800
Time (minutes)

(b) Vicat penetration test results of Cement C3

161
45
Cement C4 Control paste
40 LS-C1
Initial setting SNF-S1

Depth of penetration (mm)


35
SNF-D2
30 SNF-C2
SMF-S1
25 PCE-D1
20
15
10
5
0
400 800 1200 1600 2000
Time (minutes)

(c) Vicat penetration test results of Cement C4

Fig. 6.2 Influence of type of cement on setting

The results of Kumar (2006) on setting with cement C1 shows the same trend as

observed in the present paste study; the SNF-S1 gives lower setting time than PCE

and SNF-D2 based superplasticized paste. The results of Das (2007) on setting are

also comparable with the present results; he showed that concrete with the

superplasticizer SNF-S1 exhibited higher setting time with cement C4 whereas SNF-

D2 had a lower setting time, when compared with identical concretes with the cement

C3.

6.5 STUDY OF THE ADSORPTION OF SUPERPLASTICIZERS

6.5.1 Background

As discussed earlier, the superplasticizer is adsorbed on the cement surface; the rate

and amount of this adsorption varies with type of cement and superplasticizer, which

in turn affects the fresh state properties of concrete. Consequently, adsorption plays an

important role in the initial flow behaviour, the maintenance of fluidity as well as the

setting behaviour of pastes. Hence an understanding of adsorption of different

162
superplasticizers on different types of cement is required for answering several

questions like why some superplasticizers that have good initial fluidity can not

maintain it even for 60-minutes. Also, it could explain the variation in saturation

dosage of the same superplasticizer for different types of cements.

The increase in workability due to the addition of superplasticizers occurs through the

dispersion of agglomerated cement particles. However, the quantification of this

mechanism is a difficult task and is further complicated by the hydration reactions.

The adsorbed superplasticizer causes fluidity and the superplasticizer that is

remaining in solution is responsible for fluidity retention, as explained in Chapter 2.

UV absorbance techniques have been used to understand the amount of adsorption by

several researchers, who have concluded that the method is highly precise and the

results obtained are reproducible (Daimon and Roy, 1978, Andersen et al., 1987,

Krishna, 1996, Kim et al., 2000).

In the present study, the adsorption behaviour of four families of superplasticizers and

its relation to the fluidity of cement paste has been investigated at the saturation

dosage of superplasticizer using UV absorbance technique.

6.5.2 Materials used

Superplasticizers belonging to the four families of products, namely lignosulphonates

(LS), melamines (SMF), naphthalenes (SNF) and polycarboxylates (PCE), were used.

The details of the superplasticizers are given in Table 3.4. OPC 53 grade cements C1,

C2,C3 and C4, with the characteristics given in Table 6.1, were used in the study.

6.5.3 Experimental Procedure

The concentration of polymer in the liquid phase was determined through a

quantification of the UV absorbance using a JASCO V-570 UV spectrophotometer in

163
the 190-400 nm region. According to the Beer-Lambert’s law, the absorbance is

directly proportional to the concentration of the absorbing material in solution. In

absorption spectroscopy, the intensity of light of a particular wavelength that is

absorbed by the sample is measured and used to determine the absorbance using the

following equations:

I0
A = log10 (6.1)
I1

where A is the absorbance, A = α l c (6.2)

α is the absorption coefficient, l is the distance that the light travels through the

material, c is the concentration of the absorbing species in the material, I0 is the

intensity of incident light, and I1 is the intensity after passing through the material.

UV spectrometry is used here for the measurement of the absorbance of different

superplasticizers. The test procedure is as follows.

Step 1: Determination of the characteristic peak wavelength (λmax) of the


superplasticizer

The first step in the method is to determine the wavelength corresponding to

maximum absorbance for a particular superplasticizer; absorbance measurements

made at that wavelength are expected to have the least errors. 0.1 ml of a solution of

water and superplasticizer, with a known superplasticizer concentration, is taken and

further diluted as 250X, 500X, 1000X, 10000X, etc. Water, used as the reference

liquid for the measurements, was placed in another couette along with the diluted

sample in the UV-visible spectrometer. Measurements are made for each of the

diluted solutions until an absorbance peak is obtained within the measurable range of

the instrument (maximum of 5 for the equipment used here). The UV spectra for the

different superplasticizers are shown in Figure 6.3, showing different peak absorbance

164
values and the corresponding wavelengths (denoted as λmax). Note that the dilutions

are different in all the cases so that absorbance values are not directly comparable. It

is observed that the λmax is 205 nm for LS-C1, 215 nm for SMF-S1, 225 nm for SNF-

S1 and 195nm for PCE-D1. The absorbance by the superplasticizer in the aqueous

solution was determined hereafter at the corresponding λmax.

3.5
205nm
3.0 LS-C1
SNF-S1
2.5 225nm SMF-S1
Absorbance

PCE-D1
2.0

1.5

1.0

0.5 215nm
195nm
0.0
200 250 300 350 400
Wavelength (nm)

Fig. 6.3 Absorbance spectra for different superplasticizers

Step 2: Determination of the absorbance of the superplasticizers in cement paste

The cement pastes were prepared with the saturation dosage of superplasticizer. The

aqueous solution was extracted by centrifuging the paste at 6000 rpm for 10 minutes.

The extracted solution was diluted with deionized water to obtain a solution with an

absorbance in the measureable range. The absorbance spectrum is obtained and the

peak absorbance was noted for the corresponding wavelength, as explained in step 1.

6.5.4 Results and discussion of absorbance tests

The absorbance spectrum of the aqueous solution extracted from the paste with

cement C1 and lignosulphonate based superplasticizer just after mixing is given in


165
Figure 6.4. The comparison of the same with Fig. 6.3 shows that peak absorbance for

this superplasticizer occurs at the same wavelength of 205 nm for both the solution in

water and the aqueous solution extracted from the cement paste. It is also seen that

there is no other peak in the spectrum, which indicates that there is no other species in

the aqueous solution of the paste with significant absorbance other than the

superplasticizer. The same is true for all the other superplasticizers, as seen in the

other plots of Figure. 6.4.

3.0
LS-C1 with cement C1
2.5 SNF-S1 with cement C1
SMF-S1 with cement C1
PCE-D1 with cement C1
2.0
Absorbance

Pure cement(C1) paste

1.5

1.0

0.5

0.0

200 250 300 350 400


Wavelength (nm)

Fig. 6.4 Absorbance spectrum of the diluted aqueous solutions


of superplasticized pastes extracted after mixing

The absorbance spectrum of the solution extracted from the pure cement paste is also

given in the same graph. Note that each of the curves is for a different dilution and are

not, therefore, directly comparable; the dilutions for the corresponding aqueous

solutions are: LS-C1: 250X (i.e., 0.1 ml of the aqueous solution is diluted in 25 ml of

water), SNF-S1: 500X, SMF-S1: 1000X and PCE-D1: 250X. The absorbance spectra

of the superplasticized cement pastes at 90 minutes were also obtained in the same

166
manner. Note that the age of the aqueous solution at the time of the measurements is

estimated as 15 minutes.

The absorbance values for tests at 15-minutes and 90-minutes for the different

superplasticizers in combination with cement C1 are summarized in Table 6.8. The

absorbance at the saturation dosage for different pastes depends on both the family of

superplasticizer and chemical composition of cement. The absorbance of SNF-S1 and

SMF-S1 is lower at 90-minutes compared to that at 15-minutes showing that the

adsorption of the superplasticizer has increased with time. In the other cases, the

absorbance is almost the same at both 15 and 90 minutes indicating that the

adsorption of those superplasticizers does not increase over this time range. It should

be noted that the presence of cement particles in the aqueous solution and the possible

desorption of the superplasticizer from these fine particles during subsequent dilution

process could distort the absorbance values by giving falsely high values that could be

attributed to the unadsorbed admixtures (Daimon and Roy, 1978). This is a major

limitation of the procedure and further studies are needed to get a better estimate of

the amount of adsorbed superplasticizer on the cement particles and its evolution with

time.

Table 6.8 Summary of absorbance test results


Dilution Absorbance
Combination 15 90 15 90
minutes minutes minutes minutes
Cement C1+LS-C1 250X 250X 2.7 2.9
Cement C1+SNF-D1 500X 500X 2.6 1.9
Cement C1+SMF-S1 1000X 1000X 1.1 0.8
Cement C1+PCE-D1 250X 250X 1.3 2.1
Cement C1 +PCE-D1 (0.05%) 250X 250X 1.1 1.2
Cement C2 +LS-C1 250X 250X 2.2 2.0
Cement C2+PCE-D1 250X 250X 0.4 0.4
Cement C4+PCE-D1 250X 250X 1.3 1.0

167
6.5.4.1 Variation in Superplasticizer Absorbance with Type of Cement

The absorbance for the pastes with cement C2 and LS-C1 is less than that with cement

C1 and LS-C1 for the same dilution (see Figure 6.5). This suggests that the initial

adsorption may be more in the case of the paste with cement C2 and LS-C1 than in

the paste with cement C1 and LS-C1, resulting in the higher loss of fluidity, as

described earlier. This could be attributed to the higher surface area of cement C2 and

its low alkali content, both of which can increase the adsorption. Similarly, for the

paste with cement C2 and the PCE, the absorbance at 15 and 90-minutes is low

compared to the C1-PCE-D1 combination, supporting the results shown earlier where

there is more loss of fluidity and higher superplasticizer demand in the former.

2.5

15 minutes
2.0 90 minutes
Absorbance

1.5

1.0

0.5

0.0
200 250 300 350 400
Wavelength (nm)

Fig. 6.5 Absorbance spectra for cement C2 and LS-C1

6.6 IDENTIFYING COMPATIBLE CEMENT-SUPERPLASTICIZER


COMBINATIONS

As indicated in the preceding chapters, a number of mechanisms and interactions can

contribute to the so-called incompatibility between the cement and superplasticizer.

All of these mechanisms are complex and interrelated. The results show that a single
168
test is not sufficient to determine a compatible cement-superplasticizer combination.

Consequently, a simple methodology that combines different test procedures is

developed based to select the compatible combinations.

A flow chart representing the methodology for the selection of an appropriate cement-

superplasticizer combination based on the present interaction study is shown in

Figure 6.6. The entire methodology is divided into two steps; the method of selection

of a compatible combination of cement and superplasticizer from the paste studies is

described in step 1. Step 2 describes the methodology for the final selection of

cement-superplasticizer combination based on workability, strength and cost by

conducting trial tests on concrete.

Step 1

• The water demand and setting time of cement, and the density, solid content

and pH of superplasticizer are to be determined for recording the basic

properties for the quality checks, as well as for comparing different batches of

products.

• The Hobart mixer or a similar intensive mixer is considered to be essential for

the preparation of paste for selecting a compatible combination based on the

paste studies as well as for correlating with concrete results.

• The first criterion in the selection process is based on the ability to obtain a

well-defined saturation dosage from the Marsh cone test that is within the

maximum dosage of the superplasticizer as recommended by the supplier. The

study of the flow behaviour of superplasticized paste has shown that results

from the mini-slump and viscometer tests can be correlated with those of the

Marsh cone test results and, therefore, need not be carried out since they

169
require more sophisticated equipment or may not be as sensitive. The

correlation between the flow behaviour of paste, mortar and concrete has

shown that the dosage obtained from the paste tests can be used as the

guideline for selecting the dosage of superplasticizer for concrete.

• As seen from the studies on the loss of flow and UV absorbance, the flow

behaviour can vary significantly with time in the fresh paste. Therefore, it is

recommended that the saturation dosage be obtained at the age that

corresponds to the actual initiation of the placing of concrete (e.g., age when

the concrete is expected to be delivered to the site, in the case of ready-mixed

concrete).

• The influence of superplasticizer on the setting behaviour of the paste is the

second criterion for compatibility. Here, the final setting time is limited to 16

hours as obtained in the Vicat penetration test, unless there is deliberate

retardation through the use of a retarder or a superplasticizer with retarding

properties. The results of the study suggest that if this criterion is satisfied, the

FST of concrete would be not more than 12 hours since the difference between

the FSTs of paste and concrete was found to be normally not more than 4

hours. Since the electrical resistivity measurements require more technical

expertise than the Vicat penetration apparatus and there is only limited

experience with such tests, it is deemed to be unnecessary for determining the

compatibility. In any case, the IST obtained with Vicat apparatus correlates

well with the dormant stage extension obtained from electrical resistivity

measurements.

• The microstructural analysis is not included in the methodology since the

studies previously discussed show that a paste with adequate workability and

normal setting behaviour generally has a uniform microstructure.

170
Step 2

• Final trials are done on concrete, within this selection procedure, to

complement the paste tests and, more importantly, to ensure a mix with the

required slump and compressive strength. It is assumed that the paste volume

and the aggregate skeleton are optimized independently, for example with a

method based on packing density or other, as in the methods of Toralles-

Carbonari et al. (1996) and Gomes et al. (2001).

• The study has shown that the fines content in the aggregates can increase the

superplasticizer demand to more than the saturation dosage of the

superplasticizer, which also supports the necessity for the final trials on

concrete. In such cases, a higher dosage than the saturation point may be used

for achieving the desired workability; however, it is recommended that the

dosage is not more than 150% of the saturation dosage in order to limit the

retardation and possible segregation, as well as cost. Obviously, the

superplasticizer dosage can be reduced if the slump obtained is more than the

desired value when the saturation dosage is used.

• From the point of view of productivity, a minimum value of 50% of the 28-

day compressive strength is recommended at 3 days. This criterion can be

waived when the early age strength is not critical.

• The final choice of the superplasticizer can be based on the minimum cost

considering the dosage of the superplasticizer for obtaining the desired slump

among the compatible combinations. Superplasticizers can also be short-listed

for further testing based on tests on pastes, for example, by comparing flow

times and saturation dosages.

171
• Determine density, pH and solid content of superplasticizer, and compare with previous
data and manufacturer’s specifications. If there is disagreement, check with another batch.
• Determine the consistency and setting time of cement and compare with previous data and
manufacturer’s specifications. If there is disagreement, check with another batch.
• In case of anomalies, reject corresponding batch.

Mixing of the paste with Hobart mixer or similar

Determination of saturation dosage of superplasticizer using Marsh cone test,


at the required age.

• Is there a well-defined saturation point?


• Is the saturation dosage within the No
Step 1 manufacturer’s recommended limit?

Yes
Incompatible
Determine the final setting time (FST) combination
using Vicat test at saturation dosage

No
FST < 16 hours

Compatible combination

Mix for concrete with saturation dosage of the


superplasticizer

Check for the required slump. If slump is higher than required, decrease the dosage. If slump
is lower than required, increase the dosage up to a maximum of 150% of the saturation
dosage; if the required slump is still not obtained, reject the superplasticizer.

Determine the compressive strength at 3 days.


Reject if the strength is less than 50% of the
Step 2 target 28 day strength.

• Determine the cost of the superplasticizer per m3 of


the concrete
• Select the superplasticizer with the minimum cost
for the dosage required for the desired slump

Fig. 6.6 Methodology to select the Compatible Cement-Superplasticizer Combination

172
The summary of the paste test results along with the compatibility criteria derived

from the methodology for different combinations at 0-minutes and 60-minutes are

given in Tables 6.9 and 6.10, respectively. The nomenclature used for the mixes is as

follows; the first letter represents the type of cement followed by the admixture code,

e.g.,C1/LS-C1 is the mix with cement C1 and superplasticizer, LS-C1.

Table 6.9 Summary of Superplasticized paste combinations at 0-minutes

Saturation dosage
within the Final
Mix
manufacturer’s Setting Compatible?
compositions
recommended ≤ 16 hr
limit at 0 min
C1/LS-C1 Yes Yes Yes
C1/SNF-S1 Yes Yes Yes
C1/SNF-D1 Yes Yes Yes
C1/SNF-D2 Yes Yes Yes
C1/SNF-C1 Yes Yes Yes
C1/SMF-S1 Yes Yes Yes
C1/SMF-C1 Yes Yes Yes
C1/PCE S1 Yes Yes Yes
C1/PCE-D1 Yes Yes Yes
C2/LS-C1 No * No
C2/SNF-S1 Yes * Yes
C2/SNF-D2 Yes * Yes
C2/SMF-S1 Yes * Yes
C2/PCE-D1 Yes * Yes
C3/LS-C1 Yes Yes Yes
C3/SNF-S1 Yes Yes Yes
C3/SNF-D2 Yes Yes Yes
C3/SNF-C2 Yes Yes Yes
C3/SMF-S1 Yes Yes Yes
C3/PCE-D1 Yes Yes Yes
C4/LS-C1 No Yes No
C4/SNF-S1 Yes Yes Yes
C4/SNF-D2 Yes Yes Yes
C4/SNF-C2 No No No
C4/SMF-S1 Yes Yes Yes
C4/PCE-D1 Yes Yes Yes
*Could not be determined

The compatibility of different combinations are studied as per the criteria given in

step 1. Accordingly, C2/LS-C1 and C4/LS-C1 seem to be incompatible at 0-minutes

as the saturation dosages are more than the recommended limit. C4/SNF-C2 is also
173
incompatible at 0-minutes as per both criteria; the saturation dosage is above the

recommended limit and the setting time is more than 16 hours. The incompatible

combinations are marked with bold letters.

The test results at 60-minutes give more incompatible combinations due to the

significant loss of fluidity in some combinations (see Figure 6.1(a-d)). Accordingly,

C1/SMF-C1, C3/LS-C1 become incompatible as their saturation dosages are above

the manufacturer’s recommended limit whereas the absence of a well-defined

saturation dosage makes C2/LS-C1, C4/LS-C1 and C4/SMF-S1 incompatible. The

SNF-C2 is also incompatible due to the same reasons as mentioned for 0-minutes.

Table 6.10 Summary of Superplasticized paste combinations at 60-minutes

Saturation dosage within the Final


Mix
manufacturer’s recommended Setting Compatible?
Compositions
limit at 60 min ≤ 16 hr
C1/LS-C1 Yes Yes Yes
C1/SNF-S1 Yes Yes yes
C1/SNF-D1 Yes Yes Yes
C1/SNF-D2 Yes Yes Yes
C1/SNF-C1 Yes Yes Yes
C1/SMF-S1 Yes Yes Yes
C1/SMF-C1 No Yes No
C1/PCE S1 Yes Yes Yes
C1/PCE-D1 Yes Yes Yes
C2/LS-C1 No * No
C2/SNF-S1 Yes * Yes
C2/SNF-D2 Yes * Yes
C2/SMF-S1 Yes * Yes
C2/PCED1 Yes * Yes
C3/LS-C1 No Yes No
C3/SNF-S1 Yes Yes Yes
C3/SNF-D2 Yes Yes Yes
C3/SNF-C2 Yes Yes Yes
C3/SMF-S1 Yes Yes Yes
C3/PCE-D1 Yes Yes Yes
C4/LS-C1 No Yes No
C4/SNF-S1 Yes Yes Yes
C4/SNF-D2 Yes Yes yes
C4/SNF-C2 No No No
C4/SMF-S1 No Yes No
C4/PCE-D1 Yes Yes Yes
*could not be determined
174
The literature gives other approaches to identify the compatible cement-

superplasticizer combinations. According to Aïtcin (1998), in a cement-

superplasticizer combination, if the 0 minutes Marsh cone flow time curve and 60

minutes flow time curve intersect or are close to each other, it can be considered as a

compatible combination, and the saturation dosage is taken as the dosage at the point

of intersection of the curves. According to Aïtcin's criterion, all the combinations

studied here are compatible, except those with the superplasticizer LS-C1 with the

cements C2, C3 and C4. The combinations that are incompatible according this

criterion are also incompatible as per the proposed selection criteria, which is

however, more stringent (by taking into consideration the setting of the paste) leading

to more incompatible combinations.

Furthermore, Nkinamubanzi et al. (2000) has classified the compatibility of different

cements through their chemical compositions. They suggested the use of the SO3/C3A

ratio and alkali content for selecting compatible combinations. Figure 6.7 attempts to

classify the tested cements accordingly. According to Nkinamubanzi et al. (2000), a

high SO3/C3A ratio and high alkali content are considered as the requirements for

compatibility. Consequently, the cements C1, C2 and C3 are classified as compatible

and cement C4 is less compatible due to a low SO3/C3A ratio. Note that the SO3

present is considered to be completely soluble, which may not be always the case.

According to the proposed method (see test results in Tables 6.9 and 6.10), the cement

C4 has more incompatible combinations, and therefore, can be considered as less

compatible than the other cements, which is in accordance with the classification of

Nkinamubanzi et al. (2000).

175
Fig. 6.7 Classification of cement based on its chemical composition
(based on Nkinamubanzi, et al., (2000)

6.7 SUMMARY

The influence of the chemical composition of cement and family of superplasticizers

on the flow and setting behaviour of different pastes are analysed in the present

chapter. It is confirmed that both the fresh and hardening properties vary for different

combinations. In order to understand this difference in nature, the absorbance of the

admixtures has been determined, as the rate and amount of adsorption of

superplasticizers on cement surfaces is responsible for the effective dispersion. Also,

the UV absorbance at different times helps understand the mechanisms that cause the

loss of fluidity of superplasticized pastes. This study confirms that a good correlation

exists between the loss of fluidity with initial adsorption. In addition to that, the λmax

is found to be a characteristic of a particular superplasticizer and hence UV

absorbance can be used as a simple test to identify the family of superplasticizer.

Finally, a simple methodology based on the flow and setting behaviour of pastes is

suggested to identify the compatible cement-superplasticizer combinations and is

validated with other approaches from the literature.

176
CHAPTER 7

FLOW BEHAVIOUR OF SUPERPLASTICIZED PASTES


INCORPORATING OTHER ADMIXTURES

7.1 GENERAL

An objective procedure for determining the saturation dosage of superplasticizer in a

cement paste using the Marsh cone test has been explained in Chapter 3 for

understanding the flow behaviour of the paste. The experimental and analytical

correlation between the Marsh cone flow times with rheological parameters has also

been presented. The flow behaviour of the paste was found to relate well to the fresh

state properties of normal concrete showing the relevance of conducting the tests on

paste. These procedures applied for understanding the behaviour of superplasticized

cement paste are extended in the present Chapter to pastes with other admixtures such

as metakaolin and viscosity modifying agents.

The incorporation of mineral admixtures, viscosity modifying admixtures (VMA) and

superplasticizers in the paste composition results in a balance between the cohesion

and fluidity of special cement-based materials such as cement grouts, shotcrete,

underwater concrete and self compacting concrete (SCC) (Khayat and Guizani, 1997;

Lachemi et al., 2004, 2007). Superplasticizers generally decrease the yield stress

whereas viscosity modifying admixtures tend to increase the viscosity at the same w/c

ratio (Ferraris et al., 2001; Sonebi, 2006). Moreover, the interaction between different

admixtures in a cementitious system sometimes results in delays in setting time and

rapid slump loss, which demand a compatible combination to achieve the best

rheological properties (Khayat and Yahia, 1997).


The objective of this chapter is to understand the flow behaviour of superplasticized

paste with components such as mineral admixtures and VMAs. The work is of

particular interest for the development of self compacting concrete (SCC) as it

generally incorporates both the superplasticizer and viscosity modifying agent.

The present chapter details the procedures and results of tests performed with the

Marsh cone, mini-slump and viscometer on superplasticized pastes with metakaolin or

a VMA. For correlating the results from the tests on pastes, an experimental study of

the fresh behaviour of SCC was done through its characteristic tests like slump flow,

V-Funnel, J-Ring in addition to compression tests. Based on the results, the

methodology of Gomes et al. (2001) and Gettu et al. (2004) for the mix design is

extended to SCC incorporating a VMA.

7.2 MATERIALS USED IN THE STUDY

53 grade ordinary portland cement (cement C1), PCE-D1 and SNF-D2

superplasticizers, metakaolin and four types of VMAs are used in the study. The

characteristics of cement and superplasticizers have already been reported in

Chapter 3.

The commercial availability of metakaolin in India has increased in the last few years

due to extensive kaolin sources across the country (Kumar and Kaushik, 2003).

Metakaolin is a primary product formed by the calcination of kaolin clay at a

temperature of 650-800°C. It belongs to the chemical family of China clay-

aluminium silicate (Al2O3 2SiO2). It contains 50-55% SiO2 and 40-45% Al2O3. Its

specific gravity is 2.6, bulk density is 300 kg/m3 and specific surface is 12000m2/kg.

Additional, C-S-H is produced due to its reaction with Ca(OH)2 as indicated in

Equation 7.1:

AS2 + 6CH + 9H Æ C4AH13 + 2C-S-H (7.1)

178
Four types of VMAs were used in the study with the properties given in Table 7.1.

VMA 4 is diutangum whereas the others are liquids based on polysaccharides.

Table 7.1 Properties of viscosity modifying agents

Density
Name of VMA Solid content (%) appearance
(gm/ml)
VMA 1 1.02 2 liquid
VMA 2 1.00 2 liquid
VMA 3 1.01 2 liquid
VMA 4 - 100 powder

7.3 FLOW BEHAVIOUR OF PASTES WITH METAKAOLIN

Metakaolin is said to increase the total volume of finer pores and decrease the volume

of coarser pores; however, the total volume of pores is reported to be less than in pure

cement paste (Khatib and Wild, 1996). Its density is lower than that of cement

whereas the surface area is higher with irregular shape particles. Consequently, the

flow behaviour is affected by the incorporation of metakaolin with higher loss of

fluidity and more shear thickening when compared with pure cement paste (Cyr et al.,

2000; Corinaldesi and Moriconi, 2003; Li and Ding, 2003).

It has also been reported that metakaolin increases the water demand and apparent

viscosity of paste due to an increase in the solid fraction of paste. The superplasticizer

type appears to significantly affect the workability and compressive strength of mortar

and concrete incorporating metakaolin as the rate of pozzolanic reaction and hydration

depends on the type of superplasticizer; Kim et al. (2003) found that SNF blends and

PCE cause slump retention whereas the SNF and SMF cause slump loss. It was also

reported that the PCE retards the hydration reaction and hence causes a reduction in

compressive strength.

179
7.3.1 Effect of Metakaolin in Paste incorporating PCE Based Superplasticizer

For the pastes, cement and metakaolin were taken in the ratio of 90:10 and the

water/powder ratio (w/p) was taken as 0.35. The Marsh cone, mini-slump and

viscometer tests were conducted immediately after mixing (0 minutes) and after 60

minutes. The trend observed in the Marsh cone flow time with metakaolin is same as

that of cement-PCE combination with a decrease in flow time with an increase in the

dosage of superplasticizer. All the combinations seem to exhibit well defined

saturation points. The addition of metakaolin increases the saturation superplasticizer

dosage from 0.1% to 0.2% (in terms of solid content of sp/c%), as shown in Figure

7.1, which may be due to the increase in solid fraction and surface area in the paste.

The loss of fluidity is also observed to be higher than in the cement-superplasticizer

combination, which confirms earlier results (Li and Ding, 2003). The level of flow at

the saturation dosages are, however, practically the same.

1.22
1.20 Cement-PCE-0-minutes
Cement-PCE-60-minutes
1.18 Cement-metakaolin-PCE-0-minutes
Cement-metakaolin-PCE-60-minutes
1.16
log (flow time,sec)

1.14
1.12
1.10
1.08
1.06
1.04
1.02
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 7.1 Marsh cone flow time curves for pastes with
metakaolin and PCE based superplasticizer

180
The mini-slump spread increases with an increase in the dosage of superplasticizer in

the cement-superplasticizer-metakaolin combination, as seen in Figure 7.2. At 0

minutes, both the pastes exhibit bleeding beyond the saturation dosage, and hence the

corresponding data are not plotted in the graph. The addition of metakaolin decreases

the mini-slump spread, as shown in Figure 7.2.

200

180
Mini-slump spread (mm)

160

140

120 Cement-PCE-0-minutes
Cement-PCE-60-minutes
Cement-metakaolin-PCE-0-minutes
100 Cement-metakaolin-PCE-60-minutes

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %
Fig. 7.2 Variation of mini-slump spread with dosage of PCE for
metakaolin-PCE combination

The rheological study of the pastes showed that the addition of metakaolin results in

an increase in the yield stress at low dosages and remains constant beyond the

saturation dosage of superplasticizer, as seen in Figure 7.3. The plastic viscosity also

increases with the addition of metakaolin, especially at low dosages of

superplasticizer, as shown in Figure 7.4. With regards to both the parameters, the

addition of metakaolin does not produce any significant difference in the rheological

parameter beyond the saturation dosage. The observed trends are in disagreement with

the reported results of Ferraris et al. (2001), who did not find any significant increase

in the yield stress or viscosity due to the addition of metakaolin mix. However, the

181
present data are in accordance with the results of Cyr and Mouret (2003). It is possible

that the variations in the properties of the metakaolin lead to the differences in the

conclusions drawn from the different studies.

Cement-PCE 0
30 Cement-PCE 60
Cement-metakaolin-PCE 0
Cement-metakaolin-PCE 60
Yield stress (Pa)

20

10

0
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 7.3 Variation of yield stress with dosage of PCE for


metakaolin-PCE combination

200
Cement-PCE 0
180 Cement-PCE 60
Cement-metakaolin-PCE 0
Plastic Viscosity (cP)

160 Cement-metakaolin-PCE 60

140

120

100

80

60

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %
Fig. 7.4. Variation of plastic viscosity with dosage of PCE for
Metakaolin-PCE combination

182
7.3.2 Effect of Metakaolin in Paste Incorporating SNF Based Superplasticizer

The flow curves at 0-minutes and 60-minutes for the cement-metakaolin-SNF based

paste are shown in Figure 7.5. The flow curves at 0-minutes shows that the addition of

metakaolin increases the saturation dosage of SNF. In spite of the retarding effect of

the superplasticizer SNF-D2, there is a significant loss of fluidity within 60-minutes in

the paste with metakaolin and a well defined saturation dosage cannot be obtained for

the flow curve corresponding to 60-minutes. This may be either due to incompatibility

between the metakaolin and SNF, as reported by Kim et al. (2003), caused by the

absorption of SNF by the metakaolin. The addition of metakaolin decreases the mini

slump spread at 0-minutes as well as for 60-minutes (see Figure 7.6) confirming the

results of Kim et al. (2003). The viscometer test results also shows the same trend,

with an increase in yield stress and plastic viscosity with an increase in the time after

mixing (see Figures 7.7 and 7.8).

The comparison of test results shows that SNF causes more loss of fluidity than PCE

based superplasticizer and is significantly less compatible with the metakaolin than

the PCE. The results also show that a compatible cement-superplasticizer combination

can give incompatibility related problems, such as loss of fluidity, when a mineral

admixture is incorporated.

183
1.5
Cement-SNF-0-minutes
Cement-SNF-60-minutes
1.4 Cement-metakaolin-SNF-0-minutes
Cement-metakaolin-SNF-60-minutes
log (flow time, sec) 1.3

1.2

1.1

1.0

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %

Fig. 7.5 Marsh cone flow time curves for pastes with metakaolin and SNF-D2

Cement-SNF-0-minutes
180
Cement-SNF-60-minutes
Cement-metakaolin-SNF-0-minutes
mini-slump spread (mm)

160 Cement-metakaolin-SNF-60-minutes

140

120

100

80
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 7.6 Variation of mini-slump spread for pastes with metakaolin and SNF-D2

184
30 Cement-SNF-0-minutes
Cement-SNF-60-minutes
25 Cement-metakaolin-SNF-0-minutes
Cement-metakaolin-SNF-60-minutes
Yield stress (Pa)
20

15

10

0.0 0.2 0.4 0.6 0.8 1.0


sp/c %

Fig. 7.7 Variation of yield stress for paste with metakaolin and SNF-D2

270 Cement-SNF-0-minutes
240 Cement-SNF-60-minutes
Cement-metakaolin-SNF-0-minutes
210 Cement-metakaolin-SNF-60-minutes
Plastic viscosity (cP)

180

150

120

90

60

30
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

Fig. 7.8 Variation of plastic viscosity for paste with metakaolin and SNF-D2

7.4 INFLUENCE OF VMA ON THE FLOW BEHAVIOUR OF


SUPERPLASTICIZED PASTE

7.4.1 Characteristics of VMA

VMAs are water soluble polymers that increase the viscosity of the paste and enhance

the ability of the paste to retain its constituents in suspension (Khayat and Yahia,

1997; Lachemi et al., 2007). The main function of the VMA is to modify the

rheological properties of cement paste. As mentioned earlier, the rheological

behaviour of fluid cement pastes can be described by the yield stress and plastic

viscosity. The VMA changes the rheology by increasing the plastic viscosity but with

185
a small increase in yield shear stress. On the other hand, superplasticizers tend to

decrease the yield stress. In conjunction, the superplasticizer and VMA can be used to

optimize the yield shear stress. The test methods used are those explained earlier were

used here to understand the flow behaviour of paste incorporating VMA.

VMAs are either cellulose or polysaccharide based, and they are classified as natural,

semi-synthetic or synthetic polymers. Natural polymers include starches, guar gum,

locust bean gum, alginates, agar, gum Arabic, welan gum, xanthan gum, rhamsan

gum, and gellan gum, as well as plant protein. Semi-synthetic polymers include

decomposed starch and its derivatives, cellulose-ether derivatives such as

hydroxypropyl methyl cellulose (HPMC), hydroxyethyl cellulose (HEC) and carboxy

methyl cellulose (CMC), as well as electrolytes such as sodium alginate and

propylene glycol alginate. Synthetic polymers consist of polymers based on ethylene

such as polyethylene oxide, polyacrylamide, polyacrylate and those based on vinyl

such as polyvinyl alcohol (Khayat and Yahia, 1998). According to their physical

action, VMAs are classified into five categories (Rixom and Mailvaganam, 1999):

Class A: Water-soluble synthetic and natural organic polymers that increase the

viscosity of the mixing water. These types of materials include cellulose ethers,

polyethylene oxides, polyacrylamide, polyvinyl alcohol etc.

Class B: Organic water-soluble flocculants that become adsorbed on cement grains

and increase the viscosity due to enhanced interparticle attraction between cement

grains. It includes styrene copolymers with carboxyl groups, synthetic

polyelectrolytes and natural gums.

186
Class C: Emulsions of different organic materials that enhance interparticle attraction

and supply additional superfine particles in the cement paste. The materials belonging

to Class C are acrylic emulsions and aqueous clay dispersions.

Class D: Water-soluble inorganic materials having high surface area that increase the

water retaining capacity of the paste such as bentonites, silica fume, and milled

asbestos.

Class E: Inorganic materials having high surface area which increase the content of

fine particles in paste and hence the thixotropy. They include fly ash, hydrated lime,

kaolin, various rock dusts, and diatomaceous earth, etc.

7.4.2 Experimental Details

The methodology used for the mix design of SCC in the present study is an extension

of an approach proposed by Toralles-Carbonari et al. (1996) for high strength silica

fume concrete and later modified by Gomes et al. (2001) and Gettu et al. (2004) for

SCC. Within this approach, concrete is considered as a two phase material consisting

of a paste that provides fluidity and cohesion, and an aggregate skeleton that provides

the mechanical integrity. The flow behaviour of paste and concrete is studied in the

present work by varying the type and dosage of superplasticizer and VMA, while

keeping the other parameters such as temperature and cement type constant.

7.4.2.1 Study of the Interaction between Superplasticizer and VMA

The mixing method for the preparation of paste is same as that used in the preparation

of superplasticized cement paste, except that the VMA is added after three minutes

and it is further mixed for 2 minutes. The flow behaviour of the paste was determined

using Marsh cone, mini-slump and viscometer. Preliminary tests were conducted with

three different types of VMAs (VMA 1, VMA 2 and VMA 3) and a PCE based

superplasticizer at two w/c ratios (w/c = 0.5 and w/c = 0.35) to understand the
187
superplasticizer-VMA interaction. The saturation dosage of the superplasticizer was

determined as 0.2% and trials were conducted with a low dosage of 0.3% of different

VMAs. This study identified the VMA leading to the highest Marsh cone flow time

and the least mini-slump spread as VMA 1; see Figures 7.9 and 7.10 (VMA 0 denotes

the mix without VMA). This was considered as the most effective in terms of

achieving the cohesion compared to the other VMAs and hence it is used for further

studies. Also, it is observed that the effectiveness of the VMA increases with a

decrease in w/c as shown in Figures 7.9 and 7.10. This may be because, the molecules

in adjacent polymer chains can develop attractive forces at low w/c, further blocking

the movement of water, causing gel formation and a consequent increase in viscosity

(Khayat and Guizani, 1997). The plastic viscosity also shows the same trend as the

Marsh cone flow time and mini-slump spread and is given in Figure 7.11.

40
PCE- 0.2% VMA 0
35
0.3% VMA 1
Marsh cone flow time (sec)

30 0.3% VMA 2
0.3% VMA 3
25

20

15

10

0
w/c = 0.5 w/c = 0.35

Fig. 7.9 Marsh cone flow time for different viscosity modifying agents

188
180
PCE - 0.2% VMA 0
160
0.3% VMA 1
140
mini slump spread (mm)
0.3% VMA 2
120 0.3% VMA 3
100

80

60

40

20

0
w/c = 0.5 w/c = 0.35

Fig. 7.10 Mini slump spread for different viscosity modifying agents

18 100
PCE - 0.2%
w/c = 0.35 90
16 yield stress
14 plastic viscosity 80
70 Plastic viscosity(cP)
yield stress(Pa)

12
w/c = 0.5 60
10
50
8
40
6
30
4 20
2 10
0 0
VMA 0 VMA 1 VMA 2 VMA 3 VMA 0 VMA 1 VMA 2 VMA 3

type of VMA

Fig. 7.11 Values of yield stress and plastic viscosity for


different viscosity modifying agents

189
With the w/c ratio set at 0.35, Marsh cone flow time curves were determined for low

and high dosages of VMA 1 with different dosages of superplasticizer (PCE-D1) The

test results are given in Figure 7.12; where it can be observed that Marsh cone flow

time increases with an increase in the dosage of the VMA. For comparison, results are

also reported for a Welan gum based VMA, namely VMA 4, at the dosages that are

normally used. It is seen that VMA 4 leads to higher flow times than VMA 1 at both

dosages.

2.0 0.3% VMA 1


1% VMA 1
1.9 0.01% VMA 4
0.03% VMA 4
log ( flow time, sec)

1.8

1.7

1.6

1.5

1.4
0.0 0.1 0.2 0.3 0.4 0.5 0.6
sp/c%
Figure 7.12 Determination of saturation dosage of superplasticizer
at different dosages of VMA

The mini-slump spread, the spread time to reach a diameter of 115 mm (t115) and

rheological parameters were determined for different dosages of VMA with the

corresponding saturation dosages of superplasticizers from paste tests. The results are

summarized in Table 7.2. Gettu et al. (2004) recommend a value of about 2 seconds

for t115 to obtain a paste with adequate cohesion for SCC, which is obtained in all the

present cases. However, their requirement of a minimum spread of about 170 mm is

not met, except in the case of VMA 4 at a dosage of 0.01%.

190
Table 7.2 Flow characteristics of pastes with different types and dosages of VMA

Saturation Type and Flow Mini- t 115 Yield Plastic


dosage of dosage of time slump (sec) stress viscosity
SP (sp/c VMA (%) (sec) spread (Pa) (cP)
%) (mm)
0.2 VMA 1 - 0.3 34 133 2 4 126
0.3 VMA 1 - 1.0 66 112 3 3 130
0.2 VMA 4 - 0.01 49 168 2 0 282
0.2 VMA 4 - 0.03 68 140 2 * *
* values could not be measured

7.4.2.2 Effect of Time on the Flow Behaviour of Paste Incorporating


Superplasticizer and VMA

The effect of VMA to increase as well as to maintain the viscosity depends on the

type of superplasticizer and type of VMA (Khayat and Yahia, 1997). In order to

understand the efficiency of VMA to maintain the cohesion for longer period, flow

tests were conducted at 0, 30, 60 and 90 minutes with PCE-D1 based superplasticizer

and VMA 1 and VMA 4 at a w/c of 0.35. The saturation dosage of superplasticizer for

different combinations were determined as explained in Section 7.4.2.1.

The rheological parameters were also investigated at low and high dosages of VMA.

The Marsh cone test results of different combinations are shown in Figure 7.13 and the

summary of flow test results are given in Table 7.3. The test results at different ages

show that only in the paste with PCE-D1 and VMA 4, the Marsh cone flow time

increases with time. It is observed that in other combinations, instead of increasing, the

Marsh cone flow time decreases with time. The mini-slump test result also supports the

Marsh cone results; the mini-slump spread decreases with time in PCE-D1-VMA 4

combination. The t115 value also increases with time in the same combination. This

shows that even though VMA 1 is effective in giving cohesion and stability to the paste

initially, it is not effective in maintaining the cohesion and fluidity even for a period of

60-minutes. This is because the yield stress of paste in this combination is very low and

the mix loses its cohesiveness as the plastic viscosity decreases with time.
191
Here, it is seen that VMA 4 is more effective in increasing, as well as maintaining the

viscosity of the paste than VMA 1. This may be due to the difference in the chemical

structure of the two VMAs.

90
VMA 1 0.3%
80 VMA 1 1%
VMA 4 0.01%
Marsh cone flow time (sec)

70 VMA 4 0.03%

60
50
40
30
20
10
0
0 30 60 90
Time after mixing (minutes)

Fig. 7.13 Effect of time on Marsh cone flow curves of SP-VMA mix

Table 7.3 Influence of time on SP-VMA mix

Type and Mini-


Type and Time after Flow Plastic
dosage of slump t115 Yield
dosage of mixing time viscosity
SP spread (sec) stress (Pa)
VMA (%) (min) (sec) (cP)
(sp/c%) (mm)
0 27 115 2 4.5 121
PCE- D1 VMA 1 30 18 125 1 3.8 82.4
0.2% 0.3% 60 16 122 1 3.6 85.7
90 18 140 1 3.3 84.4
0 43 147 3 0 125
PCE- D1 VMA 1 30 38 155 2 0 107
0.3% 1% 60 37 145 1.5 0 93
90 35 154 1 0 95
0 36 148 1.5 1.6 235
PCE-D1 VMA 4 30 43 149 2 2.9 263
0.2% 0.01% 60 51 135 3 5.4 260
90 89 109 6 8.6 285
0 64 138 2 * *
PCE-D1 VMA 4 30 75 128 3 * *
0.2% 0.03% 60 84 123 4 * *
90 87 115 5 * *
* data could not be determined

192
7.5 COMPARISON OF THE FLOW BEHAVIOUR OF PASTE AND SELF
COMPACTING CONCRETE

7.5.1 Mix design and testing of SCC

The procedure for the determination of the saturation dosage of superplasticizer in the

paste phase with VMA has been explained in Section 7.4, with the intention of

obtaining a paste with good cohesion and fluidity, which are essential for the

flowability and resistance to segregation of highly workable concretes. Since adequate

flow behaviour of paste is fundamental for achieving concrete with adequate fluidity

and viscosity, tests have been performed on SCC to validate the paste test results, as

done for normal concrete in Chapter 4. The mix proportioning of SCC is performed

by separately determining the saturation dosage of superplasticizer for prescribed

dosages of VMA in the paste phase, optimizing the aggregate proportion for minimum

void content, and then the paste content of the concrete to get the final composition,

following the method of Gomes et al. (2001).

As the first step in the mix design of SCC, the paste phase is optimized for high

fluidity and adequate cohesion. In the method of Gomes et al. (2001), a paste with the

saturation dosage of superplasticizer, as determined from the Marsh cone test, and

meeting the criteria of mini-slump spread of about 170 mm and a t115 of about 2

seconds (Gettu et al., 2004) is considered to be appropriate for SCC. The pastes

studied in the previous section with the superplasticizer and VMA dosages given in

Table 7.2, and with a w/c of 0.35, satisfied the above-mentioned criteria for pastes,

except for the requirement of a spread of about 170 mm that is met only in the case of

0.01% dosage of VMA 4. Nevertheless, SCC mixes were developed for all the four

pastes for further evaluation.

193
For the definition of the aggregate skeleton, the maximum aggregate size was limited

to 20 mm in order to have high fluidity without segregation. The optimum aggregate

combination is selected by considering the packing density and lowest void content

with the assumption that minimum void content leads to the minimum paste volume,

porosity and shrinkage. However, concrete requires more paste content than the

minimum void content to ensure cohesion and fluidity. The aggregate phase is

optimized here by measuring the dry uncompacted density of aggregate mixes and

choosing the mix with least void content (Gomes et al., 2001). The method is same as

that given in ASTM C29/C29M (2007), except that the uncompacted weight is taken,

which is more suitable for SCC.

In the present study, the sand, and coarse aggregates with maximum grain sizes of 20

mm and 10 mm were dry mixed manually in a tray and placed in a 15 litre container

without any compaction. The proportions of the 5-10 mm : 10-20 mm fractions were

fixed as 3:2. The density or unit weight was measured to determine the void content,

as shown in Figure 7.14. It was observed that the sand content that gave the minimum

void content was 45 %, with a void content of 28%.

46 2000
44 Void ratio
42 Bulk density 1900

40
Bulk density (kg/m )
3

1800
Void content (%)

38
36
1700
34
32 1600
30
28 1500
26
0 20 40 60 80 100
Sand (%)

Fig. 7.14 Determination of aggregate proportions for concrete

194
The SCC has to satisfy certain requirements in the fresh and hardened state .The

filling ability and stability of SCC in the fresh state can be defined by four key

characteristics, which can be determined by the following test methods recommended

by EFNARC Guide lines (2005), Italian UNI standards and ASTM C1621 (2006)

code.

Flowability Slump flow test

Viscosity and segregation resistance T 500 or V-Funnel test

Passing ability J-Ring test

The following are the acceptance criteria of SCC that were prescribed based on the

above test methods.

Table 7.4 Acceptance criteria for SCC

Test values
Slump flow 550-700 mm
Slump flow spread time (T 500) 2-5 sec
V-funnel flow time 10 ± 3 sec
Difference in V-funnel flow time between ≤ 3 sec
measurements made immediately and after 5
minutes
Difference between J ring Flow and slump < 50 mm
flow
28 days compressive strength >50 MPa

In the present study, the slump flow, V-funnel and J-ring tests have been conducted

for characterizing the self compacting concrete (EFNARC Guidelines for SCC, 2005).

In addition to that the compressive strength of concrete has been determined at 3 days,

7 days and 28 days.

Slump flow test is a simple method to evaluate the filling ability of SCC under its own

weight. The test method is shown in Figure 7.15. The slump cone is filled with

concrete without compaction and is lifted to allow the concrete to flow over a plain

195
surface. The time for 50cm spread (t50) is measured along with the final spread as the

average of two perpendicular diameters. The concrete is visually examined for the

segregation and bleeding.

Fig. 7.15 (a) Slump cone measurements (b) Slump Cone test

A combination of slump flow with a ring denoted as J-Ring has been used to study the

blockage due to reinforcement (see Figure 7.16). As in the case of slump flow test, the

slump cone kept inside the J-Ring is filled with concrete. The concrete flows through

the gaps as the cone is lifted. The final spread is measured in two perpendicular

directions which gives an idea about the passing ability of concrete.

(a)

196
(b)

Fig. 7.16 (a) J-Ring measurement (b) J-Ring test

The V-funnel test method consists of measuring the time taken for a certain volume

(approximately 10 litres) of concrete to flow through a funnel (see Figure 7.17). The

funnel is again filled with the same concrete and waited for 5 minutes. The outlet is

opened after 5 minutes and the time taken for it to flow is noted. The difference

between the initial and final time should be less than 5 seconds so as to satisfy the

characteristic of SCC. Thus the test gives a good indication of viscosity of the paste as

well as an indication of segregation.

Fig. 7.17 V-Funnel test


197
The trial mixes showed that adequate paste content is needed for achieving a

satisfactory SCC. Here, the paste volume was set as 38%, which is about 10% more

than the void content determined in the aggregate compaction tests. The mix

proportions used for the SCC tests are given in Table 7.5; the paste compositions are

the same those characterized in Section 7.4 except that the superplasticizer dosage has

been varied to evaluate the influence of the same. The aggregates used have been

described already in Chapter 5.

Table 7.5 Proportion of materials used in the SCC trials

Fine Coarse aggregate Superplasticizer –


Cement Water
aggregate PCE-D1 VMA (%)
Kg/m3 Kg/m3 10 mm 20 mm
Kg/m3 (sp/c %)
Kg/m3 Kg/m3
0.3%
0.1,0.2,0.25, 0.3
VMA 1
1%
0.25, 0.3, 0.35, 0.4
VMA 1
550 193 800 494 332
0.01%
0.1,0.2,0.25, 0.3
VMA 4
0.03%
0.2, 0.25, 0.3, 0.4
VMA 4

SCC was prepared with different dosages of superplasticizers for the different SP-

VMA combinations, and with low and high dosages of VMA in each combination.

The mixing procedure is same as that of normal concrete except that the total mixing

time is 7 minutes in which VMA was added after 4 minutes. The slump flow, V-

funnel and J-ring tests are used for studying the fresh state properties. In addition, 3, 7

and 28-day compressive strength were also determined.

The results are given in Table 7.6, with the mixes satisfying the acceptance criteria of

SCC denoted in bold letters. It is observed that in all the combinations, a dosage equal
198
to or higher than the saturation dosage is needed for obtaining an adequate slump

flow. Also, other acceptance criteria are not satisfied if the dosage is lower than the

saturation superplasticizer dosage. The superplasticizer dosage slightly higher than the

saturation dosage results in t50 value less than one second, showing that the mix is not

viscous enough for a stable concrete. Also the passing ability of concrete is not

improved with the extra superplasticizer. An overdosage of the PCE based

superplasticizer results in a reduced slump flow in some combinations as reported in

the paste study. This is attributed to the shear thickening and immediate stiffening of

the paste.

Table 7.6 Test results of SCC with various SP-VMA combinations

Saturation Slump V- V- J ring Compressive


Combi- sp/c T50
dosage of Flow Funnel Funnel spread strength (MPa)
nation (%) (Sec)
sp/c from (mm) Time0 Time5 (mm) 3 7 28
paste(%) (Sec) (Sec) days day days
0.1 290 * * * * 33 38 43
VMA 1 0.2 665 2.1 9 9.1 590 36 47 56
0.3% 0.2
0.25 675 1 4.3 5.2 610 36 39 49
0.3 715 <1 4.81 5.93 710 42 39 54
0.25 460 * 6.5 9.3 450 36 44 49
VMA 1 0.3 475 * 7.5 11 455 39 44 56
1% 0.3
0.35 580 2 5.1 5.6 575 38 43 59
0.4 565 1.41 4.9 5.7 555 37 43 52
0.1 * * 23.7 * * 27 37 47
VMA 4 0.2 600 3.2 7.1 9 570 38 44 52
0.01% 0.2
0.25 680 1.5 7 9.3 668 37 42 52
0.3 775 <1 6 6.2 650 39 44 53
0.2 545 2 12 14.3 460 39 48 54
VMA 4 0.25 645 2.2 15.4 15.6 605 40 50 58
0.03% 0.2
0.3 713 1.5 13 13.2 630 40 48 59
0.4 663 2.1 12 14 663 40 49 60

The best performance for each combination is observed at or slightly higher than the

saturation dosage of superplasticizer obtained from the paste studies; an increase in

dosage of 0.05% was required for the PCE at higher dosage of VMAs. The only

combination that satisfies all the criteria is with 0.01% of VMA 4, which was also the

199
paste that came closest to satisfying the criteria of Gettu et al. (2004). The

compressive strength results show that the strengths achieved are satisfactory in all

the mixes at the corresponding saturation dosages. The study shows that paste test

results can be used as guidelines for selecting the best combination of SP and VMA

with appropriate dosages for developing SCC.

7.6 SUMMARY

• The flow behaviour of paste composed of cement, superplasticizer, mineral

admixture and VMA was investigated using the Marsh cone, mini-slump and

viscometer tests. It is seen that Marsh cone test along with mini-slump spread

is required to understand the complete flow behaviour of the paste with these

compositions.

• The addition of mineral admixtures and VMA generally increases the flow

time with the corresponding change in rheological properties (i.e., the yield

stress and plastic viscosity increase).

• The study confirms that the interaction between other admixtures and

superplasticizer depends on the type and dosage of admixture and

superplasticizer, which in turn affects the fluidity and slump retention.

Generally, the paste with the PCE has good initial fluidity and maintains the

fluidity better than the paste with the SNF in all the tested compositions with

the mineral admixture. Among the different SP-VMA combinations studied, it

is seen that the PCE-VMA 4 combination is able to maintain the cohesion over

time unlike the other combinations. Consequently, it is emphasized that the

compatibility of the SP-VMA combination should be studied prior to use in

concrete since it can affect the time dependent behaviour of the fresh concrete

and result in loss of cohesion with time, in as early as 30-minutes. The slump

200
flow values gradually decrease and the viscosity increases as time passes, in

most of the cases.

• A mix design procedure for developing SCC with VMA has been presented

based on the flow tests of pastes. The study shows that the type and dosage of

superplasticizer and VMA together govern the viscosity of SCC. The

saturation dosage of superplasticizer obtained from paste studies were found to

give good results in SCC. Consequently, it can be confirmed that the

optimization of superplasticized cement paste can be used as a guideline in the

preliminary selection of chemical admixtures for SCC.

201
CHAPTER 8

CONCLUSIONS AND RECOMMENDATIONS FOR


FURTHER RESEARCH

8.1 GENERAL CONCLUSIONS

Overall, the study was able to improve the understanding of the mechanisms of

cement–superplasticizer interaction and their influence on concrete performance.

Moreover, the effectiveness of the Marsh cone test in characterising the fluidity of

cement paste and the factors to be considered while establishing the correlation

between flow behaviour of paste and concrete were brought out. The salient

conclusions that correspond to the fulfillment of the principal objectives of this work

are given below.

• The preparation of superplasticized cement paste with four mixing methods

and the comparison of the saturation dosages obtained from these mixing

methods with “ball milling” to simulate the mixing of concrete, demonstrate

the influence of mixing on the flow behaviour of the paste. The dependence of

the fluidity of the paste and the saturation dosage of the superplasticizer on the

intensity of mixing suggests that a paste prepared using Hobart with B-flat

beater is representative of the paste phase of concrete.

• The establishing of a correlation between the empirical methods with the

rheological parameters shows that the same trends are obtained in the Marsh

cone flow time, the yield stress and plastic viscosity when there is an increase

in the dosage of the superplasticizer. The prediction of Marsh cone flow time

using rheological parameters further confirms the correlation.


• The correlation between the fluidity and setting of the fresh paste and that of

the mortar and concrete through standard tests shows that the trends are

similar in all cases. The comparison of the flow behaviour at 60-minutes

shows that the aggregates increase the loss of fluidity. Further, the presence of

fines in the coarse aggregate increases the superplasticizer demand for

adequate workability. However, it is confirmed that the optimization of the

paste phase is a preliminary step in the mix design of concrete and the

appropriate type and dosage of the superplasticizer can be selected on the basis

of tests on pastes.

• The use of multiple characterization techniques instead of a single method is

an effective approach for the analysis of the effect of superplasticizers on the

hydration processes as well as on the microstructural development. The

complementary use of XRD, SEM, DTA/TG/DTG and Si NMR has given an

insight into the properties of hardened cement paste. The retardation of the

setting of the superplasticized paste is attributed to the delay in the formation

of ettringite, as well as with the reduction in the silicate polymerization in the

presence of superplasticizers. The delay in hydration is also reflected by the

presence of more unhydrated particles in the micrographs of the

superplasticized pastes.

• The major characteristics of cement and superplasticizer affecting the cement-

superplasticizer interaction have been identified. The mechanisms of action of

the superplasticizer affect its effectiveness and consequently its compatibility

with cement. The major characteristics of the cement affecting the

compatibility are identified as SO3/C3A ratio, alkali content and fineness.

  203
Finally, a methodology, based on the test results, is suggested as a guideline

for the selection of an appropriate cement-superplasticizer combination.

8.2 SPECIFIC CONCLUSIONS

The specific conclusions derived from the experimental and analytical investigations

on cement-superplasticizer interaction are classified under the following topics: (i)

selection of a suitable mixing method for the preparation of paste, (ii) selection of

type and dosage of superplasticizer based on Marsh cone test, (iii) rheological

characterization of superplasticized paste using viscometer and rheometer, (iv)

establishing experimental and analytical correlation between Marsh cone test with

rheological parameters, (v) validation of paste test results with normal concrete and

self compacting concrete, and (vi) influence of superplasticizers on hardening and

hardened state properties of concrete.

A number of tests on paste have been performed using the Marsh cone and

Viscometer to confirm the effectiveness of simple test methods in bringing out the

influence of superplasticizers on the flow behaviour of paste. The rheological

parameters like yield stress and plastic viscosity provided ample information about the

flow behaviour of paste immediately after mixing as well as at 60-minutes. The

complementary use of multiple techniques to understand the effect of

superplasticizers on hardened properties validates the test results even with a few

number of samples.

8.2.1 Influence of Mixing Method on the Flow Behaviour of Paste

• The type of mixing affects the fluidity as well as the saturation dosage of

superplasticizer. Hence, the selection of a proper mixing method for the

preparation of paste is required while correlating the superplasticized paste

  204
flow behaviour with that of concrete.

• The study helps to classify the different mixing methods as less intensive,

intensive, and more intensive; less intensive mixing like hand mixing leads to

higher saturation dosage and low fluidity, intensive mixing using Hobart with

B-flat beater, reduce the superplasticizer demand in terms of saturation

dosage, more intensive mixing like blender mixing even though leads to

highest fluidity, and gives higher saturation dosage than the intensive mixing.

• The study confirmed that the paste prepared with the Hobart mixer with B-flat

beater simulates the paste phase of concrete. This clarifies the conflicting

views in the literature regarding the preparation of paste with blender and

Hobart mixer, for establishing a correlation between paste and concrete

behaviour.

8.2.2 Influence of Superplasticizers on the Flow Behaviour of Paste

• A detailed study of the influence of type and dosage of superplasticizers on the

flow properties of different compositions along with the determination of the

saturation dosage of superplasticizer has been conducted. The Marsh cone test,

a simple but effective method, has been used for measuring the fluidity. The

results of this test are found to have trends comparable to the rheological

parameters (i.e., yield stress and plastic viscosity) obtained from viscometric

tests.

• Even though the mini-slump test is also used to understand the flow behaviour,

the study showed that, in pastes with high fluidity, it is not useful for

identifying the influence of superplasticizers effectively.

  205
• The saturation dosages of a polycarboxylate based superplasticizer with the

tested cements are confirmed to be less than lignosulphonates, melamines and

naphthalenes. However, the loss of slump in concrete with a polycarboxylate

based superplasticizer is more than in concrete with a naphthalene based

superplasticizer.

• The influence of the saturation dosage of superplasticizer on the loss of

fluidity, and overdosage on settling and bleeding throws light on the selection

of appropriate dosage of superplasticizer in any combination. The study on

loss of fluidity of paste shows that if the fluidity at a later stage is more critical

to the application than the immediate fluidity, then the saturation dosage

corresponding to 60-minutes or any other critical time is required for

maintaining the fluidity for that duration.

• The tests using a rheometer with parallel plate attachment showed that the

nature of the superplasticized paste changes from viscoelastic fluid-like

behaviour to linearly viscous flow with an increase in the dosage of the

superplasticizer.

• The addition of mineral admixtures and VMA makes the paste more cohesive,

and increases the yield stress and viscosity as their dosages increase.

Therefore, the saturation dosage of the studied superplasticizers were all

observed to increase with the addition of the admixtures.

8.2.3 Comparison between Marsh Cone Flow time and Rheological Parameters

• By applying the Bingham and Herschel Bulkley models to the flow behaviour

of superplasticized paste, it is seen that an increase in the dosage of the

  206
superplasticizer leads to a significant decrease in the yield stress and plastic

viscosity up to saturation dosage, after which they remain practically constant.

• The prediction of Marsh cone flow time using rheological models shows that

the Marsh cone test results have a fundamental basis. Therefore, it is justified

that it is suitable for evaluating the relative fluidity of superplasticized paste.

8.2.4 Effect of Type and Dosage of superplasticizer on the Retardation of Setting

• The Vicat penetration test results confirmed that the superplasticizer retards

the setting of cement paste due to its effect on hydration. The present study

with saturation dosage of superplasticizer showed that amount of retardation

mainly depends on the type of superplasticizer. The PCE and LS based

superplasticizers retard more than other families.

• Electrical conductivity measurements on the cement paste have been used to

characterize the setting behaviour of cement paste. It has been observed that

the superplasticizer leads to retardation of the setting process, which is mainly

manifested as an increase in the duration of the dormant stage; the

superplasticizer that gives the highest initial setting time leads to the longest

dormant period.

• Even though the superplasticizer retards the initial setting, the duration

between the initial and final setting is less than in pure cement paste and is

comparable for different superplasticizers. This shows that superplasticizer

does not prolong the setting process once it is initiated.

8.2.5 Correlation between Paste, Mortar and Concrete

• It is found that the Marsh cone test is a satisfactory tool for comparing the

flow behaviour of different pastes and their components, and for defining the

  207
saturation dosage of the superplasticizer. The saturation dosage of

superplasticizer immediately after mixing and after 60-minutes is observed to

be generally different. Hence, the saturation dosage corresponding to 60-

minutes flow time curves is more suitable to select a superplasticizer for

maintaining the fluidity for more than one hour.

• The mortar flow tests show that the incorporation of sand does not increase the

saturation dosage significantly in the context of the present study. However,

the loss of fluidity of mortar is higher than that of paste at low dosages and has

to be considered while comparing the mortar behaviour to that of paste phase.

• Flow tests on normal concrete and self compacting concrete show that a slight

increase in the saturation dosage of superplasticizer is required for adequate

workability in all combinations. This is attributed to the excess of fines in the

coarse aggregate used here.

• The comparison of the setting of paste and concrete showed that even though

the setting times are not directly related, the trends of the setting of paste and

concrete with superplasticizers are comparable. That is, SMF-S1 causes the

least retardation and PCE-D1 results in higher retardation than SNF-S1 in both

paste and concrete.

• The compressive strength increases with an increase in dosage of

superplasticizer and is maximum for the highest slump obtained without any

segregation or bleeding. This emphasizes the importance of good workability

in achieving adequate strength. The decrease in strength observed beyond

certain dosage is due to bleeding and segregation.

  208
• The cost of concrete derived from the individual prices of superplasticizer,

cement and aggregate shows that low unit cost of superplasticizer need not

always result in cost effective concrete. For example, in the studied

combinations, even though the unit cost (cost/kg) of SMF-S1 is less than PCE-

D1, the former leads to more cost per m3 of concrete because the dosage of the

superplasticizer required for a particular slump is higher for SMF-S1.

8.2.6 Effect of Superplasticizer on the Hydration process and on the


Microstructural Development

• Instead of a single technique, multiple techniques (XRD, SEM, thermal

methods and Si NMR) have been used complementarily for understanding the

effect of superplasticizers on the hydration process and consequently on the

microstructure.

• The XRD study gives a qualitative evaluation of the effects of

superplasticizers on the hydration. It confirms that the superplasticizer affects

hydration and delays the formation of ettringite. In addition to that, it shows

that secondary ettringite is present even upto 28 days in some combinations.

The evaluation shows that the amount of C-H increases with the age of curing.

The presence of C-H peaks at half setting shows that the formation of C-H is

not delayed as the formation of ettringite in the presence of the

superplasticizers.

• The SEM study shows that the incorporation of a superplasticizer leads to more

dense and uniform structure than pure cement paste. However, the amount of

unhydrated particles at 3 days in the superplasticized paste is more than that of

pure cement paste confirming that the superplasticizer retards the hydration.

  209
• The DTA/TG/DTG analysis also confirms the XRD results that the C-H

formation is less affected by the addition of superplasticizers. However, the

mass percentage of C-H is very low for a quantitative comparison.

• The NMR results that show that all the superplasticizers studied here decrease

the rate of silicate polymerization and lead to a lower amount of polymerized

silicates at 3 and 28-days which is indicative of the change in the morphology

of C-S-H gel in the presence of superplasticizers.

• As explained in the previous section, the addition of superplasticizer always

leads to a slightly higher strength in concrete. However, the physico-chemical

analyses indicate that polymerization of silicates is retarded and the amount of

C-H formed is lower at early ages. Hence, it can be concluded from the

microstructural studies that more uniform and dense structure of

superplasticized paste results in slightly higher compressive strength of

concrete.

8.2.7 Influence of Chemical Composition of Cement on Compatibility

• Based on this study it is recommended that specification for the alkali content

be included in IS code for the chemical composition of cement as the amount

of superplasticizer required varies with alkali content; low alkali cement

requires more amount of superplasticizer due to more initial adsorption.

• The absorbance study conducted using a UV spectrophotometer shows that it is a

good characterization technique for identifying the superplasticizers. This study

also suggests that the reason for the increased demand of certain superplasticizers

is that there is higher initial adsorption of these superplasticizers.

  210
8.3 Guidelines for Selecting the Best Combination of Cement and
Superplasticizer

Based on the present study, a methodology is recommended for the selection of the

superplasticizer for concrete mix design. The methodology uses criteria based on the

Marsh cone and Vicat penetration tests to select a superplasticizer that is compatible

with the cement used. A compatible combination should have a well defined

saturation point and the dosage corresponding to the same should be within the

manufacturer’s recommended limit. Also, the final setting time of superplasticized

paste should be less than 16 hours, which is expected to lead to a concrete with a

setting time less than 12 hours. However, final trials have to be conducted on concrete

to select a cost-effective superplasticizer. The combinations studied here are used to

illustrate the evaluation for compatibility based on this methodology, and the results

are compared with other approaches in the literature.

8.4 Recommendations for Further Research

The study clarifies many of the conflicting views on the flow behaviour of paste,

effectiveness of the Marsh cone test in representing the flow behaviour of the

superplasticized paste, the correlation between the flow behaviour of paste and

concrete and the influence of superplasticizers on the hardened properties of concrete.

Continued research on the following aspects can give better understanding and be of

more use to the construction industry:

• The influence of the fines content in the aggregates on the saturation dosage as

well as the loss of fluidity with time.

  211
• Regarding the microstructure and hydration processes, the hydration processes

could be studied further by incorporating the quantitative measurement of the

different phases involved in the hydration through XRD.

• The determination of the Ca/Si ratio of the C-S-H and image analysis of more

samples can give a clearer picture of the effect of superplasticizers on the

composition of C-S-H and pores in the microstructure. An extensive study on

the pore structure of the superplasticized paste is essential for understanding

the effect of superplasticizers on the shrinkage of concrete.

• The prediction of the Marsh cone time with the available rheological models is

suitable only for a certain range of fluidity. In the cases of pastes with low

flow, the models seem to be inappropriate. In addition to that, the viscoelastic

characterization shows the change in nature of the paste from non-linear

viscoelastic to viscous with variation in the type and dosage of

superplasticizer. Hence, an appropriate constitutive model is required for the

whole range of flow.

• The interaction of the superplasticizer with mineral admixtures and VMA

requires further study for the optimization of paste as well for understanding

the loss of fluidity. In addition, the influence of other admixtures like

shrinkage reducing and accelerators on the cement-superplasticizer interaction

need further attention.

• Absorbance based studies of superplasticizer adsorption can be extended for

understanding the influence of mineral admixtures and VMAs on the amount

and rate of adsorption of superplasticizers on cement surface which in turn

  212
affects the loss of fluidity. The zeta potential of different pastes could be

measured for validating the adsorption study.

Finally, the number of tests on fresh cement paste are quite satisfactory for definite

conclusions to be made whereas in the studies performed on hardened cement paste,

the number of tests has been quite limited, which makes unambiguous conclusions

difficult. Therefore, more extensive tests can be performed on the properties of

hardened concrete for confirming the trends seen in paste.

  213
APPENDIX A

Table A.1 Test results of three trials in the selection of mixing method

Optimum superplasticizer
Type of mixing dosage (sp/c%)
trial 1 trial 2 trial 3
Hobart with B-flat beater 0.10 0.13 0.09
Hobart with D-wire whip 0.19 0.16 0.12
Blender 0.13 0.14 0.16
Grinder 0.15 0.16 0.07
Hand 0.39 0.20 0.06
Ball mill mixing (steel balls) 0.07 0.16 0.11
Ball mill mixing (Glass marbles) 0.15 0.16 0.14

Table A.2 Comparison of superplasticizer dosages


(data from Marsh cone and viscometer tests)

Super- Dosage Dosage Optimum Dosage Dosage from


plasticizer for for dosage corresponding Aïtcin
constant constant corresponding to 140º (at 60 method using
yield plastic to 140º (at 0 minutes) flow-time
stress viscosity minutes) curves at 0
and 60
minutes
LS-C1 0.4 0.4 0.25 0.26 0.3
SNF-S1 0.3 0.3 0.20 0.34 0.3
SNF- D1 0.4 0.5 0.24 0.23 0.3
SNF-D2 0.4 0.4 0.23 0.46 0.4
SNF- C1 0.3 0.3 0.16 0.20 0.3
SMF-S1 0.3 0.3 0.23 0.30 0.5
SMF-C1 0.5 0.5 0.20 0.30 0.5
PCE-S1 0.3 0.3 0.20 0.24 0.4
PCE-D1 0.1 0.1 0.07 0.17 0.2
14
LS-C1 LS-C1 120
1.3 log (flow time, sec)
1.3 Marsh cone flow time 12

plastic viscosity (cP)


plastic viscosity

log (flow time, sec)


log (flow time, sec) yield stress 100
10

yield stress (Pa)


1.2 80
1.2 8
6 60
1.1
1.1 4
40
2
1.0 20
1.0 0 0.0 0.2 0.4 0.6 0.8 1.0
0.0 0.2 0.4 0.6 0.8 1.0
sp/c %
sp/c%

(a) Correlation between Marsh cone flow time and rheological parameters for LS-C1

1.5
SNF-D2 1.5
SNF-D2 200
Marsh cone flow time 16
1.4
log (flow time, sec)

Yield stress Marsh cone flow time

plastic viscosity (cP)


1.4 160
yield stress (Pa)

log (flow time, sec) Plastic viscosity


12
1.3 120
1.3
8
1.2 80
4 1.2
40
1.1
0 1.1
0.0 0.2 0.4 0.6 0.8 1.0 0
sp/c % 0.0 0.2 0.4 0.6 0.8 1.0
sp/c %

(b) Correlation between Marsh cone flow time and rheological parameters for SNF-D2

35
SMF-C1 SMF-C1 200
1.20 30 1.20

plastic viscosity (cP)


Marsh cone flow time
log (flow time, sec)

log (flow time, sec)

25 Marsh cone flow time 160


yield stress (Pa)

yield stress
1.16 1.16 Plastic viscosity
20
120
15 1.12
1.12 80
10
5 1.08
40
1.08
0
1.04 0
0.2 0.4 0.6 0.8 1.0 0.2 0.4 0.6 0.8 1.0
sp/c % sp/c %

(c) Correlation between Marsh cone flow time and rheological parameters for SMF-C1

215
PCE-S1 PCE-S1 160
1.4 16 1.4
Marsh cone flow time 140

plastic viscosity (cP)


Marsh cone flow time
log (flow time, sec)

log (flow time, sec)


yield stress (Pa)
Yield stress Plastic viscosity
1.3 12 120
1.3
100
8
1.2 1.2 80
4 60
1.1 1.1
0 40
1.0 1.0 20
0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0
sp/c % sp/c%

(d) Correlation between Marsh cone flow time and rheological parameters for PCE-S1

Fig. A.1 Correlation between Marsh cone flow time and


rheological parameters of superplasticizers

35 0.25
30
0.20
Plastic viscosity (cP)
25
Yield stress (Pa)

20 0.15
15
0.10
10
5 0.05

0 0.00
0 10 20 30 40 50 60 70 0 10 20 30 40 50 60 70
Flow time (sec) Flow time (sec)

(a) Correlation between Mash cone flow time, yield stress and
plastic viscosity for different combinations

35 0.25
30
0.20
Plastic viscosity (cP)

25
Yield stress (Pa)

20 0.15
15
0.10
10
5 0.05
0
0.00
80 100 120 140 160 180 200 80 100 120 140 160 180 200
mini-slump spread (mm) mini-slump spread (mm)

(b) Correlation between Mini-slump spread, yield stress and


plastic viscosity for different combinations

Fig. A.2 Correlation between Marsh cone flow time and mini-slump spread
with rheological parameters for different compositions

216
APPENDIX B
Matlab programme to calculate flow time as per Nguyen model

%to calculate the velocity at the exit of the nozzle


%to calculate the time for falling the level from h1 to h2
clc
n=input('enter value of n = ');
R=0.004;
rho=input('enter value of rho = ');
g=9.81;
H=0.06;
tanphi=0.267;
lamda=0.96;
tuo0=input('enter value of tuo0 = ');
k=input('enter value of k = ');
alpha1= (1-(1.1*(tuo0/(rho*g*H*tanphi))));
alpha2=(1+(R/(3*n*H*tanphi)));
alpha=alpha1/alpha2;
syms x real
deltap=((alpha*rho*g)*(1+x/H)-rho*g)*H;
r0=2*tuo0/(rho*g+(deltap/H));
f1=1+((2*n)/(2*n+1))*(r0/R);
f2=((2*n^2)/((n+1)*(2*n+1)))*(r0/R)^2;
f3=(1-(r0/R))^((n+1)/n);
f=(f1+f2)*f3;
v1=n*R/(3*n+1);
v2=(alpha*rho*g*R)/(2*k);
v3=(H+x)/(H);
v=(v1*(v2*v3)^(1/n))*f;
Re=(rho*((v)^(2-n))*(2*R)^n)/k;
B1=(1/(2*lamda))*sqrt((alpha*R)/H);
B2=(n/(2*(3*n+1)))^(n/2);
B3=(1-(exp(-0.1*sqrt(Re))));
B4=(sqrt(Re)+(1.5/n)*(Re^(1/4)));
B=(B1*B2*B3*B4)+1;
vb= v/B;
t1=(R+(x*tanphi))^2;

217
t2=(vb)*R^2;
t3=t1/t2;
ii=linspace(0.224,0.165,100);
stepsize=ii(2)-ii(1);
for jj=1:length(ii)
fval(jj)=subs(t3,ii(jj));
end
cumsum=0;
for jj=2:length(ii)
cumsum=cumsum+(fval(jj)+fval(jj-1));
end
time=-((stepsize/2)*cumsum)

218
APPENDIX C

INFLUENCE OF SUPERPLASTICIZER ON THE NON-


NEWTONIAN CHARACTERISTICS OF CEMENT PASTE

C.1 GENERAL

The superplasticized cement paste usually exhibits non-Newtonian characteristics, which

depend on the type of superplasticizer and its dosage. The superplasticizer causes an

increase in the fluidity of the paste and shear thinning behaviour as explained in Chapter

3. However, in some instances of incompatibility, excessive amount of superplasticizer is

added, which can result in possible shear thickening. This shear thickening may be due to

high concentration of solids in nonflocculated suspensions as the hydration process is

accelerated during shearing (Struble and Sun 1995; Cyr et al. 2000). Also, an

agglomeration of excessive amount of superplasticizer available in the solution causes

more elasticity and possible shear thickening (Papo and Piani 2004).

Typically rotational viscometers have been used for characterizing the rheological

behaviour of cement paste as explained earlier (Roy and Asaga 1979; Masood and

Agarwal 1994; Ferraris et al. 2001; Park et al. 2005; Schwartzentruber et al. 2006). Most

of these investigations subject the material to a cycle consisting of shear rate increasing

and decreasing in a stepwise manner. Normally the cycles are repeated and the shear

stress-shear rate relationship in the down ramp of either the first, second or third cycle is

considered for further analysis (Papo and Piani 2004; Roussel and Roy 2005). Current

attempts in the cement paste rheology literature have mainly focused on reducing the data

from steady shear experiments to standard constitutive models, and exploring the

influence of superplasticizer dosage and type based on these model parameters as seen in

219
Chapter 3. A variety of models such as Bingham, Herschel- Bulkley, Power law models

etc., are used for characterizing the response of the material. The main intent of most of

these investigations have been to study the influence of these admixtures on some of

these model based parameters such as plastic viscosity, yield stress, consistency, flow

index etc. For instance, the influence of the dosage of superplasticizer on the Bingham

model parameters like yield stress and plastic viscosity shows that both of these

parameters decrease with increase in dosage of superplasticizer up to saturation dosage

and remains constant thereafter (Asaga and Roy 1980; Banfill 1981; Nehdi and Rahman

2004). Similar is the case with Herschel-Bulkley parameters like yield stress, consistency

and consistency index (Atzeni et al. 1985; Nehdi and Rahman 2004). There are also

contradictory observations reported in the literature for the same mixture of materials. For

instance, while the Bingham model reports negative yield stress for high dosage of

superplasticizers, Herschel-Bulkley model reports a non–negative yield stress and the

reason for this is normally ascribed to the so called non-linearity between shear stress and

shear rate of Herschel-Bulkley model (de Larrard et al. 1998).

As discussed earlier, the cement paste during the pre-induction period can exhibit shear

thinning when sheared in a coaxial viscometer. The rate at which the shear thinning

characteristic of the material changes depends on the time at which the material is

sheared. After pre-induction period, the same material can exhibit shear thickening

behaviour. The scenario is more complex for the superplasticizer-cement paste mixture.

Issues such as compatibility/incompatibility and low and high dosage change the

rheological behaviour considerably (Mailvaganam 1999). A compatible system at a low

dosage of superplasticizer normally will keep the cement particles in a state of dispersion

even up to 60 minutes (Agullo et al. 1999). This can be observed by the more or less

220
Newtonian behaviour in steady shear experiment, in a rotational viscometer. However if

the same material is sheared in a dynamic shear rheometer it is possible for it to exhibit

high normal stress differences thus conclusively proving the non–Newtonian response

characteristics (Mansoutre et al.1999; Min et al. 1994). Also, the same material can

exhibit creep and recovery and stress relaxation behaviour thus showing a viscoelastic

material response (Struble and Schultz 1993; Sun et al., 2006).

Significant insight into the nature of cement paste has been gained recently due to the

investigations of Van Damme et al. (2002). Essentially one can visualize the paste as a

dense assembly of deformable or rigid objects close to a jamming situation. These dense

assemblies are lubricated by the hydrodynamic or physical-chemical interactions of the

interstitial fluid, a mixture of water and superplasticizer. One important aspect of pastes

when viewed as an assemblage of granular particles is the property of dilatancy.

Essentially, pastes when subjected to shear strain will exhibit the tendency to expand in

the direction perpendicular to the shear plane. In a typical dynamic shear rheometer

experiment where the gap between the two plates is kept fixed, this essentially results in

the development of normal stresses. Van Damme et al. (2002) postulate that the two

important properties that give complete description of the mechanical behaviour of the

paste is the effect of dilatancy and inter-granular friction; the latter described very closely

by the ‘yield stress’. They used these concepts to study the rheological behaviour of

flocculated calcium silicate pastes and emphatically observe the need to measure normal

force in shear flow in addition to the ‘yield stress’ for a complete characterization of

superplasticized cement paste.

221
A proper constitutive model taking into account all the internal structural changes taking

place during the hydration process is lacking currently. This is not surprising as the

cement paste-superplasticizer mixture exhibits diverse and complex response

characteristics. However, even before one attempts in building a constitutive model, it is

imperative to understand through careful experimental investigations the transition in the

viscometric properties over a period of time. This information is necessary in building a

proper constitutive model. The intent of this investigation is precisely that.

It has been observed that the Marsh cone is extensively used in the optimization of high

performance concrete mixture design (see Chapter 3). Since the idea of using

superplasticizer is to impart linearly viscous behavior for a short period of time to the

paste and the mixture, the flow time measured in a Marsh cone has been linked to the

superplasticizer dosage beyond which there is no appreciable change in flow time is

normally taken as the optimum dosage. There are several interesting issues related to this

aspect. First, for a given cement paste – superplasticizer mixture, the optimum dosage

essentially relates the transition from a non-Newtonian to a Newtonian behavior. If one

assumes the Bingham model for the cement paste – superplasticizer mixture for dosages

less than the optimum dosages, the equations to be solved are the Buckingham-Reiner

equation and for the dosage equal to and greater than the optimum dosages, the classical

Hagen-Poiseuille equation are required to be solved. Also, discussions in the earlier

paragraphs clearly indicated that it is possible for the cement paste – superplasticizer

mixture to exhibit shear thickening, a typical non-Newtonian behavior for over dosage.

Hence, it is expected that the flow time might start increasing beyond the optimum

dosage as seen in the Chapter 3. It is also to be noted that this optimum dosage may not

be a sharp transition point, it is normally asymptotic in nature and hence it will be natural

222
that there can be several dosages near the optimum dosage which will result in the

transition from non-Newtonian to Newtonian. It will also be helpful if this transition

point is arrived at by appealing to kinematics of fluid flow rather than inferring it through

model dependent parameters.

In this investigation, several combinations of cement paste and superplasticizer are tested

in a Brookfield viscometer (details are given in Chapter 3) and Anton Paar dynamic shear

rheometer. The raw experimental data in terms of torque - shear stress – shear rate are

examined and detailed analysis is made on the various interesting manifestations of the

rheology of cement paste. The main thrust here is that the non-Newtonian behavior of

cement paste and superplasticizer mixture is characterized based on more fundamental

experimental investigations without recourse to any specific model. Steady shear

experiments, creep and recovery and stress relaxation experiments were conducted on a

variety of superplasticizer and cement paste combinations and the transition from non-

linear viscoelastic to non-Newtonian to linearly viscous fluid as the dosage rate is

increased is captured. Additionally the saturation dosages obtained from Marsh cone tests

were compared and it was found that dosage rates determined using Marsh cone were

closer to the dosage rates where the transition to the linearly viscous fluid was seen.

C.2 Materials used

The physical and chemical characteristics of cement and four types of superplasticizers

used for the study are same as given in Section 3.2

223
Table C.1–Mix details for steady shear experiments

Type of Dosage of Saturation dosage of


superplasticizer superplasticizer superplasticizer from
(sp/c %) Marsh cone test
(sp/c %)
SNF-S1 0.1, 0.2, 0.3, 0.4, 1 0.20
SNF-D2 0.1, 0.2, 0.4, 0.6, 1 0.23
SMF-S1 0.1, 0.2, 0.3, 1 0.23
PCE-D1 0.05, 0.1, 0.2, 0.4, 1 0.07

C.3 Experimental Procedure

The mixing procedure for the preparation of the paste is same as given in Section 3.5.The

Marsh cone test was used to evaluate the relative fluidity and the saturation

superplasticizer dosage. Table 1 gives The saturation superplasticizer dosages for the four

products used in this study are given in Table C.1. The details of the steady shear

experiments using Brookfield HA DV II +pro were explained in detail in Section 3.5. The

rheological experiments were also conducted with an Anton Paar dynamic shear

rheometer for studying the creep and recovery, and stress relaxation responses. The tests

were conducted with a parallel plate attachment with a gap of 1 mm (see Figure C.1). The

details of these tests are given in the respective sections that follow.

Fig. C.1(a) Anton Paar Rheometer (b) Shearing of cement paste with parallel plate
attachment

224
C.4 Results and Discussions

C.4.1 Marsh cone test on pastes

The saturation dosages obtained from Marsh cone test results for four superplasticized

pastes, already presented in Chapter 3, are summarized in Table C.1.

C.4.2 Steady shear tests with viscometer

Superplasticized cement paste generally shows shear thinning behaviour with an increase

in shear rate. This is due to the breakdown of the hydrated products formed around the

cement grain due to continuous mixing leading to the reduction in the apparent viscosity.

During unloading apparent viscosity remains constant during the down curve of the

hysteresis cycles for the PCE based product as shown in Figure C.2. It is observed that

the area within the loop decreases from the first cycle to the second cycle, because the

energy required for breaking down the hydration products decreases with time.

7 PCE-D1 0.1%

6
Shear stress (Pa)

3
First cycle
2 Second cycle
Area of first cycle = 86.49 J
1 Area of second cycle = 48.36 J

0
20 40 60 80 100 120 140 160 180
Shear rate (1/sec)

Fig. C.2- Loading-unloading cycles for the superplasticizer PCE-D1


at the dosage of 0.1%

Viscometric test results on paste shows that the behaviour changes from shear thinning

nature to Newtonian and shear thickening depending on the type and dosage of

superplasticizer (Lootens et al. 2004; Papo and Piani 2004; Chen et al. 2006). The
225
reported results on shear thickening are mainly based on oscillatory tests in terms of

storage modulus and viscous modulus (Chen et al., 2006; Chougnet et al. 2007). The

model parameters derived from viscometric tests and hysteresis cycles are used to capture

the change from shear thinning to shear thickening behaviour in Chapter 3.

The shear thinning and thickening nature of materials can also be classified based on the

area within the loading-unloading cycle as it reflects the energy required for breaking

down of the structure of the paste. Figures C.2 and C.3 show the work done in

deflocculating the particles for shear thinning and shear thickening compositions,

respectively. The area in the loop decreases from the first cycle to the second cycle for

shear thinning material because the energy required for breaking down of the particles

decreases. In the case of shear thickening material, the area within the loop increases.

30
PCE-D1 1%
25
Shear stress (Pa)

20
First cycle
15 Second cycle

Area of first cycle = 166.59 J


10 Area of second cycle = 188.21 J

5
20 40 60 80 100 120 140 160 180
Shear rate (1/sec)
Fig. C.3- Loading-unloading cycles for the superplasticizer PCE-D1 at the dosage of 1%

Table C.2 shows the areas within the loading-unloading loops for the different pastes

studied, along with the saturation dosages for different compositions and it can be seen

that more work is required at low dosages of superplasticizer to make the paste to flow,

and that the area of the loops decreases with an increase in the dosage of superplasticizer.

226
It can be observed that the area calculation is reliable upto the saturation dosage of

superplasticizer and the maximum percentage reduction occurs approximately at the

saturation dosage of superplasticizer. The paste with PCE-D1 shows shear thickening at

1% and hence work required for structural breakdown increases with time of shearing.

Thus, the reduction in the area of hysteresis cycles indicates the effectiveness of different

superplasticizers as it represents the amount of work required to deflocculate the

particles. From the perspective of rheology of cement paste-superplasticizer mixture, this

reduction in area is related to the transition from non-Newtonian to Newtonian regime.

However, the trend is not clear in other superplasticizers.

Table C.2 Work done in different loading-unloading cycles

Type of sp/c Saturation Loop area of Loop area of


SP % dosage first cycle second cycle
(sp/c%) (N/m2s) (N/m2s)
0.1 234 230
SNF-S1 0.2 108 92
0.3 0.20 28 3
0.4 85 65
1 44 7
0.1 667 356
SNF-D2 0.2 115 55
0.4 0.23 143 135
0.6 140 135
1 124 112
0.1 672 299
SMF-S1 0.2 0.23 81 41
0.3 118 58
1 130 85
0.05 1813 820
0.1 86 48
PCE-D1 0.2 0.07 91 22
0.4 37 24
1 167 188

227
C.4.2.1 Shear thinning index

Superplasticized paste exhibits shear thinning nature as the structural breakdown is faster

than structural build up. The degree of shear thinning of non Newtonian fluids for

different combinations can be represented by the shear thinning index (ASTM D 2196

(2005)); the shear thinning index is defined as the apparent viscosity at low speed to that

at high speed. Accordingly, in this case the shear thinning index is calculated based on

the apparent viscosity at 25 rpm and 175 rpm.

Table C.3 shows the influence of the type and dosage of superplasticizer on the shear

thinning index. Higher the shear thinning index, higher is the degree of shear thinning. It

is observed that generally the shear thinning index decreases with an increase in the

dosage of superplasticizer, which shows that the effect is predominant at low dosages.

PCE D1 shows considerable shear thinning even at low dosages and at overdosage,

causes some shear thickening (i.e., shear thinning index is less than 1). For all

compositions, shear thinning occurs upto the saturation dosage and after that the index

remains almost constant. The shear thinning index of less than one in all cases shows that

the apparent viscosity at high shear rate is slightly higher than that at low shear rate. This

generally occurs at overdosage of superplasticizer because of the excess amount of

superplasticizer available in the solution which forms a network. Hence, it is proposed

that the shear thinning index of one can be used as a limit for the dosage of the

superplasticizer. Interestingly, the dosage corresponding to this limit is practically

comparable with the saturation dosage obtained with Marsh cone and coaxial cylinder

viscometer (except SNF-D2). In the case of PCE based superplasticizer, a small increase

in dosage of 0.05% causes a large reduction in the shear thinning index, showing its

effectiveness at low dosages.

228
Table C.3 Shear thinning indices for different types and dosages of superplasticizers

Type of Dosage of SP Shear thinning


superplasticizer index
SNF-S1 0.1 4
0.2 3
0.3 2
0.4 1
1 1
0.1 3
0.2 3
SNF-D2 0.4 1
0.6 1
1 1
0.1 3
0.2 3
SMF-S1 0.3 2
1 1
0.05 3
0.1 1
PCE-D1 0.2 1
0.4 1
1 1

C.4.3 Creep and Recovery and Stress Relaxation Experiments

The behaviour of cement pastes under steady shear has been determined using different

procedures (Lapasin et al. 1979; Lapasin et al. 1983; Atzeni et al. 1995; Geiker 2002).

These studies explain that the material behaviour changes from solid-like to fluid-like

with the variation of shear stress with time. They consider the yield stress as a criterion

for the change of material behaviour. The concept of ‘yield stress’ is discussed in detail in

the excellent review article by Barnes (1999) which also discusses various issues related

to whether ‘yield stress’ is a model derived parameter or a true phenomenon. In the

present study, tests were conducted with an Anton Paar rheometer using five different

compositions - pure cement paste and cement paste with four types of superplasticizers to

229
understand the influence of different superplasticizers on the flow behaviour (see Table

C.4). The details of the testing methods are given in Table C.5.

Table C.4 Mix details for creep and recovery and stress relaxation experiments

Composition of paste w/c Dosage of superplasticizer


(sp/c%)
Pure cement paste 0.35 Nil
Cement paste with SNF-S1 0.35 0.05, 0.20
Cement paste with SNF-D2 0.35 0.05, 0.23
Cement paste with SMF-S1 0.35 0.05, 0.23
Cement paste with PCE-D1 0.35 0.01, 0.025, 0.05, 0.07, 0.1

Table C.5 Details of the different data acquisition frequencies

Details of test Number of Measuring point Testing details


data points duration
1. Creep and Recovery Measure Strain (Є) vs t
Creep (10 sec) 200 0.05 sec
Recovery (100 sec) 40 0.05 to 5 sec varied
linearly during 100 sec σ

10 sec t
100 sec

2. Stress relaxation Measure σ vs t


Ramping of strain 100 5×10-3 sec
(0.5 sec)
Stress relaxation 10 5×10-3 sec to 20 sec Є
(33 sec) varied logarithmically
during 33 sec
0.5 sec t

C.4.3.1 Creep and Recovery Test on Cement Paste

Most flow studies of paste and models used consider the material as viscous only. In this

section an attempt is made to observe whether the material shows any viscoelastic nature

and whether the addition of superplasticizer results in a transition to non-Newtonian

behaviour.
230
Struble and Leit (1995), and Struble and Schultz (1993) demonstrated that creep and

recovery behaviour of cement paste undergoes a transition from solid-like behaviour at

low stress to fluid like behaviour at high stress. However, at high stress, the material

behaves as a viscous liquid with a linear increase in strain throughout the duration of

stress with no recovery. This stress level associated such a transition agrees with the yield

stress estimated from flow curves and oscillatory tests in their study. Creep and recovery

tests have also been used to monitor the setting of cement paste based on the development

of yield stress with time (Struble and Leit 1995).

Generally, in a stress controlled test, when the material is subjected to instantaneous

stress followed by a constant stress, an instantaneous increase in strain occurs initially

and then there is a continuous increase in strain at a non-constant rate, i.e., an elastic

response followed by combination of elastic and viscous response. If the material is fluid-

like, the strain rate asymptotically approaches a constant value. While releasing the stress,

some instantaneous strain recovery followed by delayed recovery occurs (Wineman and

Rajagopal, 2000).

In the present study, creep and recovery tests were done on pure cement paste and

superplasticized cement pastes. Tests were conducted at different stress levels in order to

understand how the nature of the material changes at different stress levels for different

types of superplasticizers and dosage rates. Figure C.4 shows the test results for pure

cement paste, which shows a typical viscoelastic nature with some residual strain at

different stress levels. The stress levels used include 50 Pa, 75 Pa and 100 Pa and it can

be observed that the strain produced at higher stress levels is higher. Also, the recovery is

less for lower stress.

231
0.007
Cement paste with w/c 0.35
0.006
50 Pa
75 Pa
0.005 100 Pa

0.004
Strain 0.003

0.002

0.001

0.000
20 40 60 80 100 120
Time (sec)

Fig. C.4 Creep and recovery test for pure cement paste

Influence of type of superplasticizer on creep and recovery

Figure C.5 (a-d) shows the influence of types of superplasticizers on the rheology of

cement paste. Tests were conducted for four superplasticizers (see Table C.4) at constant

dosage of 0.05%. the pastes with the SNF and SMF based superplasticizers show the

same trend as the pure cement paste; however, strain recovery is less compared to pure

cement paste. Moreover, the cement paste with PCE based superplasticizer shows an

entirely different nature. It shows linear strain initially and no instantaneous strain or

recovery. This is because the paste becomes viscous at that dosage due to the better

dispersion produced by the combined action of electrostatic repulsion and steric

hindrance.

232
0.8 SNF-D2 - 0.05%
0.35 SNF-S1 - 0.05%
50 Pa
0.30 75 Pa 0.7
100 Pa
0.6 50 Pa
0.25
0.5 75 Pa

Strain
Strain

0.20 0.4
100 Pa

0.15 0.3
0.10 0.2
0.05 0.1
0.00 0.0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(a) Creep and recovery of SNF-S1 (b) Creep and recovery of SNF-D2
600
SMF-S1 - 0.05% 50 Pa PCE-D1 - 0.05%
0.4 75 Pa 500
100 Pa 50 Pa
0.3 400 75 Pa
100 Pa
Strain
Strain

300
0.2
200
0.1
100

0.0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(c) Creep and recovery of SMF-S1 (d) Creep and recovery of PCE-D1

Fig. C.5 (a-d) Creep and recovery test result of cement paste and with
constant dosage of different superplasticizers

Influence of dosage of superplasticizer on creep and recovery

In order to understand the influence of dosage of superplasticizer on creep and recovery,

tests were conducted with the most effective superplasticizer PCE-D1 (determined from

Marsh cone test) at different dosages. The test results are given in Figure C.6 (a-c). It is

observed that the nature of the paste changes from viscoelastic to viscous as the dosage of

superplasticizer increases from low to saturation. Also, the strain/deformation of the

material increases with an increase in dosage of superplasticizer.

233
0.08 500
PCE-D1 - 0.025%
PCE-D1 - 0.01% 50 Pa
0.07 75 Pa
100 Pa
400
0.06
50 Pa
0.05 300

Strain
75 Pa
Strain

0.04 100 Pa

0.03 200
0.02
100
0.01
0.00 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(a) Creep and recovery of PCE-D1, 0.01% (b) Creep and recovery of PCE-D1, 0.025%
600 600
PCE-D1 - 0.05% PCE-D1 - 0.1%

500 500

400 50 Pa 400 50 Pa
75 Pa 75 Pa
Strain

Strain

300 100 Pa 300 100 Pa

200 200

100 100

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(c)Creep and recovery of PCE-D1, 0.05% (d) Creep and recovery of PCE-D1, 0.1%

Fig. C.6 (a-d) Variation in creep and recovery behaviour


with different dosages of PCE-D1

Influence of saturation dosage of superplasticizer on creep and recovery

A comparison of the flow behaviour of pastes was obtained by conducting creep tests at

the saturation dosage of superplasticizers (see Figure C.7 (a-d)). It is observed that all

pastes show a viscous nature with a linear increase in strain initially and no recovery

while releasing the stress at their saturation dosages. Thus a transition in the nature of

material occurs at saturation dosages, which is necessary to have a flowable concrete at

the saturation dosages.

234
SNF-S1 - 0.20% SNF-D2 - 0.23%
600 1200
500 1000
400 800

Strain
Strain

300 600
200 400
50 Pa 50 Pa
100 75 Pa 200 75 Pa
100 Pa 100 Pa
0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(a) SNF-S1, at saturation dosage, (b) SNF-D2, at saturation dosage ,

SMF-S1- 0.23% 1800 PCE-D1 0.07%


500 50 Pa
75 Pa
1500
400 100 Pa
1200
Strain
Strain

300
900
200 600
50 Pa
100 75 Pa
300 100 Pa

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

(c) SMF-S1, at saturation dosage , (b) PCE-D1, at saturation dosage

Fig. C.7 (a-d) Creep test result at saturation dosage of superplasticizers

Check for linearity of response

The simplest modeling concept for viscoelastic materials for which the experimental data

can be characterized is called linearity of response (Wineman and Rajagopal, 2000). It

consists of two conditions (1) linear scaling and (2) superposition of separate responses.

If the response of a viscoelastic material displays both linear scaling and possibility

superposition of separate responses for all histories, the response is said to be linear. If

the stress (σ0) is changed by a factor α, and if the deformation is also changed

235
consequently by the same factor α, then the material shows linear scaling. In our current

investigation, the stress levels are 50 Pa, 75Pa and 100Pa and for checking linear

response, 50 Pa is multiplied by a factor of 1.5 and 2. If the material satisfies linear

scaling, the deformation response obtained by increasing the stress by these factors

should be same as that of the deformation produced by 75 Pa and 100 Pa respectively.

However, Figure C 8 (a and b) shows that the pure cement paste as well as that with

superplasticizer at a low dosage, violate the condition of linear scaling in the tested

compositions. The same behavior was found for all the combinations tested.

0.007
Cement paste with w/c 0.35 SMF-S1 - 0.05% Actual
0.4
0.006 Observed
Actual σ= 100 Pa Expected
0.005 Observed 0.3
Expected
Strain

0.004
Strain

0.003 0.2
σ=100Pa
0.002
0.1 σ=2*50 Pa
σ=2*50 Pa
0.001 σ=50 Pa
σ=50 Pa
0.000 0.0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
Time (sec) Time (sec)

Fig. C.8 Test for linear scaling

If the material follows the principle of superposition, then the instantaneous deformation

due to the applied load at time t = 0 and the instantaneous deformation recovered on the

release of the load should be same. The instantaneous deformation due to loading and

instantaneous deformation due to unloading as observed from above Figure C.8 and

Table C.6 at a dosage of 0.05%, are different. This shows that the cement paste tested in

this investigation also violates the condition of superposition of separate responses. These

two checks show that pure cement paste and superplasticized pastes have a non-linear

236
viscoelastic nature at dosages less than saturation. However, the recovery could not be

observed for superplasticizers at saturation dosages.

Table C.6 Superposition of separate responses

Composition Stress (Pa) Instantaneous Recovered


Strain on strain on
loading unloading
50 0.00037 0.000171
Cement paste 75 0.00038 0.00017
100 0.0026 0.00114
Cement paste 50 0.0107 0.0015
with SNF-S1 75 0.0119 0.003
100 0.0147 0.0045
Cement paste 50 0.0098 0.002
with SNF-D2 75 0.0141 0.003
100 0.0214 0.0035
Cement paste 50 0.0072 0.0014
with SMF-S1 75 0.0141 0.003
100 0.0171 0.0045
Cement paste 50 0.0127 0
with PCE-D1 75 0.0194 0
100 0.0254 0

C.4.3.2 Stress Relaxation

In the case of a viscoelastic material, when strain is instantaneously increased at t = 0 and

then maintained, the stress instantaneously increases and then gradually decreases with

time. Tests are conducted on different compositions as described in creep and recovery

tests. In the case of pure cement paste, the material shows viscoelastic nature as the stress

reduces asymptotically to zero (see Figure C.9).

237
1000 Cement paste with w/c 0.35

Shear stress (Pa)


Strain 0.25%
800 Strain 0.5%
Strain 0.75%
600

400

200

0
0 10 20 30
Time (sec)

Fig. C.9 Stress relaxation of pure cement paste

Influence of type of superplasticizer on stress relaxation

Figure C.10 (a-d) shows stress relaxation test results for different types of

superplasticizers at a constant dosage of 0.05%.The SNF and SMF based

superplasticizers show viscoelastic nature as in the case of pure cement paste. In the case

of PCE based paste, no external force is required to maintain the strain because the

dosage is high enough to ensure fluid-like behaviour. This is because of the better

dispersion produced due to PCE based superplasticizer.

35
SNF-S1- 0.05% SNF-D2 - 0.05%
30 0.25% 120 0.25%
Shear stress (Pa)

0.5% 0.5%
100
Shear stress (Pa)

25 0.75% 0.75%

20 80

15 60

10 40

5 20

0 0
0 5 10 15 20 25 30 35
0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(a) Stress relaxation of SNF-S1 (b) Stress relaxation of SNF-D2

238
10
50 SMF-S1 - 0.05% PCE-D1 - 0.05%
0.25%
Shear Stress (Pa) 0.25%

Shear stress (Pa)


0.5%
5 0.5%
40 0.75%
0.75%

30
0
20
-5
10

0 -10
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(c) Stress relaxation of SMF-S1 (b) Stress relaxation of PCE-D1

Fig. C.10 (a-d) Stress relaxation test results for pastes with 0.05%
dosage for different superplasticizers

Influence of dosage of superplasticizer on stress relaxation

Tests were performed with different dosages of PCE-D1. It is seen that the shear stress

decrease with increase in dosage of PCE based superplasticizer, as in Figure C.11 (a-e).

Even though the paste shows stress relaxation at low dosages of 0.01% and 0.025%, the

behaviour is viscous at higher dosages.

120 PCE-D1 - 0.025%


PCE-D1 - 0.01% 50
Shear stress (Pa)

0.25%
Shear Stress (Pa)

100 0.25%
0.5% 0.5%
0.75% 40 0.75%
80
30
60

40 20

20 10

0 0
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(a) Stress relaxation of PCE-D1, 0.01% (b) Stress relaxation of PCE-D1, 0.025%

239
10 10
PCE-D1 - 0.05% PCE-D1 - 0.07%
Shear stress (Pa) 0.25% 0.25%

Shear stress (Pa)


5 0.5% 5 0.5%
0.75% 0.75%

0 0

-5 -5

-10 -10
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(c) Stress relaxation of PCE-D1, 0.025% (d) Stress relaxation of PCE-D1, 0.07%

10
PCE-D1 - 0.1%
Shear stress (Pa)

0.25%
5 0.5%
0.75%

-5

-10
0 5 10 15 20 25 30 35
Time (sec)

(e) Stress relaxation of PCE-D1, 0.1%

Fig. C.11 (a-e) Stress relaxation at different dosages of PCE-D1

Influence of saturation dosage of superplasticizer on stress relaxation

It is clear that the dosage of the superplasticizer affects the relaxation. From the tests on

pastes with PCE-D1, it appears that at dosages much lower than the saturation dosages,

there is appreciable relaxation with time whereas the behaviour is viscous at dosages

higher than saturation dosages. At the saturation dosage, the paste with PCE-D1 has a

practically viscous response. The stress relaxation shown by SNF-S1 and SMF-S1 at

saturation dosage is appreciable but small whereas there is zero stress relaxation for the

240
paste with SNF-D2 as in the case of PCE-D1 as shown in Figure C.12 (a-d). This results

further confirms the viscous nature of superplasticized pastes at and beyond saturation

dosages.

25 10
SNF-S1 - 0.20% SNF-D2 - 0.23%
20 0.25%

Shear stress (Pa)


Shear stress (Pa)

0.25%
5 0.5%
15 0.5%
0.75% 0.75%
10
5 0
0
-5 -5
-10
-15 -10
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(a) SNF-S1, at saturation dosage (b) SNF-D2, at saturation dosage

25 10
SMF-S1 - 0.23% PCE-D1 - 0.07%
Shear stress (Pa)

0.25% 0.25%
Shear stress (Pa)

20 0.5% 0.5%
0.75% 0.75%
15
0
10

0 -10
0 5 10 15 20 25 30 35 0 5 10 15 20 25 30 35
Time (sec) Time (sec)

(c) SMF-S1, at saturation dosage (d) PCE-D1, at saturation dosage

Fig. C.12 (a-d) Stress relaxation of pastes at saturation dosage of superplasticizers

C.5 Comparison of rheological tests

The creep and recovery and stress relaxation tests shows that a transition from non-

linear viscoelastic to viscous nature occurs for pastes as the superplasticizer dosage

changes from low to saturation and beyond. The saturation dosages obtained from

Marsh cone tests are shown to yield viscous behaviour in all cases studied.

241
C.6. SUMMARY

In this investigation, steady shear, creep and recovery, and stress relaxation experiments

were conducted on cement pastes with different superplasticizers of varying dosage rates.

This helps develop an understanding of the transition of these mixtures from viscoelastic

fluid-like materials to non-Newtonian fluids to linearly viscous fluids, as the

superplasticizer dosage rate is increased. The work also increases the understanding of

shear thinning characteristics as a function of superplasticizer type and dosage rates. The

area of hysteresis cycles and shear thinning index were used for characterizing the shear

thinning nature. As superplasticizer dosage increases, in general, there is a transition in

the behaviour from viscoelastic to linearly viscous fluid behaviour. In the creep and

recovery experiment, this is reflected as a linear change of strain with zero recovery

during unloading. In the case of stress relaxation, the classical stress spike with zero

stress after application of constant strain was seen. The dosages at which the change

occurs compares well with the saturation dosages determined using the Marsh cone.

242
REFERENCES

ACI 116 R (2000) Cement and concrete terminology. ACI Publicaton, USA.

ACI E4 (2003) Eduction Bulletin, Chemical admixtures for concrete. ACI


Publication, USA.

Agarwal, S.K., I. Masood, and S.K. Malhotra (2000) Compatibility of


superplasticizers with different cements. Construction and Building Materials, 14,
253-259.

Agullo, L., B.T. Carbonari, R. Gettu, and A. Aguado (1999) Fluidity of cement
pastes with mineral admixtures and superplasticizer - a study based on the Marsh cone
test. Materials and Structures, 32, 479-485.

Aїtcin, P.C., S.L. Sarka, M. Regourd, and D. Volant (1987) Retardation effect of
superplasticizers on different cement fractions. Cement and Concrete Research, 17,
995-999.

Aїtcin, P.C., C. Jolicoeur, and J.G. Macgregor (1994) Superplasticizers: How they
work and why they occasionally don’t. Concrete International, 16, 45-52.

Aïtcin, P.C. High performance concrete. E& FN Spon, London, 1998.

Aїtcin, P.C., S. Jiang, B.G. Kim, P.C. Nkinamubanzi, and N. Petrov (2001)
Cement/ superplasticizer interaction: The case of polysulphonates. Bulletin Des
Laboratories Ponts et Chausses, July-August, 89-99.

Akman, M.S and Z. Cavdar, Efflorescence problem in superplasticized concretes


pp. 316-329. J.G. Cabrera, and R.R. Villarreal (eds) Proceedings of International
Symposium on the Role of Admixtures in High Performance Concrete, RILEM,
Mexico, 1999.

Andersen, P.J., D.M. Roy, and J.M. Gaidis (1987) The effects of adsorption of
superplasticizers on the surface of cement. Cement and Concrete Research, 17, 805-
813.

Asaga, K. and D.M. Roy (1980) Rheological properties of cement mixes: 1V. Effects
of superplasticizers on viscosity and yield stress. Cement and Concrete Research, 10,
287-295.

Assaad, J., K.H. Khayat, and H. Mesbah (2003) Assessment of thixotropy of


flowable and self-Consolidating Concrete. ACI Materials Journal, 100, 99-107.

ASTM C 29/29M (2007) Standard test method for bulk density and voids in
aggregate. Annual book of ASTM standards, American Society of Testing and
Materials, USA.

ASTM C 109/109M (1998) Standard test method for compressive strength of


hydraulic cement mortars. Annual book of ASTM standards, American Society of
Testing and Materials, USA.
ASTM C 150 (2007) Standard specification for Portland cement. Annual book of
ASTM standards, American Society of Testing and Materials, USA.

ASTM C 187 (2004) Standard test method for normal consistency of hydraulic
cement. Annual book of ASTM standards, American Society of Testing and Materials,
USA.

ASTM C 191 (2001a) Standard test method for time of setting of hydraulic cement by
Vicat needle. Annual book of ASTM standards, American Society of Testing and
Materials, USA.

ASTM C 204 (2007) Standard test method for fineness of hydraulic cement by air
permeability apparatus. Annual book of ASTM standards, American Society of
Testing and Materials, USA.

ASTM C 305 (1989) Standard practice for mechanical mixing of hydraulic cement
pastes and mortars of plastic consistency. American Society for Testing and Materials,
USA.

ASTM C 494/C 494M (2008a) Standard specification for chemical admixtures for
concrete. Annual book of ASTM standards, American Society for Testing and
Materials, USA.

ASTM C 939 (1987) Standard test method for flow of grout for preplaced-aggregate
concrete (flow cone method), Annual book of ASTM standards, American Society for
Testing and Materials, USA.

ASTM C 1621/C 1621M (2006) Standard test method for passing ability of self
consolidating concrete by J-ring Annual book of ASTM standards, American Society
for Testing and Materials, USA.

ASTM D 2196 (2005) Standard test methods for rheological properties of non-
Newtonian materials by rotational (Brookfield Type) viscometer, American Society
for Testing and Materials, USA.

ASTM D 4643 (1987) Standard test method for determination of water (moisture)
content of soil by the microwave oven method, Annual book of ASTM standards,
American Society for Testing and Materials, USA.

Atzeni, C., L. Massida, and U. Sanna (1985) Comparison between rheological


models for portland cement pastes. Cement and Concrete Research, 15, 511-519.

Banfill, P.F.G. (1981) A viscometric study of cement pastes containing


superplasticizers with a note on experimental techniques. Magazine of Concrete
Research, 33, 37-47.

Banfill, P.F.G. and D.C. Saunders (1981) On the viscometric examination of cement
pastes. Cement and Concrete Research, 11, 363-370.

Banfill, P.F.G. (2003) The rheology of fresh cement and concrete - A review.
Proceedings of 11th International congress on Cement Chemistry, Durban, South
Africa, 50-62.

244
Barnes, H.A. (1999) The Yield stress – a review, Journal of Non-Newtonian Fluid
Mechanics 81, 133-178.

Basile, F., S. Biagini, G. Ferrari, and M. Collepardi, Influence of different


sulphonated naphthalene polymers on the fluidity of cement paste. pp. 209-220, V.M
Malhotra (ed) Third International Conference on Superplasticizers and Other
Chemical Admixtures in Concrete, American Concrete Institute, Detroit, USA, 1989.

Basu, P.C. and S. Choudhury (2005) Influence of minor constituents of portland


cement on rheology of mortar for self compacting concrete. The Structural
Engineering Convention (SEC 2005), Indian Institute of Science, Bangalore, India,
December, 209-214.

Basu, P.C. and S. Choudhury (2007) Impact of fine aggregate particle size on
mortar rheology for SCC. The Indian concrete Journal, January, 7-14.

Baum, H. and M. Ben-Bassat Effect of superplasticizers on some properties of fresh


and hardened concrete mixes. pp. 482-492. J.G. Cabrera, and R.R. Villarreal (eds)
Proceedings of International Symposium on the Role of Admixtures in High
Performance Concrete, RILEM, Mexico, 1999.

Beaupre, D. and S. Mindess, Rheology of fresh concrete: Principles, measurement


and applications. pp. 149-190. J. Skalny and S. Mindess (eds) Materials Science of
Concrete V, The American Ceramic Society Publication, USA, 1998.

Bedard, C.P.E. and N.P. Mailvaganam (2005) The use of chemical admixtures in
concrete: Part1: Admixture –Cement compatibility. ASCE Journal of Performance of
Constructed Facilities, 19, 263-266.

Bedard, C.P.E. and N.P. Mailvaganam (2006) The use of chemical admixtures in
concrete: Part 11: Admixture-admixture compatibility and practical problems. ASCE
Journal of Performance of Constructed Facilities, 20, 2-5.

Bentur, A. (2002) Cementitious materials-Nine millennia and a new century: Past,


present and future. Journal of Materials in Civil Engineering, January/February, 2-22.

Bhatty, J.I. and P.F.G. Banfill (1982) Sedimentation behaviour in cement pastes
subjected to continuous shear in rotational viscometers. Cement and Concrete
Research, 12, 69-78.

Billberg, P., The effect of mineral and chemical admixtures on fine mortar rheology.
pp. 301-320. V.M. Malhotra (ed) Proceedings of Fifth CANMET/ACI International
Conference on superplasticizers and other chemical admixtures in concrete,
American Concrete Institute, USA, 1997.

Bonen, D. and S.L. Sarkar (1995) The superplasticizers adsorption capacity of


cement pastes, pore solution composition, and parameters affecting flow loss. Cement
and Concrete Research, 25, 1423-1434.

Brooks, J.J., M.A.M. Johari, and M. Mazloom (2000) Effect of admixtures on


the setting times of high-strength concrete. Cement and Concrete Composites, 22,
293-301

245
Calleja, J. (1952a) Determination of setting and hardening time of high-alumina
cements by electrical resistance techniques. Journal of American Concrete Institute, 4,
329-330.

Calleja, J. (1952b) New techniques in the study of setting and hardening of hydraulic
materials. Journal of American Concrete Institute, 23, 525-536.

Chang, P.K. and Y.N. Peng (2001) Influence of mixing techniques on properties of
high performance concrete. Cement and Concrete Research, 31, 87-95.

Chandra, S. and J. Bjornstrom (2002a) Influence of cement and superplasticizers


type and dosage on the fluidity of cement mortars- Part 1. Cement and Concrete
Research, 32, 1605-1611.

Chandra, S. and J. Bjornstrom (2002b) Influence of superplasticizers type and


dosage on the slump loss of Portland cement mortars- Part 11. Cement and Concrete
Research, 32, 1613-1619.

Chen, C.T., L.J. Struble, and H. Zhang (2006) Using dynamic rheology to measure
cement-admixture interactions. Journal of ASTM International, 3, 1-13.

Chougnet A., A. Audibert, and M. Moan (2007) Linear and non-linear rheological
behaviour of cement and silica suspensions - Effect of polymer addition. Rheologica
Acta, 46, 793-802.

Claisse, P.A., P. Lorimer, and M.H. Al-Omari, The effect of changes in cement on
the properties of cement grouts with superplasticizing admixtures, pp. 19-33. J.G.
Cabrera, and R.R. Villarreal (eds) Proceedings of International Symposium on the
Role of Admixtures in High Performance Concrete, Mexico, 1999.

Collepardi, M. (1998) Admixtures used to enhance placing characteristics of


concrete. Cement and Concrete Composites, 20, 103-112.

Collepardi, M., Admixtures enhancing concrete performance. pp. 217-230. R.K.


Dhir, P.C. Hewlett and M.D Newlands (eds) Proceedings of International
Conference on Admixtures-Enhancing Concrete Performance, Thomas Telford,
Dundee, UK, 2005.

Corinaldesi,V. and G. Moriconi, The influence of mineral additions on the rheology


of self compacting concrete. pp. 227-240. V.M Malhotra (ed) Seventh CANMET/
ACI International Conference on Superplasticizers and Other Chemical Admixtures in
Concrete, Berlin, Germany, 2003

Corradi, M., R. Khurana, and Magarotto, New superplasticizers for the total
control of performances of fresh and hardened concrete. pp. 1-10. R.K. Dhir, P.C.
Hewlett and M.D. Newlands (eds),Proceedings of International Conference on
Admixtures-Enhancing Concrete Performance, Thomas Telford, London, 2005.

Cyr, M., C. Legrand, and M. Mouret (2000) Study of the shear thickening effect of
superplasticizers on the rheological behaviour of cement pastes containing or not
mineral additives. Cement and Concrete Research, 30, 1477-1483.

246
Cyr, M. and M. Mouret, Rheological characterization of superplasticized cement
pastes containing mineral admixtures: Consequences on self compacting concrete
design pp. 241-256. V.M Malhotra (ed) Seventh CANMET/ ACI International
Conference on Superplasticizers and Other Chemical Admixtures in Concrete, Berlin,
Germany, 2003

Daimon, M. and D. M. Roy (1978) Rheological properties of mixes: I. methods,


preliminary experiments, and adsorption studies. Cement and Concrete Research, 8,
753-764.

Das, A.C. (2007) Admixture-cement Compatibility, M.Tech Project report, Indian


Institute of Technology, Madras, India.

de Larrard, F (1990) A method for proportioning high strength concrete mixtures.


Cement, Concrete and Aggregates, 12, 47-52.

de Larrard, F., C.F. Ferraris, and T. Sedran (1998) Fresh concrete: A Herschel-
Bulkley material. Materials and Structures, 31, 494-498.

de Larrard, F.(1999) Why rheology matters? Concrete International, 21, 79-81.

Dhir, R.K. and A.W.F. Yap (1983) Superplasticized high workability concrete:
some properties in the fresh and hardened states. Magazine of Concrete Research, 35,
215-228.

Diamond, S. (2005) The patch microstructure in concrete: effect of mixing time.


Cement and Concrete Research, 35, 1014-1016.

Diop, M.B., A.A. Ndiaya, S. Ndiaye, B. Lo, O. Sakho, P.D. Tall, A.C. Beye, and
W. Soboyejo, Evaluation of phosphogypsum as an alternative of gypsum in the
fabrication of CEM II/A-M 32.5 R Portland cement. R.K. Dhir, T.A. Harrison and
M.D. Newlands (eds) Proceedings of International conference on Cement
Combinations for Durable Concrete, University of Dundee, UK, 2005.

Dodson, V.H. and T.D. Hayden (1989) Another look at the Portland cement
/chemical admixture incompatibility problem. Cement Concrete and Aggregates, 11,
52-56.

Domone, P. (2006) Mortar tests for self consolidating concrete- A contribution from
ACI committee 236, Concrete International, April, 39-45.

EFNARC (2005) The European Guidelines for Self Compacting Concrete:


Specification, Production and Use. The European Federation of Specialist
Construction Chemicals and Concrete Systems. www.efnarc.org visited on 25/7/2007.

Emoto, T. and T.A. Bier (2007) Rheological behaviour as influenced by plasticizers


and hydration kinetics. Cement and Concrete Research, 37: 647-654.

Erdogdu, S. (2000) Compatibility of superplastcizers with cements different in


composition. Cement and Concrete Research, 30, 767-773.

Falikman, V.R. Y.V Sorokin, A.Y.Vainer, and N.F Bashlykov, New high
performance polycarboxylate superplasticizers based on derivative copolymers of
247
maleinic acid. pp. 41-46. R.K. Dhir, P.C. Hewlett and M.D Newlands (eds)
Proceedings of International Conference on Admixtures-Enhancing Concrete
Performance. Thomas Telford, Dundee, UK, 2005.

Famy, C., K.L. Scrivener, A. Atkinson, and A.R. Brough (2002) Effects of an early
or late heat treatment on the microstructure and composition of inner C-S-H products
of portland cement mortars. Cement and Concrete Research, 32, 269-278.

Faroug, F., J. Szwabowski, and S. Wild (1999) Influence of superplasticizers on


workability of concrete, Journal of materials in civil engineering, May, 151-157.

Felekoglu, B., B. Baradan, H. Sarikahya, and L. Yuceer, Compatibility of a new


generation polycarboxylate-type superplasticizer with different set accelerators. pp.
89-96. R.K. Dhir, P.C. Hewlett and M.D Newlands (eds) Proceedings of
International Conference on Admixtures-Enhancing Concrete Performance. Thomas
Telford, Dundee, UK, 2005.

Felekoglu, B. and H. Sarikahya (2007) Effect of chemical structure of


polycarboxylate based superplasticizers on workability retention of self compacting
concrete. Construction and Building Materials (Article in press).

Fernandez-Altable, V. and I. Casanova (2006) Influence of mixing sequence and


superplasticizer dosage on the rheological response of cement pastes at different
temperatures. Cement and Concrete Research, 36, 1222-1230.

Fernon, V., A. Vichot, N. Le Goanvic, P. Colombet, F. Corazza, and U. Costa,


Interaction between Portland cement hydrates and polynaphthalene sulfonates. pp.
225-248. V.M. Malhotra (ed) Proceedings of 5th Canmet/ACI International
Conference on Superplasticizers and Other Chemical Admixtures in Concrete,
American Concrete Institute, Detroit, 1997.

Ferraris, C.F. and J.M Gaidis (1992) Connection between the rheology of concrete
and rheology of cement paste. ACI Materials Journal, 88, 388- 393.

Ferraris, C.F. (1999) Measurement of rheological properties of HPC: State of art


report. Journal of research of the National Institute of Standards and Technology,
104, 461-478.

Ferraris, C.F. (2001) Concrete mixing methods and concrete mixers: State of the art
report. Journal of Research of the National Institute of Standards and Technology,
106, 391-399.

Ferraris, C.F., K.H. Obla, and R. Hill (2001) The influence of mineral admixtures
on the rheology of cement paste and concrete. Cement and Concrete Research, 31,
245-255.

Flatt, R.J., Y.F. Houst, P. Bowen, H. Hofmann, J. Widmer, U. Sulser, U. Maeder,


and T.A. Bürge, Interaction of superplasticizers with model powders in a highly
alkaline medium. pp. 743-762. V.M. Malhotra (ed) Proceedings of 5th
CANMET/ACI International Conference on Superplasticizers and Other Chemical
Admixtures in Concrete, American Concrete Institute, Detroit, 1997.

248
Flatt, R.J., Y.F. Houst, P. Bowen, H. Hofmann, J. Widmer, U. Sulser, U. Maeder,
and T.A. Bürge (1998) Analysis of superplasticizers used in concrete. Analusis
Magazine, 26, 28-34.

Flatt R.J. and Y.F. Houst (2001) A simplified view on chemical effects perturbing
the action of superplasticizers. Cement and Concrete Research, 31, 1169-1176.

Flatt, R. J. (2004) Dispersion forces in cement suspensions. Cement and Concrete


Research, 34, 399- 408.

Gaidis J.M. and E.M. Gartner, Hydration Mechanisms II. pp. 9-39. J. Skalny and
S. Mindess (eds), Materials Science of concrete II, The American Ceramic Society,
USA, 1991.

Geiker M.R., B. Mari, L. Nyholm, and N. Lauge (2002) On the effect of coarse
aggregate fraction and shape on the rheological properties of self compacting
concrete. Cement, Concrete and Aggregates, 24, 3-6.

Gettu, R., A. Aguado, L. Agullo, B. Toralles-Carbonari, and J. Roncero,


Characterization of cement pastes with silica fume and superplasticizer as components
of high performance concretes. pp. 331-344. P.K Mehta (ed) Mario Collepardi
Symposium on Advances in Concrete Science and Technology, 1997.

Gettu, R., J. Roncero, and M.A. Martin (2002) Long term behaviour of concrete
incorporating a shrinkage-reducing admixture. The Indian Concrete Journal,
February, 1-7.

Gettu, R., P.C.C. Gomes, L. Agullo, and A. Josa, High strength self compacting
concrete with fly ash: Development and utilization. pp. 507-522. V.M. Malhotra (ed)
Eighth CANMET/ACI International Conference on Flyash, Silica Fume, Slag and
Natural Pozzolans in Concrete, ACI International, USA, 2004.

Gettu, R., B. Barragan, T. Garcia, J. Ortiz, and R. Justa (2006) Fiber Concrete
tunnel lining. Concrete International, August, 63-69.

Giaccio, G. and R. Zerbino (2002) Optimum superplasticizer dosage for systems


with different cementitious materials. The Indian Concrete Journal, September, 553-
557.

Goisis M., A. Buscema, and T. De Marco, Characterization of the rheological


properties of cement paste prepared with polycarboxylate type superplasticizer. pp. 1-
15. V.M Malhotra (ed) Seventh CANMET/ ACI International Conference on
Superplasticizers and Other Chemical Admixtures in Concrete, Berlin, Germany,
2003.

Golaszewski J. and J. Szwabowski (2004), Influence of superplasticizers on


rheological behaviour of fresh cement mortars. Cement and Concrete Research, 34,
235-248.

Gomes, P.C.C., R. Gettu, L. Agullo, and C. Bernad, Experimental optimization of


high strength self compacting concrete. pp. 377-386. K. Ozawa, and M. Ouchi (eds)
Proceedings of Second International Symposium on SCC, Kochi, Japan, 2001.

249
Gomes, P.C.C. (2002) Optimization and characterization of high strength self
compacting concrete. Doctoral thesis, Universitat Politecnica De Catalunya,
Barcelona, Spain.

Gopinath, R. and R. Gettu, Influence of shrinkage reducing admixtures on early age


shrinkage. pp.1132-1142. P.J. Rao, V. Ramakrishnan, I. Patnaikuni, and V.S.
Parameswaran (eds) Proceedings of Second International Conference on Advances
in Concrete and Construction, Hyderabad, India, 2008.

Grabiec, A.M. and Z. Piasta (2004) Study on compatibility of cement –


superplasticizer assisted by multicriteria statistical optimization. Journal of Materials
Processing Technology, 152, 197-203.

Griesser, A. (2002) Cement-superplasticizer interactions at ambient temperatures,


Doctoral thesis, Swiss Federal Institute of Technology, Zurich.

Grzeszczyk, S., and M. Sudol, The influence of temperature on the fluidity of fresh
self compacting concrete. pp. 165-170. R.K. Dhir, P.C. Hewlett and M.D Newlands
(eds) Proceedings of the International Conference on Admixtures-Enhancing
Concrete Performance, Thomas Telford, Dundee, UK, 2005.

Hanna, E., M. Ostiguy, K. Khalifé, O. Stoica, B.-G. Kim, C. Bédard, M. Saric-


Coric, M. Baalbaki, S. Jiang, P.C. Nkinamubanzi, P.C. Aїtcin, and N. Petrov
(2000) The importance of superplasticizers in modern concrete technology.
Superplasticizers conference, Nice, 2-18.

Hanehara, S. and K. Yamada (1999) Interaction between cement and chemical


admixture from the point of cement hydration, absorption behaviour of admixture and
paste rheology. Cement and Concrete Research, 29, 1159-1165.

Hanehara, S. and K. Yamada (2008) Rheology and early age properties of


cement systems. Cement and Concrete Research, 21, 175-195.

Hewlett P.C. (ed) (2004) Lea’s chemistry of cement and concrete. Elsevier Ltd,
Burlington, UK.

Hidalgo, J., C.T. Chen, and L.J. Struble (2008) Correlation between paste and
concrete flow behaviour, ACI Materials Journal, 105, 281-288.

Houst, Y.F., R.J. Flatt, H. Hofmann, U. Mader, J. Widmer, U. Sulser, and T.A.
Burge, Influence of superplasticizer adsorption on the rheology of cement paste,
pp.71-78, J.G. Cabrera and R.R. Villarreal (eds) Proceedings of International
Symposium on the Role of Admixtures in High Performance Concrete, Mexico, 1999a.

Houst,Y.F., R.J. Flatt, P. Bowen, and H. Hofmann, Optimization of


superplasticizers: From research to application, pp. 121-134, J.G. Cabrera, and R.R.
Villarreal (eds) Proceedings of International Symposium on the Role of Admixtures in
High Performance Concrete, Mexico, 1999b.

Houst, Y.F., P. Bowen, and F. Perche, Towards tailored superplasticizers. pp. 11-20.
R.K. Dhir, P.C. Hewlett and M.D. Newlands (eds) Proceedings of International

250
Conference on Admixtures-Enhancing Concrete Performance, , Thomas Telford,
London, 2005.

Hsu, K.C., J.J. Chiu, S.D. Chen, and Y.C. Tseng (1999) Effect of addition time of a
superplasticizer on cement adsorption and on concrete workability. Cement and
Concrete Composites, 21, 425-430.

Hui, W.U., G. Huiling, L. Jiaheng, Z.Rongguo, and L Yong (2007) Research on


synthesis and action mechanism of polycarboxylate superplasticizer. Front. Chem.
China, 2, 322-325.

IS 383 (2002) Indian standard specification for coarse and fine aggregates from
natural sources for concrete. Bureau of Indian Standards, New Delhi.

IS 1199 (2004) Indian standard specification for methods of sampling and analysis of
concrete. Bureau of Indian Standards, New Delhi.

IS 2386 (2007) Indian standard specification for methods of test for aggregates of
concrete. Bureau of Indian Standards, New Delhi.

IS 4031 – Parts 2, 4 and 5 (2005) Indian standard specification for method of


physical tests for hydraulic cement. Bureau of Indian Standards, New Delhi.

IS 8142 (2002) Indian standard specification for method of test for determining
setting time of concrete by penetration resistance. Bureau of Indian Standards, New
Delhi.

IS 9103 (2004) Indian standard specification for Concrete Admixtures. Bureau of


Indian Standards, New Delhi.

IS 10262 (2004) Recommended guidelines for concrete mix design, Bureau of Indian
Standards, New Delhi.

IS 12269 (2004) Indian standard specification for 53 grade ordinary Portland cement.
Bureau of Indian Standards, New Delhi.

Jiang, S., B.G. Kim, and P.C. Aїtcin (1999) Importance of adequate soluble alkali
content to ensure cement/superplasticizer compatibility. Cement and Concrete
Research, 29, 71-78.

Johnston, C.D. (1987) Admixture-cement incompatibility: A case history. Concrete


International, 9, 51-60.

Jolicoeur, C. and M.A. Simard (1998) Chemical admixture–cement interactions:


Phenomenology and physico-chemical concepts. Cement and Concrete Composites,
20, 87-101.

Justnes, H., M.V. Dooren, and D.V. Gemert, Reasons for workability loss in
cementitious binders. pp 53-65. V.M. Malhotra (ed) Seventh CANMET/ACI
International Conference on Superplasticizers and Other Chemical Admixtures in
Concrete, Berlin, Germany, 2003

251
Kantro, D.L. (1980) Influence of water-reducing admixtures on properties of cement
paste-A miniature slump test. Cement, Concrete and Aggregates, 2,
95- 102.

Kapelko, A. (2006) The possibility of adjusting concrete mixtures’ fluidity by means


of superplasticizer SNF. Archives of Civil and Mechanical Engineering, VI, Wroclaw
University of technology, 3, 37- 53.

Khayat, K. H. and Z. Guizani (1997) Use of viscosity modifying admixture to


enhance stability of fluid concrete. ACI Materials journal, 94, 332-340.

Khayat, K.H. and A. Yahia (1997) Effect of welangum-high range water reducer
combinations on rheology of cement grout. ACI Materials journal, 94, 365-372.

Khayat, K.H. and A. Yahia (1998) Simple field tests to characterize fluidity and
washout resistance of structural cement grout. Cement, Concrete and Aggregates, 20,
145-156.

Khayat, K.H., M. Saric-Coric, and F. Liotta (2002) Influence of thixotropy


onstability characteristics of cement grout and concrete. ACI Materials Journal, 99,
234-241.

Khayat, K.H., A. Yahia, and M. Sayed (2008) Effect of supplementary cementitious


materials on rheological properties, bleeding and strength of structural grout. ACI
Materials Journal, 105, 585-593.

Khatib, J.M. and S. Wild (1996) Pore size distribution of metakaolin paste. Cement
and Concrete Research, 26, 1545-1553.

Khatib, J.M. and P.S. Mangat (1999) Influence of superplasticizer and curing on
porosity and pore structure of cement paste. Cement and Concrete Composites, 21,
431-437.

Kim, B.G., S. Jiang, and P.C. Aїtcin, Influence of molecular weight of PNS
superplasticizers on the properties of cement pastes containing different alkali
contents. J.G. Cabrera, and R.R. Villarreal (eds) Proceedings of International
Symposium on the Role of Admixtures in High Performance Concrete, Mexico, 1999.

Kim, B.G., S. Jiang, C. Jolicoeur, and P.C. Aїtcin (2000) The adsorption behaviour
of PNS superplasticizer and its relation to fluidity of cement paste. Cement and
Concrete Research, 30, 887-893.

Kim, B.G. and P.C. Aїtcin, Dispersing mechanism of PNS superplasticizers in high
performance concrete. G. Grieve and G. Owens (eds) Proceedings of 11th
international congress on the chemistry of cement, South Africa, 11-16 May, Durban,
South Africa, 2003.

Kim, B.G., T.H Ahn,. B.G. Kang and Y.T. Kim, Effect of superplasticizer type on
the properties of high performance concrete incorporating metakaolin,. pp. 83-
97.Supplementary papers compiled by P.C. Nkinamubanzi, Seventh CANMET/ACI
International Conference on Superplasticizers and Other Chemical Admixtures in
Concrete, Berlin, Germany, 2003.

252
Kirca, O., L. Turanli, and T.Y. Erdogan (2002) Effects of retempering on
consistency and compressive strength of concrete subjected to prolonged mixing.
Cement and Concrete Research, 32, 441-445.

Kjellsen, K.O., R.J. Detwiler, and O.E. Gjorv (1991) Backscattered Electron image
analysis of cement paste specimens: specimen preparation and analytical methods.
Cement and Concrete Research, 21, 388-390.

Krishna, R.N. (1993) Compatibility and strength characteristics of superplasticized


high strength concrete. Doctoral thesis, Anna University, Madras, India.

Krishna, R.N. (1996) Dispersing action of a superplasticizer with different grades of


cements and flyash. ACI materials Journal, 93, 351-355.

Kumar, P. and S.K. Kaushik (2003) Some trends in the use of concrete: Indian
Scenario. The Indian Concrete Journal, December, 1503-1508.

Kumar, J.V. (2006) Study of cement-superplasticizer compatibility in concrete.


M.tech Project Report, Indian Institute of Technology, Madras, India.

Lachemi, M., K.M.A Hossain, V. Lambros, P.C. Nkinamubanzi, and N.


Bouzouba (2004) Self compacting concrete incorporating new Viscosity modifying
admixtures. Cement and Concrete Research, 34, 917-926.

Lachemi, M., K.M.A. Hossain, R. Patel, M. Shehata, and N. Bouzoubaa (2007)


Influence of paste/mortar rheology on the flow characteristics of high volume fly ash
self consolidating concrete. Magazine of Concrete Research, 59, 517- 528.

Lapasin, R., A. Papo, S. Rajgelj (1983) The phenomenological description of the


thixotropic behaviour of fresh cement pastes. Rheologica Acta, 22, 410-416.

Laskar, A.I. and S. Talukdar (2007) Correlation between compressive strength and
rheological parameters of high performance concrete. Research Letters in Materials
Science, Article ID 45869, 1-4.

Li, Z., and Z. Ding (2003) Property improvement of portland cement by


incorporating with metakaolin and slag. Cement and Concrete Research, 33, 579-584.

Li, G., A. Tagnit-Hamou, and P.C. Aïtcin, Improving cement-superplasticizer


compatibility by using soluble alkalis as a chemical additive in concrete. pp.655-665.
G. Grieve and G. Owens (eds) Proceedings of 11th international congress on the
chemistry of cement, South Africa, 11-16 May, Durban, South Africa, 2003.

Liao, T.S., C.L. Hwang, Y.S Ye, K.C. Hsu (2006) Effects of a carboxylic acid/
sulfonic acid copolymer on the material properties of cementitious materials. Cement
and Concrete Research, 36, 650-655.

Lim, G.G., S.S. Hong, D.S. Kim, B.J. Lee, and J.S. Rho (1999) Slump loss control
of cement paste by adding polycarboxylic type slump- releasing dispersant. Cement
and Concrete Research, 29, 223-229

253
Lootens D., P. Hebraud, E. Lecolier, and H.V. Damme (2004) Gelation, shear-
thinning and shear-thickening in cement slurries. Oil and Gas Science andTechnology
– Rev. IFP, 59, 31-40.

Maeder, U. and I. Schober, Performance of blends of polycarboxylate polymers in


different cements. G. Grieve and G. Owens (ed) Proceedings of 11th international
congress on the chemistry of cement, 11-16 May, Durban, South Africa, 2003.

Magarotto, R., I. Torresan, and N. Zeminian, Influence of molecular weight of


polycarboxylate ether superplasticizers on the rheological properties of fresh cement
pastes, mortar and concrete. pp. 514-526. G. Grieve and G. Owens (eds) Proceedings
of 11th international congress on the chemistry of cement. 11-16 May, Durban, South
Africa, 2003.

Mailvaganam, N.P. (1999) Chemical admixtures in concrete- side effects and


compatibility problems, The Indian Concrete Journal, 73, 367-374.

Mailvaganam, N. P. (2001) How chemical admixtures produce their effects in


concrete? Indian Concrete Journal, 75, 331-334.

Mansoutre S., P. Colombet, H. Van Damme (1999) Water retention and granular
rheological behaviour of fresh C3S paste as a function of concentration. Cement and
Concrete Research, 29, 1441-1453

Masood, I. and Agarwal (1994) Effect of various superplasticizers on rheological


properties of cement paste and mortars. Cement and Concrete Research, 24, 291- 302.

McCarter, W.J. and P.N. Curran (1984) The electrical response characteristics of
setting of cement paste. Magazine of Concrete Research, 36, 42-49.

McCarter, W. J (1987) Gel formation during early hydration. Cement and Concrete
Research, 17, 55-64.

McCarter, W.J. and D. Tran (1996) Monitoring pozzolanic activity by direct


activation with calcium hydroxide. Construction and Building materials, 10, 179-184.

McCarter, W.J., T.M. Chrisp, and G. Starrs (1999) The early hydration of alkali-
activated slag: Developments in monitoring techniques. Cement and Concrete
Composites, 21, 277-283.

McCarter, W.J., T.M. Chrisp, G. Starrs, and J. Blewett (2003) Characterisation


and monitoring of cement based systems using intrinsic electrical property
measurements. Cement and Concrete Research, 33, 197-206.

Mehta, P.K. and P.J.M. Monteiro, Concrete-Microstructure, properties and


materials, Indian Concrete Institute, Chennai, India, 2005.

Meyer, L.M. and W.F. Perenchio (1979) Theory of concrete slump loss as related to
the use of chemical admixtures, Concrete International, 1, 36-43.

Min B.H., L. Erwin, and H.M. Jennings (1994) Rheological behaviour of fresh
cement paste as measured by squeeze flow, Journals of Materials Science, 29, 1374-
1381.
254
Mollah, M.Y.A., P. Palta, and T.R. Hess, R.K. Vempati, and D. L. Cocke (1995)
Chemical and physical effects of sodium lignosulphonate superplasticizer on the
hydration of portland cement and solidification/stabilization consequences. Cement
and Concrete Research, 25, 671-682.

Monte R., and A.D. Figueiredo, Comparative evaluation of test methods for
determination of superplasticizers saturation dosages in cement pastes. pp 147-159.
V.M. Malhotra (ed) Seventh CANMET/ACI International Conference on
Superplasticizers and Other Chemical Admixtures in Concrete, Berlin, Germany,
2003.

Mork, J.H. and O.E Gjoerv (1997) Effect of gypsum-hemihydrate ratio in cement
on rheological properties of fresh concrete. ACI Materials Journal, March-April, 142-
146.

Mullick, A.K. (2008) Cement-superplasticizer compatibility and method of


evaluation, The Indian Concrete Journal, 82, 8-16.

Nagataki, S., E. Sakai, and T. Takeuchi (1984) The fluidity of flyash-cement paste
with superplasticizer. Cement and Concrete Research, 14, 631-638.

Nakajima, Y. and K. Yamada (2004) The effect of the kind of calcium sulphate in
cements on the dispersing ability of poly β naphthalene sulphonate condensate
superplasticizer. Cement and Concrete Research, 34, 839-844.

Nawa, T., H. Ichiboji, and M. Kinoshita, Influence of temperature on fluidity of


cement paste containing superplasticizer with polyethylene oxide graft chains. Pp.
195-210. V.M Malhotra (ed) Sixth CANMET/ ACI International Conference on
Superplasticizers and Other Chemical Admixtures in Concrete, Ottawa, Canada, 2000

Nehdi, M and M.A. Rahman (2004) Estimating rheological properties of cement


pastes using various rheological models for different test geometry, gap and surface
friction. Cement and Concrete Research, 34, 1993-2007.

Neville, A.M. Properties of concrete, ELBS with Longman, Harlow, UK, 2000.

Nguyen, V.H., S. Remond, J.L. Gallias, J.P. Bigas, and P. Muller (2006) Flow of
Herschel- Bulkley fluids through the Marsh cone, Journal of Non Newtonian Fluid
mechanics, 139, 128- 134.

Nkinamubanzi, P.C., B.G. Kim, and P.C. Aїtcin, Some key cement factors that
control the compatibility between naphthalene-based superplasticizers and ordinary
Portland cements. pp.33-54 V.M. Malhotra (ed) 6th CANMET/ACI International
Conference on superplasticizers and other chemical admixtures in concrete, Paris,
2000.

Nkinamubanzi, P.C. and P.C. Aїtcin (2004) Cement and superplasticizer


combinations: Compatibility and Robustness. Cement, Concrete and Aggregates, 26,
1-8.

Ohta, A., T. Sugiyama, and T. Uomoto, Study of dispersing effects of


polycarboxylate based dispersant on fine particles, pp. 211-227. V.M Malhotra (ed)

255
Sixth CANMET/ ACI International Conference on Superplasticizers and Other
Chemical Admixtures in Concrete, Ottawa, Canada, 2000

P 18-358 (1985) Adjuvants pour bètons, mortiers et coulis. Coulis courants


d’injection pour prècontrainte. Measure de la fluiditè et da la reduction d’eau,
Association francaise de normalization (afnor), Juillet (French Code).

Page, M., P.C Nkinamubanzi, and P.C. Aїtcin, The cement/superplasticizer


compatibility: A headache for superplasticizer manufacturers. J.G. Cabrera, and
R.R. Villarreal (eds) Proceedings of International Symposium on the Role of
Admixtures in High Performance Concrete, Mexico, 1999.

Papayianni, I., G. Tsohos, N. Oikonomou, and P. Mavria (2005) Influence of


superplasticizer type and mix design parameters on the performance of them in
concrete mixtures. Cement and Concrete Composites, 27, 217-222.

Papo, A., and L. Piani (2004) Effect of various superplasticizers on the rheological
properties of Portland cement pastes. Cement and Concrete Research, 34, 2097-2101.

Park C.K., M.H. Noh, and T.H. Park (2005) Rheological properties of cementitious
materials containing mineral admixtures. Cement and Concrete Research, 35, 842-
849.

Phan, T.H., M. Chaouche, and M. Moranville (2006) Influence of organic


admixtures on the rheological behaviour of cement pastes. Cement and Concrete
Research, 36, 1807-1813.

Plank, J. and C. Hirsch (2007) Impact of zeta potential of early cement hydration
phases on superplasticizer adsorption. Cement and Concrete Research, 37, 537-542.

Porter, M.R. Handbook of surfactants. Blackie Academic and Professional, United


Kingdom, 1994.

Prince, W., M. Edwards-Lajnef, and P.C. Aïtcin (2002) Interaction between


ettringite and polynaphthalene sulphonate superplasticizer in cementitious paste.
Cement and Concrete Research, 32, 79-85.

Prince, W., M. Espagne, and P.C. Aїtcin (2003) Ettringite formation: A crucial step
in cement superplasticizer compatibility. Cement and Concrete Research, 33, 635-
641.

Puertas, F., H. Santos, M. Palacios, and S. Martinez-Ramirez (2005)


Polycarboxylate Superplasticizer admixtures: effects on hydration, microstructure and
rheological behaviour in cement pastes. Advances in Cement Research, 17,
77- 89.

Rahman, M.A and M. Nehdi (2005) Empirical correlations between rheological


properties of cement pastes from various models. The Indian Concrete Journal,
October, 52- 60.

Ramachandran, V S., M.S Lowery, V. Malhotra, (1995) Behaviour of ASTM Type


V cement hydrated in the presence of sulfonated melamine formaldehyde. Materials
and Structures, 28, 133-138.
256
Ramachandran, V.S., V.M Malhotra, C. Jolicoeur, and N. Spiratos,
Superplasticizers: Properties and Applications in Concrete. Materials Technology
Laboratory, CANMET, Canada, 1998.

Ramachandran, V.S. and J.J. Beaudoin, Handbook of analytical techniques in


concrete science and technology. Noyes and William Andrew Publications, USA,
2001.

Ramachandran, V.S. (2002) Concrete admixtures handbook. Standard Publishers,


New Delhi, India.

Ramezanianpour, A.A., V. Sivasundaram, and V.M. Malhotra (1995)


Superplasticizers: Their effect on the strength properties of concrete. Concrete
International, April, 30-35.

Ravina,D. and I. Soroka (1994) Slump loss and compressive strength of concrete
Made with WRR and HRWR admixtures and subjected to prolonged mixing. Cement
and Concrete Research, 24, 1455-1462.

Rixom R and N. Mailvaganam, Chemical admixtures for concrete. E&FNSpon,


London, UK, 1999.

Roberts, L.R. (1995) Dealing with cement-admixture interactions. 23 rd Annual


convention of the Institute of concrete technology ,11th April, Telford. p.16

Roncero, J., R. Gettu, P.C.C Gomes, and, L. Agullo (2000) Study of flow
behaviour of superplasticized cement paste systems and its influence on properties of
fresh concrete. pp. 273-294. High Performance Concrete: Research to practice,
American Concrete Institute, USA.

Roncero, J. (2000) Effect of superplasticizers on the behaviour of concrete in the


fresh and hardened states: Implications for high performance concretes, Doctoral
thesis, Universitat Politecnica de Catalunya, Barcelona, Spain.

Roncero, J., R. Gettu, E. Vazquez, and J.M. Torrents, Effect of superplasticizer


content and temperature on the fluidity and setting of cement pastes. pp. 343-356.
J.G. Cabrera, and R.R. Villarreal (eds) Proceedings of International Symposium on
the Role of Admixtures in High Performance Concrete, Mexico, 1999.

Roncero, J., S. Valls, and R.Gettu (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, 22, 103-108.

Roussel, N. (2005) Steady and transient flow behaviour of fresh cement pastes.
Cement and concrete Research, 35, 1656-1664.

Roussel, N. and R.L. Roy (2005) The Marsh cone: a test or a rheological apparatus?
Cement and Concrete Research, 35, 823-830.

Roussel, N and P. Coussot (2005) “Fifty cent rheometer” for yield stress
measurements: From slump to spreading flow. Journal of Rheology, 49, 705-718.

257
Roussel, N. (2006) Correlation between yield stress and slump: Comparison between
numerical simulations and concrete rheometers results. Materials and Structures, 39,
501-509.

Roy, R.L. and N. Roussel (2005) The Marsh cone as a viscometer: theoretical
analysis and practical limits. Materials and Structures, 38, 25-30.

Roy, D.M. and K. Asaga (1980) Rheological properties of cement mixes: V. The
effects of time on viscometric properties of mixes containing superplasticizers;
conclusions. Cement and Concrete Research, 10, 389-394.

Sakai, E. and M. Daimon, Mechanisms of superplastification. pp.91-111. J. Skalny


and S. Mindess (eds) The American Ceramic Society Publication, USA, 1995.

Sakai, E., K. Yamada, and A. Ohta (2003) Molecular structure and dispersion-
adsorption mechanisms of comb-type superplasticizers used in Japan. Journal of
Advanced Concrete Technology, 1, 16-25.

Sakai, E., T. Kasuga, T. Sugiyama, K. Asaga, and M. Daimon (2006) Influence of


superplasticizers on hydration of cement and pore structure of hardened cement.
Cement and Concrete Research, 36, 2049-2053.

Schmidt, G. and E. Schlegel (2002) Rheological characterization of C-S-H phases-


water suspensions. Cement and Concrete Research, 32, 593-599.

Schultz, M.A. and L.J. Struble (1993) Use of oscillatory shear to study flow
behaviour of fresh cement paste. Cement and Concrete Research, 23, 273- 282.

Schwartzentruber, L.D., R. Roy, and J. Cordin (2006) Rheological behaviour of


fresh cement pasts formulated from a self compacting concrete (SCC). Cement and
Concrete Research, 36, 1203-1213.

Scrivener, K.L., The microstructure of Concrete. pp. 127-157. J.P. Skalny (ed)
Materials science of concrete I, The American ceramic society publication,
Westerville, 1989.

Seung-Bum, P. (1999) The effects of superplasticizers on the engineering properties


of plain concrete. KCI Concrete Journal, 11, 29-43.

Shah, S. (2006) Gypsum, a valuable input for agriculture, John Harrison Publications
(visited on 8.11.08)
(www.allotment.org.uk/articles2/Gypsum_Valuable_Input_for_Agriculture.php)

Simard, M.A., P.C. Nkinambanzi, and C. Jolicoeur (1993) Calorimetry Rheology


and compressive strength of superplasticized cement pastes, Cement and Concrete
Research, 23, 939-950.

Singh, N.B., R. Sarvahi, and N.P Singh (1992) Effect of superplasticizers on the
hydration of cement. Cement and Concrete Research, 22, 725-735.

Sonebi, M. (2006) Rheological properties of grouts with viscosity modifying agents


as diutan gum and welan gum incorporating pulverized fly ash. Cement and Concrete
Research, 36, 1609-1618.
258
Soroka I., Portland Cement Paste and Concrete. The Macmillan Press Limited,
London, 1979.

Spiratos, N., M. Page, N.P. Mailvaganam, V.M. Malhotra, and C. Jolicoer,


(2003) Superplasticizers for Concrete- Fundamentals, Technology and Practice,
Supplementary Cementing materials for Sustainable Development, Ottawa, Canada.

Spiratos, N. and C. Jolicoeur (2006) Concrete chemical admixtures: Perspective and


challenges. Indian Concrete Institute Journal, 6, 23-32.

Srinath, B. (2007) Use of non destructive techniques to analyze fresh and hardened
state properties of concrete. Master of Science Thesis, Indian Institute of Technology,
Madras, India.

Struble L.J. and M.A. Schultz (1993) Using creep and recovery to study flow
behaviour of fresh cement paste. Cement and Concrete Research, 23, 1369- 1379.

Struble, L. and G.K. Sun (1995) Viscosity of Portland cement paste as a function of
concentration. Advanced cement based materials, 2, 62-69.

Struble L.J. and Leit (1995) Rheological changes associated with setting of cement
Paste. Advanced cement based materials, 2, 224-230.

Struble, L., R. Szecsy, W.G. Lei and G.K. Sun (1998) Rheology of cement paste
and concrete, Cement, Concrete and Aggregates, 20, 269-277.

Struble L.,T.Y. Kim, and H.Zhang (2001) Setting of cement and concrete. Cement,
Concrete and Aggregates, 23, 88-93.

Sugiyama, T., A. Ohta and T. Uomota, The dispersive Mechanism and applications
of polycarboxylate based superplasticizers. pp. 560-568. G. Grieve and G. Owens
(eds) Proceedings of 11 th international congress on the chemistry of cement, Durban,
South Africa, 2003.

Sun Z., T. Voigt, and S.P. Shah (2006) Rheometric and ultrasonic investigations of
viscoelastic properties of fresh Portland cement pastes. Cement and Concrete
Research, 36, 278-287

Swamy R.N. (1991) Mineral admixtures for high strength concrete. The Indian
Concrete Journal, 65, 265-271.

Tagmit–Hamou, G., A. Li, and P.C Aїtcin, (2003) Improving cement


superplasticizer compatibility by using soluble alkalis as a chemical additive in
concrete. Proceedings of 11 th international congress on the chemistry of cement,
Durban, South Africa.

Tanaka, Y., A. Ohta, and T. Sugiyama (1999) Polycarboxylate based advanced


superplasticizers for high performance concrete. pp. 135-142. J.G. Cabrera, and
R.R. Villarreal (eds) Proceedings of the International Symposium on the Role of
Admixtures in High Performance Concrete, Mexico, 1999.

259
Tang, C.W., T. Yen, C.S. Chang, and K.H. Chen (2001) Optimizing mixture
proportions for flowable high performance concrete via rheology tests. ACI Materials
Journal, 98, 493-502.

Tattersall G.H., P.F.G Banfill, The rheology of fresh concrete. Pitman Books
Limited, London, 1983.

Taylor, H.F.W., Cement Chemistry, Thomas Telford, London UK, 1997.

Toralles-Carbonari B., R. Gettu, L. Agullo, A. Aguado, and V.A. Acena (1996) A


Synthetic approach for the experimental optimization of high strength concrete. pp.
161-168. F. de Larrard and R. Lacroix (eds) Proceedings of fourth International
Symposium on Utilization of High Strength/ High Performance Concrete, Paris, 1996.

Torrents, J.M., J. Roncero, and R. Gettu (1998) Utilization of impedence


spectroscopy for studying the retarding effect of a superplasticizer on the setting of
the cement. Cement and Concrete Research, 28, 1325-1333.

Torresan, I. and R. Khurana, New superplasticizers based on modified melamine


polymer. pp. 235-253, V.M Malhotra, (ed) Recent Advances in Concrete Technology,
Proceedings Fourth CANMET/ACI/JCI International Conference, ACI International,
Tokushima, Japan, 1998.

Tregger, N., L. Ferrara, and S.P. Shah (2008) Identifying viscosity of cement paste
from mini-slump flow test. ACI Materials Journal, 105, 558-566.

Tzouvalas, G., A. Papageorgiou, S. Tsimas, and D. Papageorgio (2005) Study of


early hydration of cement pastes containing alternative calcium sulphate bearing
materials, Proceeding of International Conference on Use of Foamed Concrete in
Construction, Dundee, UK, 41-51.

Uchikawa, H., D. Sawaki, and S. Hanehara (1995) Influence of kind and added
timing of of organic admixture on the composition, Structure, and property of fresh
cement paste. Cement and Concrete Research, 25, 353-364.

Uchikawa, H., S. Hanehara, and D. Sawaki (1997) The role of steric repulsive force
in the dispersion of cement particles in fresh paste prepared with organic admixture.
Cement and Concrete Research, 27, 37-50.

Van Damme H., S. Mansoutre, P. Colombet, C. Lesaffre, and D. Picart (2002)


Pastes: Lubricated and cohesive granular media. C.R Physique, 3, 229-238

Vandanjon, P.O., F. de Larrard, B. Dehouse, G. Villain, R. Maillot, and P.


Laplante (2003) Homogenisation of concrete in a batch plant: The influence of
mixing time and method on the introduction of mineral admixtures. Magazine of
Concrete Research, 55, 105-116.

Vikan, H., H. Justnes, F. Winnefeld, and R. Figi (2007) Correlating cement


characteristics with rheology of paste. Cement and Concrete Research, 37, 1502-
1511.

Wallevik, J.E. (2005) Thixotropic investigation on cement paste: Experimental and


numerical approach. Journal of Non-Newtonian Fluid Mechanics, 132, 86-99.
260
Williams, D.A., A.W. Saak, and H.M. Jennings (1999) The influence of mixing on
the rheology of fresh cement paste. Cement and Concrete Research, 29, 1491-1496.

Wineman, A.S. and K.R. Rajagopal, Mechanical response of polymers. Cambridge


University Press, United Kingdom, 2000.

Winnefeld, F., S. Becker, J. Pakusch and T. Gotz (2007) Effects of the molecular
architecture of comb shaped superplasticizers on their performance in cementitious
system. Cement and Concrete Composites, 29, 251-262.

Xiao, L.Z., Z.J. Li and X.S. Wei (2007) Selection of superplasticizer in concrete mix
design by measuring the early electrical resistivities of pastes. Cement and Concrete
Composities, 29, 350-356.

Yamada, K., T. Takahashi, S. Hanehara, and M. Matsuhisa (2000) Effects of the


chemical structure on the properties of polycarboxylate type superplasticizer. Cement
and Concrete Research, 30, 197-207.

Yamada, K., S. Ogawa, and Hanehara (2001) Controlling of the adsorption and
dispersing force of polycarboylate-type superplasticizer by sulfate ion concentration
in aqueous phase. Cement and Concrete Research, 31, 375-383.

Yamada, K. and S. Hanehara, Working mechanism of polycarboxylate


superplasticizer considering the chemical structure and cement characteristics. pp.
538-549. G. Grieve and G. Owens (eds) Proceedings of 11 th international congress
on the chemistry of cement, Durban, South Africa, 2003.

Yamada, K., T. Sugamata, and H. Nakanishi (2006) Fluidity performance


evaluation of cement and superplasticizers. Journal of Advanced Concrete
Technology, 4, 241-249.

Yang, M., and H.M. Jennings (1995) Influences of mixing methods on the
microstructure and rheological behaviour of cement paste. Advanced Cement Based
Materials, 2, 70-78.

Yen, T., C.W. Tang, C.S. Chang, and K.H. Chen (1999) Flow behaviour of high
strength high performance concrete. Cement and Concrete Composites, 21, 413-424.

Yoshioka, K., E. Tazawa, K. Kawai, and T. Enohata (2002) Adsorption


characteristics of superplasticizers on cement component minerals. Cement and
Concrete Research, 32, 1507-1513

Young, J.F. (2008) Looking ahead from the past: The heritage of cement chemistry.
Cement and Concrete Research, 38, 111-114.

Yuan, C.Z., and W.J Guo (1987) Bond between marble and cement paste. Cement
and Concrete Research, 17, 544-552.

Zhang, T., S. Shang, F. Yin, A. Aishah, A. Salmiah and T.L. Ooi (2001)
Adsorptive behaviour of surfactants on surface of portland cement. Cement and
Concrete Research, 31, 1009-1015.

261
Zhang, X. (2007) Quantitative microstructural characterization of concrete cured
under realistic temperature conditions. Doctoral thesis, Ecole Polytechnique Federale
de Lausanne, Zwitzerland.

Zhao, H. and D. Darwin (1992) Quantitative backscattered electron analysis of


cement paste. Cement and Concrete Research, 22, 695-706.

Zhor, J. and T.W. Bremner, Advances in evaluation of lignosulphonates as concrete


admixtures. pp. 1011-1042. V.M Malhotra (ed) Recent Advances in Concrete
Technology, Proceedings of Fourth CANMET/ACI/JCI International Conference, ACI
International, Tokushima, Japan, 1998.

Zingg, A., L. Holzer, A. Kaech, F. Winnefeld, J. Pakusch, S. Becker, L. Gauckler.


(2008) The microstructure of dispersed and non-dispersed fresh cement pastes – New
insight by cryo-microscopy. Cement and Concrete Research, 38, 522-529.

262
LIST OF PAPERS SUBMITTED ON THE BASIS
OF THIS THESIS

I REFERRED JOURNALS

1. C Jayasree and Ravindra Gettu (2008) Experimental Study of the Flow


Behaviour of Superplasticized Cement Paste. Materials and Structures,
RILEM, 41, 1581-1593.

2. C Jayasree and Ravindra Gettu. Correlations between the Fresh and


Hardened Properties of Superplasticized Paste, Mortar and Concrete, Indian
Concrete Journal (Under review)

II PRESENTATIONS IN CONFERENCES

1. C Jayasree and Ravindra Gettu Influence of Mixing Method on the Fluidity


of Superplasticized Cement Paste. Proceedings of the Fifth Asian Symposium
on Polymers in Concrete, Chennai, India, 11-12 September 2006, 665-670.

2. C Jayasree and Ravindra Gettu. Optimization of Superplasticized Cement


Paste and its Correlation with Fresh Concrete Behaviour. Proceedings of the
International Conference on Advances in Concrete and Construction,
Hyderabad, India, 7-9 February 2008, 1042-1052.

3. C Jayasree and Ravindra Gettu. Mix Design of Self Compacting Concrete


with Viscosity Modifying admixtures, International Conference on Innovative
World of Concrete, 11-14 December 2008, New Delhi, India.

263

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy