Download as pdf or txt
Download as pdf or txt
You are on page 1of 420

Although acid-base cements have been known since the mid 19th

century, and have a wide variety of applications, there has been a failure
to recognize them as constituting a single, well-defined class of material.
This book remedies the situation by unifying the subject and treating this
range of materials as a single class.
These cements are defined as materials that are formed by mixing a
basic powder with an acidic liquid, and offer an alternative to polymer-
ization as a method for forming solid substances. They are quick-setting
materials, with unusual properties, which find diverse applications as
biomaterials and in industry.
Chemistry of Solid State Materials
Acid-base cements
Their biomedical and industrial applications
Chemistry of Solid State Materials
Series Editors
A. R. West, Department of Chemistry, University of Aberdeen
H. Baxter, formerly at the Laboratory of the Government Chemist,
London

1 Segal: Chemical synthesis of advanced ceramic materials


2 Colomban: Proton conductors
3 Wilson & Nicholson: Acid-base cements
Acid-base cements
Their biomedical and industrial
applications

Alan D. Wilson
formerly Head, Materials Technology, Laboratory of the Government Chemist
Senior Research Fellow, Eastman Dental Hospital

John W. Nicholson
Head, Materials Research, Laboratory of the Government Chemist

m CAMBRIDGE
0 UNIVERSITY PRESS
CAMBRIDGE UNIVERSITY PRESS
Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, Sao Paulo

Cambridge University Press

The Edinburgh Building, Cambridge CB2 2RU, UK

Published in the United States of America by Cambridge University Press, New York

www.cambridge.org

Information on this title: www.cambridge.org/9780521372220

© Cambridge University Press 1993

This book is in copyright. Subject to statutory exception


and to the provisions of relevant collective licensing agreements,
no reproduction of any part may take place without
the written permission of Cambridge University Press.
First published 1993

This digitally printed first paperback version 2005

A catalogue recordfor this publication is available from the British Library

Library of Congress Cataloguing in Publication data


Wilson, Alan D.
Acid—base cements: their biomedical and industrial applications /Alan D.
Wilson, John W. Nicholson
p. cm. - (Chemistry of solid state materials; 3)
Includes bibliographical references and index.
ISBN 0-521-37222-4
1. Adhesives. 2. Dental cements. I. Nicholson, John W.
II. Title. III. Series.
TP968.W54 1993
620.1'35-dc20 91-38946 CIP
ISBN-13 978-0-521-37222-0 hardback
ISBN-10 0-521-37222-4 hardback

ISBN-13 978-0-521-67549-9 paperback


ISBN-10 0-521-67549-9 paperback
Dedicated to the past and present members of the
Materials Technology Group at the Laboratory of the
Government Chemist
Contents

Preface xvii
Acknowledgements xix
1 Introduction 1
References 4

2 Theory of acid-base cements 5


2.1 General 5
2.2 The formation of cements 7
2.2.1 Classification 7
2.2.2 Requirements for cementitious bonding 8
2.2.3 Gelation 10
2.3 Acid-base concepts 12
2.3.1 General 12
2.3.2 History of acid-base concepts 12
2.3.3 Acid-base concepts in AB cement chemistry 14
2.3.4 Relevance of acid-base theories to AB cements 19
2.3.5 Acid-base strength 20
2.3.6 Acid-base classification 22
2.3.7 Hard and soft acids and bases (HSAB) 24
References 26

3 Water and acid-base cements 30


3.1 Introduction 30
3.1.1 Water as a solvent 30
3.1.2 Water as a component 30
3.2 Water 31
3.2.1 Constitution 31
3.2.2 Water compared with other hydrides 33
3.3 The structure of water 34
3.3.1 The concept of structure in the liquid state 34
3.3.2 The structures of ice 35
3.3.3 Liquid water 36
3.4 Water as a solvent 40

IX
Contents

3.4.1 Hydrophobic interactions 40


3.4.2 Dissolution of salts 41
3.4.3 Ion-ion interactions in water 44
3.4.4 Dissolution of polymers 45
3.5 Hydration in the solid state 47
3.5.1 Coordination of water to ions 47
3.6 The role of water in acid-base cements 48
3.6.1 Water as a solvent in AB cements 48
3.6.2 Water as a component of AB cements 48
3.6.3 Water as plasticizer 51
References 52

4 Polyelectrolytes, ion binding and gelation 56


4.1 Polyelectrolytes 56
4.1.1 General 56
4.1.2 Polyion conformation 58
4.2 Ion binding 59
4.2.1 Counterion binding 59
4.2.2 The distribution of counterions 59
4.2.3 Counterion condensation 63
4.2.4 Effect of valence and size on counterion binding 65
4.2.5 Site binding - general considerations 67
4.2.6 Effect of complex formation 69
4.2.7 Effect of the polymer characteristics on ion binding 70
4.2.8 Solvation (hydration) effects 72
4.2.9 Hydration of the polyion 73
4.2.10 Hydration and ion binding 76
4.2.11 Desolvation and precipitation 77
4.2.12 Conformational changes in polyions 79
4.2.13 Interactions between polyions 82
4.2.14 Polyion extensions, interactions and precipitation 82
4.3 Gelation 83
References 85

5 Polyalkenoate cements 90
5.1 Introduction 90
5.2 Adhesion 92
5.2.1 New attitudes 92
5.2.2 The need for adhesive materials 92
5.2.3 Acid-etching 93
5.2.4 Obstacles to adhesion 93
5.2.5 The nature of the adhesion of polyalkenoates to tooth
material 94
5.3 Preparation of poly(alkenoic acid)s 97
5.4 Setting reactions 98
Contents
5.5 Molecular structures 99
5.6 Metal oxide polyelectrolyte cements 101
5.7 Zinc polycarboxylate cement 103
5.7.1 Historical 103
5.7.2 Composition 103
5.7.3 Setting and structure 104
5.7.4 Properties 106
5.7.5 Modified materials 112
5.7.6 Conclusions 113
5.8 Mineral ionomer cements 113
5.9 Glass polyalkenoate (glass-ionomer) cement 116
5.9.1 Introduction 116
5.9.2 Glasses 117
5.9.3 Poly(alkenoic acid)s 131
5.9.4 Reaction-controlling additives 133
5.9.5 Setting 134
5.9.6 Structure 143
5.9.7 General characteristics 146
5.9.8 Physical properties 147
5.9.9 Adhesion 152
5.9.10 Erosion, ion release and water absorption 156
5.9.11 Biocompatibility 159
5.9.12 Modified and improved materials 162
5.9.13 Applications 166
5.10 Resin glass polyalkenoate cements 169
5.10.1 General 169
5.10.2 Class I hybrids 170
5.10.3 Class II hybrids 171
5.10.4 Properties 173
References 175

Phosphate bonded cements 197


6.1 General 197
6.1.1 Orthophosphoric acid solutions 197
6.1.2 Cations in phosphoric acid solutions 198
6.1.3 Reactions between oxides and phosphoric acid
solutions 201
6.1.4 Effect of cations in phosphoric acid solutions 203
6.1.5 Important cement-formers 204
6.2 Zinc phosphate cement 204
6.2.1 General 204
6.2.2 History 204
6.2.3 Composition 205
6.2.4 Cement-forming reaction 207
6.2.5 Structure 212

XI
Contents
6.2.6 Properties 214
6.2.7 Factors affecting properties 218
6.2.8 Biological effects 219
6.2.9 Modified zinc phosphate cements 219
6.2.10 Hydrophosphate cements 220
6.3 Transition-metal phosphate cements 220
6.4 Magnesium phosphate cements 222
6.4.1 General 222
6.4.2 Composition 222
6.4.3 Types 222
6.4.4 Cement formation and properties 223
6.4.5 Cement formation with phosphoric acid 223
6.4.6 Cement formation with ammonium dihydrogen
phosphate 223
6.4.7 Cement formation with diammonium hydrogen
phosphate 231
6.4.8 Cement formation with ammonium polyphosphate 232
6.4.9 Cement formation with aluminium acid phosphate 232
6.4.10 Cements formed from magnesium titanates 235
6.5 Dental silicate cement 235
6.5.1 Historical 235
6.5.2 Glasses 237
6.5.3 Liquid 241
6.5.4 Cement-forming reaction 243
6.5.5 Structure 249
6.5.6 Physical properties 253
6.5.7 Dissolution and ion release 255
6.5.8 Biological aspects 260
6.5.9 Conclusions 261
6.5.10 Modified materials 262
6.6 Silicophosphate cement 263
6.7 Mineral phosphate cements 265
References 265

Oxysalt bonded cements 283


7.1 Introduction 283
7.1.1 Components of oxysalt bonded cements 284
7.1.2 Setting of oxysalt bonded cements 284
7.2 Zinc oxychloride cements 285
7.2.1 History 285
7.2.2 Recent studies 286
7.3 Magnesium oxy chloride cements 290
7.3.1 Uses 290
7.3.2 Calcination of oxide 290
7.3.3 Setting chemistry 291

xn
Contents

7.3.4 Kinetics of cementation 293


7.3.5 Phase relationships in the MgO-MgCl2-H2O system 294
7.3.6 Consequences for practical magnesium oxychloride
cements 295
7.3.7 Impregnation with sulphur 297
7.4 Magnesium oxy sulphate cements 299
7.4.1 Setting chemistry 299
7.4.2 Phase relationships in the MgO-MgSO4-H2O system 300
7.4.3 Mechanical properties of magnesium oxysulphate
cements 302
7.5 Other oxy salt bonded cements 304
References 305

8 Miscellaneous aqueous cements 307


8.1 General 307
8.2 Miscellaneous aluminosilicate glass cements 307
8.3 Phytic acid cements 309
8.4 Poly(vinylphosphonic acid) cements 310
8.4.1 Metal oxide polyphosphonate cements 311
8.4.2 Glass polyphosphonate cements 314
8.5 Miscellaneous copper oxide and cobalt hydroxide
cements 315
References 316

9 Non-aqueous cements 318


9.1 General 318
9.2 Zinc oxide eugenol (ZOE) cements 320
9.2.1 Introduction and history 320
9.2.2 Eugenol 321
9.2.3 Zinc oxide 321
9.2.4 Cement formation 322
9.2.5 Setting 323
9.2.6 Structure 331
9.2.7 Physical properties 333
9.2.8 Biological properties 334
9.2.9 Modified cements 334
9.2.10 Impression pastes 335
9.2.11 Conclusions 335
9.3 Improved ZOE cements 336
9.3.1 General 336
9.3.2 Reinforced cements 336
9.4 2-ethoxybenzoic acid eugenol (EBA) cements 337
9.4.1 General 337
9.4.2 Development 337
9.4.3 Setting and structure 339

Xlll
Contents
9.4.4 Properties 340
9.5 EBA-methoxyhydroxybenzoate cements 342
9.5.1 EBA-vanillate and EBA-syringate cements 342
9.5.2 EBA-divanillate and polymerized vanillate cements 344
9.5.3 EBA-HV polymer cements 345
9.5.4 Conclusions 346
9.5.5 Other zinc oxide cements 347
9.6 Calcium hydroxide chelate cements 347
9.6.1 Introduction 347
9.6.2 Composition 348
9.6.3 Setting 348
9.6.4 Physical properties 350
9.6.5 Biological properties 350
9.6.6 The calcium hydroxide dimer cement 351
References 352

10 Experimental techniques for the study of acid-base


cements 359
10.1 Introduction 359
10.2 Chemical methods 360
10.2.1 Studies of cement formation 360
10.2.2 Degradative studies 361
10.3 Infrared spectroscopic analysis 361
10.3.1 Basic principles 361
10.3.2 Applications to AB cements 362
10.3.3 Fourier transform infrared spectroscopy 364
10.4 Nuclear magnetic resonance spectroscopy 364
10.4.1 Basic principles 364
10.4.2 Applications to AB cements 365
10.5 Electrical methods 366
10.6 X-ray diffraction 367
10.6.1 Basic principles 367
10.6.2 Applications to AB cements 368
10.7 Electron probe microanalysis 369
10.7.1 Basic principles 369
10.7.2 Applications to dental silicate cements 369
10.7.3 Applications to glass-ionomer cements 369
10.8 Measurement of mechanical properties 370
10.8.1 Compressive strength 371
10.8.2 Diametral compressive strength 372
10.8.3 Flexural strength 372
10.8.4 Fracture toughness 373
10.9 Setting and rheological properties 374
10.9.1 Problems of measurement 375
10.9.2 Methods of measurement 375

xiv
Contents
10.10 Erosion and leaching 378
10.10.1 Importance in dentistry 378
10.10.2 Studies of erosion 379
10.11 Optical properties 379
10.11.1 Importance in dentistry 379
10.11.2 Measurement of opacity 380
10.12 Temperature measurement 380
10.13 Other test methods 381
References 382

Index 386

xv
Preface

The senior author first became interested in acid-base cements in 1964


when he undertook to examine the deficiencies of the dental silicate cement
with a view to improving performance. At that time there was much
concern by both dental surgeon and patient at the failure of this aesthetic
material which was used to restore front teeth. Indeed, at the time, one
correspondent commenting on this problem to a newspaper remarked that
although mankind had solved the problem of nuclear energy the same
could not be said of the restoration of front teeth. At the time it was
supposed that the dental silicate cement was, as its name implied, a silicate
cement which set by the formation of silica gel. Structural studies at the
Laboratory of the Government Chemist (LGC) soon proved that this view
was incorrect and that the cement set by formation of an amorphous
aluminium phosphate salt. Thus we became aware of and intrigued by a
class of materials that set by an acid-base reaction. It appeared that there
was endless scope for the formulation of novel materials based on this
concept. And so it proved.
Over the years, from 1964 to date, a team at the LGC, with its expertise
in Materials Chemistry, has studied many of the materials described in this
book, elucidating structures, setting reactions and behaviour. This
experience has formed a strong experimental background against which
the book was written. In addition we have maintained contact with leaders
in this field throughout the world. We should mention Professor Dennis
Smith of Toronto University, who amongst his many achievements
invented the adhesive zinc polycarboxylate cement (Chapter 5); Dr G. M.
Brauer, who was for many years at the Institute for Materials Research,
National Bureau of Standards, Washington, D.C., and is the acknowl-
edged authority on cements formed by the reaction between zinc oxide
and phenolic bodies (Chapter 9); and Dr J. H. Sharp of the University of
Sheffield, who has developed magnesium phosphate cements (Chapter 6).

xvu
Preface
In particular we thank Dr J. H. Sharp for supplying original photographs
for use in the section on magnesium phosphate cements and for critically
reading the draft manuscript and making constructive suggestions. On
clinical matters we have benefited from a 20-year collaboration with
Dr J. W. McLean OBE.
Our own research at the LGC, while not confined to, has centred on,
cements formed by the reactions between acid-decomposable glasses and
various cement-forming acids (Chapters 5, 6, 8, 9). One of these materials
invented at the LGC, the glass polyalkenoate or glass-ionomer cement,
has proved of immense importance. Indeed, so successful has this material
been in general dentistry, that the Materials Technology Group earned the
Queen's Award for Technology in 1988. This material illustrates the useful
combination of properties that can be found in the acid-base cements, for
it has the aesthetic appearance of porcelain, the ability to adhere to teeth,
and also the ability to releasefluoridewith its beneficial effect of reducing
caries.
We hope that this work will encourage, stimulate and assist others
choosing to work in this interesting field.

Alan D. Wilson
John W. Nicholson

xvin
Acknowledgements

We make a particular acknowledgement to the late Dr John Longwell


CBE, Deputy Government Chemist in 1964, who encouraged the Labor-
atory to enter the field, and to the line of Government Chemists who
supported the work over the long years; the late Dr David Lewis CB, the
late Dr Harold Egan, Dr Ron Coleman CB (who became Chief Scientist
of the Department of Trade and Industry), Mr Alex Williams CB and
Dr Richard Worswick.
We note the particular contributions of Brian Kent, present Head of the
Materials Technology Group, as co-inventor of the glass polyalkenoate
cement way back in 1968, and of Dr John McLean OBE in developing
clinical applications. It was Surgeon Rear Admiral Holgate CB, OBE,
Chief Dental Officer at the Ministry of Health in 1964, who introduced Dr
McLean to the Laboratory of the Government Chemist (LGC) to initiate
a collaboration that proved so fruitful. Since then there has been constant
support from the Department of Health and its various officers and also
from the British Technology Group, particularly from G. M. Blunt and
R. A. Lane.
Most importantly we acknowledge the contribution of those who
worked at that essential place, the laboratory bench, on which everything
depends.
Our colleagues in the Materials Technology Group (formerly the Dental
Materials Group) who have worked with one or other of us since 1964 are:
R. F. Batchelor, B. G. Lewis, Mrs B. G. Scott, J. M. Paddon, G. Abel,
Dr S. Crisp, A. J. Ferner, Dr H. J. Prosser, M. A. Jennings, Mrs S. A.
Merson, M. Ambersley, D. M. Groffman, S. M. Jerome, D. R. Powis,
Mrs P. J. Brookman (nee Brant), R. P. Scott, J. C. Skinner, Dr R. G. Hill,
G. S. Sayers, Dr C. P. Warrens, Miss A. M. Jackson, Dr J. Ellis,
Miss E. A. Wasson, Miss H. M. Anstice, Dr J. H. Braybrook, Miss S. J.
Hawkins and A. D. Akinmade.

xix
Acknowledgements
In addition we have received support from members of other divisions
at the LGC: Dr R. J. Mesley, M. A. Priguer, D. Wardleworth, Dr I. K.
O'Neill, B. Stuart, R. A. Gilhooley, Dr C. P. Richards, Dr O. M. Lacy
and Dr S. L. R. Ellison.
Guest workers to the Materials Technology Group who have con-
tributed include Professor P. Hotz (Klinik fur Zahnerhaltung der Uni-
versitat, Bern), Ms T. Folleras (NIOM, Scandinavian Institute of Dental
Materials).
Workers in other Government Research Stations and the Universities
who have collaborated with us are: R. P. Miller, D. Clinton, Dr T. I.
Barry, Dr I. Seed (National Physical Laboratory); K. E. Fletcher (Build-
ings Research Station); Miss D. Poynter (Warren Spring Laboratory);
Professor L. Holliday, Dr J.H.Elliott, Dr P. R. Hornsby, Dr K. A.
Hodd, Dr A. L. Reader (Brunei University); R. Manston, Dr B. F.
Sanson, Dr W. M. Allen, P. J. Gleed (Institute for Research on Animal
Diseases); Professor Braden (London Hospital); A. C. Shorthall (Bir-
mingham University), I. M. Brook (University of Sheffield); and
R. Billington (Institute of Dental Surgery, London).
We thank Dr L. J. Pluim of the Rijksuniversiteit te Groningen for
drawing our attention to the early and neglected work of E. van Dalen on
zinc phosphate cements.
We thank Mrs Margaret Wilson for her help in checking the proofs and
the indexing.
We acknowledge the stoic forbearance of our wives in putting up with
the disturbances and neglect of domestic routines and duties occasioned
by the writing of a book.
Alan D. Wilson
John W. Nicholson

xx
1 Introduction

Acid-base (AB) cements have been known since the mid 19th century.
They are formed by the interaction of an acid and a base, a reaction which
yields a cementitious salt hydrogel (Wilson, 1978) and offers an alter-
native route to that of polymerization for the formation of macro-
molecular materials. They are quick-setting materials, some of which have
unusual properties for cements, such as adhesion and translucency.
They find diverse applications, ranging from the biomedical to the
industrial.
Despite all this there has been a failure to recognize AB cements as
constituting a single, well-defined class of material. Compared with organic
polymers, Portland cement and metal alloys, they have been neglected and,
except in specializedfields,awareness of them is minimal. In this book we
attempt to remedy the situation by unifying the subject and treating this
range of materials as a single class.
Human interest in materials stretches back into palaeolithic times when
materials taken from nature, such as wood and stone, were fashioned into
tools, weapons and other artifacts. Carving or grinding of a material is a
slow and time-consuming process so the discovery of pottery, which does
away with the need for these laborious processes, was of the greatest
significance. Here, a soft plastic body, potter's clay, is moulded into the
desired shape before being converted into a rigid substance by firing.
Pottery is but one of a group of materials which are formed by the physical
or chemical conversion of a liquid or plastic body, which can be easily
shaped by casting or moulding, into a solid substance. Other examples of
this common method of fabrication are the casting of metals and the
injection moulding of plastics.
Into this category come the water-based plasters, mortars, cements
and concretes which set at room temperature as the result of a chemical
reaction between water and a powder. Some of these have been known

1
Introduction
since antiquity. The AB cements are related to these materials except that
water is replaced by an acidic liquid.
ThefirstAB cement, the zinc oxychloride cement, was reported by Sorel
in 1855. It was prepared by mixing zinc oxide powder with a concentrated
solution of zinc chloride. Its use in dentistry was recommended by
Feichtinger in 1858 but it did not prove to be a success (Mellor, 1929).
However, other AB cements have proved to be of the utmost value to
dentistry, and their subsequent development has been closely associated
with this art (Wilson, 1978). The AB cements, developed against the
backcloth of the severe demands of dentistry, have interesting properties.
Some possess aesthetic appeal and the ability to adhere to base metals and
other reactive substrates. Most have superior properties to plasters,
mortars, and Portland cements, being quick-setting, stronger and more
resistant to erosion. These advantageous properties make them strong
candidates for other applications. In fact, one of these cements, the
magnesium oxychloride cement of Sorel (1867), is still used to surface walls
and floors on account of its marble-like appearance (Chapter 7).
In the 1870s more effective liquid cement-formers were found: ortho-
phosphoric acid and eugenol (Wilson, 1978). It was also found that an
aluminosilicate glass could replace zinc oxide, a discovery which led to the
first translucent cement. Thereafter the subject stagnated until the late
1960s when the polyelectrolyte cements were discovered by Smith (1968)
and Wilson & Kent (1971).
In recent years Sharp and his colleagues have developed the magnesium
phosphate cements - Sharp prefers the term magnesia phosphate cement
- as a material for the rapid repair of concrete runways and motorways
(Chapter 6). These applications exploit the rapid development of strength
in AB cements. This cement can also be used for flooring in refrigerated
stores where Portland cements do not set. Interestingly, this material
appears to have started life as an investment for the casting of dental alloys.
The glass polyalkenoate, a polyelectrolyte cement, of Wilson & Kent
(Chapter 5), was originally developed as a dental material but has since
found other applications. First it was used as a splint bandage material
possessing early high-strength and resistance to water. Currently, it is
being used, as a biocompatible bone cement, with a low exothermicity on
setting and the ability to adhere to bone, for the cementation of prostheses
(Jonck, Grobbelaar & Strating, 1989).
Outside thefieldof biomaterials it has been patented for use as a cement
for underwater pipelines, as a foundry sand and as a substitute for plaster
Introduction
in the slip casting of pottery. Quite often it appears as a substitute for
plaster of Paris, for it is stronger, less brittle and more resistant to water.
There are other possibilities. Its translucent nature suggests that it could be
used for the production of porcelain-like ceramics at room temperature.
Phosphate and polyelectrolyte AB cements are resistant to attack by
boiling water, steam and mild acids and this suggests that they could be
employed in technologies where these properties are important.
The ability of the polyelectrolyte-based AB cements (Chapter 5) to bond
to a variety of substrates, combined with their rapid development of
strength - they can become load-bearing within minutes of preparation -
suggests that they have applications as rapid-repair and handyman
materials.
A current area of interest is the use of AB cements as devices for the
controlled release of biologically active species (Allen et aL, 1984). AB
cements can be formulated to be degradable and to release bioactive
elements when placed in appropriate environments. These elements can be
incorporated into the cement matrix as either the cation or the anion
cement former. Special copper/cobalt phosphates/selenates have been
prepared which, when placed as boluses in the rumens of cattle and sheep,
have the ability to decompose and release the essential trace elements
copper, cobalt and selenium in a sustained fashion over many months
(Chapter 6). Although practical examples are confined to phosphate
cements, others are known which are based on a variety of anions:
polyacrylate (Chapter 5), oxychlorides and oxysulphates (Chapter 7) and
a variety of organic chelating anions (Chapter 9). The number of cements
available for this purpose is very great.
A recent development has been the incorporation of a bioactive organic
component into the AB cement during preparation. Since AB cements are
prepared at room temperature, this can be done without causing
degradation of the organic compound. In this case, the AB cement may
merely act as a carrier for the sustained release of the added bioactive
compound.
Another development has been the advent of the dual-cure resin cements.
These are hybrids of glass polyalkenoate cements and methacrylates that
set both by an acid-base cementation reaction and by vinyl polymerization
(which may be initiated by light-curing). In these materials, the solvent is
not water but a mixture of water and hydroxyethylmethacrylate which is
capable of taking dimethacrylates and poly(acrylic acid)-containing vinyl
groups into solution. In the absence of light these materials set slowly and
Introduction
have extended working times, but they set in seconds when illuminated
with an intense beam of visible light. These hybrids are in their infancy but
have created great interest.
From this account we are to expect diversification of these AB cements
both for biomedical and for industrial usages. There should be further
developments of the glass polyalkenoate cements both as bone substitutes
and as bioadhesives. We also expect more types of AB cements to be
formulated as devices for the sustained release of bioactive species. These
materials would have applications in agriculture, horticulture, animal
husbandry and human health care. In industrialfieldswe expect that there
will be continued interest in developing AB cements as materials for the
rapid repair of constructural concrete, as materials for the surfacing of
floors and walls, and as adhesives and lutes for cementation in aqueous
environments. The hybrid light-cured cements also appear to be a
promising new line of development which may give us entirely novel classes
of materials.

References
Allen, W. M., Sansom, B. F., Wilson, A. D., Prosser, H. J. & Groffman, D. M.
(1984). Release cements. British Patent GB 2,123,693 B.
Jonck, L. M., Grobbelaar, C. J. & Strating, H. (1989). The biocompatibility of
glass-ionomer cement in joint replacement: bulk testing. Clinical Materials, 4,
85-107.
Mellor, J. W. (1929). A Comprehensive Treatise on Inorganic and Theoretical
Chemistry, vol. IV, p. 546. London: Longman.
Sorel, S. (1855). Procede pour la formation d'un ciment tres-solide par 1'action
d'un chlorure sur l'oxyde de zinc. Comptes rendus hebdomadaires des seances
de T Academie des sciences, 41, 784-5.
Sorel, S. (1867). On a new magnesium cement. Comptes rendus hebdomadaires
des seances de VAcademie des sciences, 65, 102—4.
Wilson, A. D. (1978). The chemistry of dental cements. Chemical Society
Review, 7, 265-96.
2 Theory of acid-base cements

2.1 General
From the chemical point of view AB cements occupy a place in the vast
range of acid-base phenomena which occur throughout both inorganic
and organic chemistry. Like Portland cement they are prepared by mixing
a powder with a liquid. However, this liquid is not water but an acid, while
the powder, a metal oxide or silicate, is a base. Not surprisingly, the
cement-forming reaction between them is extremely rapid and a hardened
mass is formed within minutes of mixing.
AB cements may be represented by the defining equation
Base + Acid = Salt + Water
(powder) (liquid) (cement matrix)
The product of the reaction, the binding agent, is a complex salt, and
powder in excess of that required for the reaction acts as the filler. Each
cement system is a particular combination of acid and base. The number of
potential cement systems is considerable since it is a permutation of all
possible combinations of suitable acids and bases.
Cement-forming liquids are strongly hydrogen-bonded and viscous.
According to Wilson (1968), they must (1) have sufficient acidity to
decompose the basic powder and liberate cement-forming cations, (2)
contain an acid anion which forms stable complexes with these cations and
(3) act as a medium for the reaction and (4) solvate the reaction products.
Generally, cement-forming liquids are aqueous solutions of inorganic or
organic acids. These acids include phosphoric acid, multifunctional
carboxylic acids, phenolic bodies and certain metal halides and sulphates
(Table 2.1). There are also non-aqueous cement-forming liquids which are
multidentate acids with the ability to form complexes.
Potential cement-forming bases are oxides and hydroxides of di- and
Theory of acid—base cements

Table 2.1. Examples of acids used for cement formation

Protonic acids Aprotic acids


(used in aqueous solution) (used in aqueous solution)
Phosphoric acid Magnesium chloride
Poly(acrylic acid) Zinc chloride
Malic acid Copper(II) chloride
Tricarballylic acid Cobalt(II) chloride
Pyruvic acid Magnesium sulphate
Tartaric acid Zinc sulphate
Mellitic acid Copper(II) sulphate
Gallic acid Cobalt(II) sulphate
Tannic acid Magnesium selenate
Zinc selenate
Protonic acids Copper(II) selenate
(liquid non-aqueous) Cobalt(II) selenate

Eugenol
2-ethoxybenzoic acid

Table 2.2. Examples of bases used for cement formation

Copper(II) oxide
Zinc(II) oxide
Magnesium oxide
Cobalt(II) hydroxide
Cobalt(II) carbonate
Calcium aluminosilicate glasses
Gelatinizing minerals

trivalent metals, silicate minerals and aluminosilicate glasses (Table 2.2).


All cement-forming bases must be capable of releasing cations into acid
solution. The best oxides for cement formation are amphoteric (Kingery,
1950a,b) and the most versatile cement former is zinc oxide, which can
react with a wide range of aqueous solutions of acids, both inorganic and
organic, and liquid organic chelating agents. Gelatinizing minerals, that is
minerals that are decomposed by acids, can act as cement formers, as can
the acid-decomposable aluminosilicate glasses.
In this chapter the nature of the cementitious bond and the acid-base
reaction will be discussed.
The formation of cements

2.2 The formation of cements


2.2.1 Classification
Before proceeding further it is well to consider the term cement, for its
definition can be the source of some confusion. Both the Oxford English
Dictionary and Webster give two alternative definitions. One defines a
cement as a paste, prepared by mixing a powder with water, that sets to a
hard mass. In the other a cement is described as a bonding agent. These two
definitions are quite different. The first leads to a classification of cements
in terms of the setting process, while the second lays emphasis on a
property. In this book the term cement follows the sense of thefirstof these
definitions.
Cements can be classified into three broad categories:
(1) Hydraulic cements. These cements are formed from two con-
stituents one of which is water. Setting comprises a hydration and
precipitation process. Into this category fall Portland cement and
plaster of Paris.
(2) Condensation cements. Here, cement formation involves a loss of
water and the condensation of two hydroxyl groups to form a
bridging oxygen:
R-OH + HO-R = R-O-R + H2O
One example is silicate cement where orthosilicic acid, chemically
generated in solution, condenses to form a silicic acid gel. Another
is refractory cement where a cementitious product is formed by
the heat treatment of an acid orthophosphate, a process which
again involves condensation to form a polyphosphate.
(3) Acid-base cements. Cement formation involves both acid-base
and hydration reactions (Wilson, Paddon & Crisp, 1979). These
cements form the subject of this book.
This classification differs from that given by Wygant (1958), who
subdivides cements into hydraulic, precipitation and reaction cements. The
advantage of the present classification is that it clearly differentiates
phosphate cements formed by condensation from those formed by an
acid-base reaction (Kingery, 1950a). Wygant includes these in the same
category, which can be confusing. Moreover, he puts silicate cements and
the heat-treated acid phosphate cements into separate categories, although
both are condensation cements.
Theory of acid—base cements

2.2.2 Requirements for cementitious bonding


The essential property of a cementitious material is that it is cohesive.
Cohesion is characteristic of a continuous structure, which in the case of a
cement implies an isotropic three-dimensional network. Moreover, the
network bonds must be attributed to attractions on the molecular level.
Increasingly, recent research tends to show that cements are not bonded by
interlocking crystallites and that the formation of crystallites is incidental
(Steinke et al., 1988; Crisp et al., 1978). The reason is that it is difficult to
form rapidly a mass which is both cohesive and highly ordered.
Cement formation requires a continuous structure to be formed in situ
from a large number of nuclei. Moreover, this structure must be maintained
despite changes in the character of the bonds. These criteria are, obviously,
more easily satisfied by aflexiblerandom structure than by one which is
highly-ordered and rigid. Crystallinity implies well-satisfied and rigidly-
directed chemical bonds, exact stoichiometry and a highly ordered
structure. So unless crystal growth is very slow a continuous molecular
structure cannot be formed.
In random structures, stoichiometry need not be exact and adventitious
ions can be incorporated without causing disruption. Bonds are not highly
directed, and neighbouring regions of precipitation, formed around
different nuclei, can be accommodated within the structure. Continuous
networks can be formed rapidly. Thus, random structures are conducive to
cement formation and, in fact, most AB cements are essentially amorph-
ous. Indeed, it often appears that the development of crystallinity is
detrimental to cement formation.
The matrices of AB cements are gel-like, but these gels differ from the
tobermorite gel of Portland cement. In AB cements, setting is the result of
gelation by salt formation, and the cations, which cause gelation, are
extracted from an oxide or silicate by a polyacid solution. The conversion
of the sol to a gel is rapid and the cements set in 3 to 5 minutes. Two basic
processes are involved in cement formation: the release of cations from the
oxide or silicate and their interaction with polyacid. This interaction
involves ion binding and changes in the hydration state which are
associated with gelation and structure formation (Section 4.3). Thus, there
are two reaction rates to be considered: the rate of release of cations and
the rate of structure formation. These two reaction rates must be balanced.
If the rate of release of cations is too fast a non-coherent precipitate of
crystallites is formed. If too slow the gel formed will lack strength.
The formation of cements
During cement formation, domains are formed about numerous nuclei
and there must be bonding between the domains as well as within them. In
AB cements bonding within the domains is mainly ionic, with a degree of
covalency. The attractive forces between domains are those of a colloidal
type. In random structures, residual forcefieldsexist which act in a similar
fashion to polar forces and serve to bond domains. These forces must
include hydrogen bonds, for the addition offluorideions always enhances
cement strength and the fluoride-hydrogen bond is a strong one.
The structures of cement gels bear some relationship to the structure of
glasses. Spatially, the O2~ ion is dominant. The matrices are based on a
coordinated polyhedron of oxygen ions about a central glass-forming
cation (Pauling, 1945). In effect, these are anionic complexes where the
cations are small, highly charged, and capable of coordinating with oxygen
or hydroxyl ions. Examples of these polyhedra are [SiOJ, [POJ and [A1OJ.
Thus, wefindthat there are silicate, phosphate and aluminosilicate glasses
and gels.
There are, however, differences which are best illustrated by reference to
the simple example of silica glass and silica gel. In silica glass, Si4+ is four-
coordinate and the polymeric links are of the bridging type:

-«>Si—O—Si<^-

In aqueous solution, coordination increases to 6, Si-OH links are possible


as well as Si-O-Si, and H2O is a possible ligand. In silica cements the
condensation of silicic acid, Si(OH)4, to SiO2 is only partial. Silica gel
therefore contains both bridging oxygen and non-bridging hydroxyl
linkages. Again, in contrast to the situation in glasses the possibility of
hydrogen bond formation will also exist.
In AB cements the gel-forming cations are frequently Zn2+, Mg2+, Ca2+
or Al3+. As Kingery (1950b) has pointed out, it is the amphoteric cations,
for example Zn2+ and Al3+, that possess the most favourable cement-
forming properties. Their oxides are capable of glass formation, not by
themselves, but in conjunction with other glass formers. Kingery also
indicated that weakly basic cations, for example Mg2+, are less effective,
and more strongly basic cations, for example Ca2+, even less effective.
The nature of the association between cement-forming cation and anion
is important. As we shall see from theoretical considerations of the nature
of acids and bases in section 2.3, these bonds are not completely ionic in
character. Also while cement-forming cations are predominantly a-
Theory of acid-base cements
acceptors and the anions cr-donors, both have weak ^-capabilities also.
This topic is treated in more detail in the next section. Complex formation
is clearly important and this view is supported by the anomaly that B2O3
forms cements with acids, not as a result of salt formation, but because of
complex formation (Chapters 5 and 8).
A final point needs to be made. Theory has indicated that AB cements
should be amorphous. However, a degree of crystallization does sometimes
occur, its extent varying from cement to cement, and this often misled early
workers in the field who used X-ray diffraction as a principal method of
study. Although this technique readily identifies crystalline phases, it
cannot by its nature detect amorphous material, which may form the bulk
of the matrix. Thus, in early work too much emphasis was given to
crystalline structures and too little to amorphous ones. As we shall see, the
formation of crystallites, far from being evidence of cement formation, is
often the reverse, complete crystallinity being associated with a non-
cementitious product of an acid-base reaction.

2.2.3 Gelation
The formation of AB cements is an example of gelation, and the matrices
may be regarded as salt-like hydrogels. They are rigid and glass-like. A gel
has been defined by Bungenberg de Jong (1949) as 'a system of solid
character, in which the colloidal particles somehow constitute a coherent
structure'. A more exact definition is not possible, for gels are easier to
recognize than define; they include a diversity of substances. Coherence of
structure appears, however, to be a universal criterion for gels.
Flory (1974) classified gels into four types on the basis of their
structures:

(1) Well-ordered lamellar structures. The lamellae are arranged in


parallel, giving rise to long-range order. Examples are soaps,
phospholipids and clays.
(2) Covalent polymeric networks which are completely disordered.
Continuity of structure is provided by an irregular three-
dimensional network of covalent links, some of which are
crosslinks. The network is uninterrupted and has an infinite
molecular weight. Examples are vulcanized rubbers, condensation
polymers, vinyl-divinyl copolymers, alkyd and phenolic resins.

10
The formation of cemen ts
(3) Polymer networks formed through physical aggregation; these
are predominantly disordered, but have regions of local order.
Linear structures of finite length are connected by multiple-
stranded helices, which may be crystalline. Examples are gelatin
and sodium alginate gels.
(4) Particulate, disordered structures. These include flocculent pre-
cipitates where particles generally consist of fibres in brush-heap
disarray or connected in irregular networks.
Since the matrices of AB cements bear some similarity to alginate gels
they most probably fall into type 3.
The classical theory of gelation, due to Flory (1953, 1974), sees gelation
as the result of the formation of an infinite three-dimensional network.
According to Flory, the theory can be applied without ambiguity to the
type 2 (covalent) gels and is also applicable to type 3 gels. The conditions
for the formation of such an infinite network are critical. Flory conceives
the growth of a random network as a sequential condensation process
between difunctional and multifunctional units involving a branching
process. During growth, the probability of branching (a) at each potential
branching point has to reach a critical value (ac) for an infinite network to
be formed. In the case of condensation between di- and trifunctional
groups, the probability has to be more than 50 % for an infinite network to
be formed. If it is 50 % or less then an infinite network is not formed. This
theory explains why gelation occurs suddenly.
In general, the critical value for a, ac, is given by the expression

where / i s the degree of functionality of the multifunctional group.


The most investigated examples are to be found in the precipitation of
polyelectrolytes by metal ions. Here, networks are formed by the random
crosslinking of linear polymer chains, and the theory requires some
modification. The condition for the formation of an infinite network is
that, on average, there must be more than two crosslinks per chain. Thus,
the greater the length of a polymer chain the fewer crosslinks in the system
as a whole are required.

11
Theory of acid-base cements

2.3 Acid-base concepts


2.3.1 General
The cement-forming reaction is a special case of an acid-base reaction so
that concepts of acid, base and salt are central to the topic. In AB cement
theory, we are concerned with the nature of the acid-base reaction and
how the acidity and basicity of the reactants are affected by their
constitution. Thus, it is appropriate at this stage to discuss the various
definitions and theories available.
Although acids and bases have been recognized since antiquity, our
concepts of them are still the subject of debate and development (Walden,
1929; Hall, 1940; Bell, 1947, 1973; Luder, 1948; Kolthoff, 1944; Bjerrum,
1951; Day & Selbin, 1969; Jensen, 1978; Finston & Rychtman, 1982). The
history of these concepts is a long one and can be seen as a prolonged and
continuous refinement of inexact and commonsense notions into precise
scientific theories. It has been a long and difficult journey and one that is by
no means ended.
There are various definitions of acids and bases, and in discussing them
it should be emphasized that the question is not one of validity but one of
utility. Indeed, the problem of validity does not arise because of the
fundamental nature of a definition. The problem is entirely one of choosing
a definition which is of greatest use in the study of a particular field of
acid-base chemistry. One point that needs to be borne in mind is that a
concept of acids and bases is required that is neither too general nor too
restrictive for the particular field of study.

2.3.2 History of acid-base concepts


From early times acids were recognized by their properties, such as
sourness and ability to dissolve substances, often with effervescence. The
story of Cleopatra's draught of a pearl dissolved in vinegar illustrates this
point (Pattison Muir, 1883). Vinegar, known to the Greeks and Romans,
was associated with the concept of acidity and gives its name to the term
acid which comes from the Latin acetum. Boyle (1661) observed that acids
dissolve many substances, precipitate sulphur from alkaline solution,
change blue plant dyes to red and lose these properties on contact with
alkalis.
It also has been known since antiquity that aqueous extracts of the ash

12
Acid-base concepts
of certain plants have distinctive properties: slipperiness, cleansing power
in the removal of fats, oils and dirt from fabrics, and the ability to affect
plant colours (Day & Selbin, 1969; Pattison Muir, 1883). These substances
were called alkalis, a name which comes from the Arabic for plant ash, al
halja (Finston & Rychtman, 1982) or algali. The term alkali applies only
to the hydroxides and carbonates of sodium and potassium, and it was
Rouelle in 1744 who extended the concept to include the alkaline earth
analogues and used the term base to categorize them (Walden, 1929; Day
& Selbin, 1969).
Salt formation as a criterion for an acid-base interaction has a long
history (Walden, 1929). Rudolph Glauber in 1648 stated that acids and
alkalis were opposed to each other and that salts were composed of these
two components. Otto Tachenius in 1666 considered that all salts could be
broken into an acid and an alkali. Boyle (1661) and the founder of the
phlogistic theory, Stahl, observed that when an acid reacts with an alkali
the properties of both disappear and a new substance, a salt, is produced
with a new set of properties. Rouelle in 1744 and 1754 and William Lewis
in 1746 clearly defined a salt as a substance that is formed by the union of
an acid and a base.
It can be seen that these definitions are derived from experimental
observation and are no more than classifications based on a set of
properties shared by a group of substances. They are scientifically
inadequate for the interpretation of results, which requires a definition
based on concepts. Historically, the attempt to provide a model rather than
a classification comes in the form of a search for underlying universal
principles. It seems that the alchemists recognized vague principles of
acidity and alkalinity, and in the 17th century the iatrochemists made these
the basis of chemical medicine. Disease was attributed to a predominance
of one or other of these principles (Pattison Muir, 1883).
Boyle (1661) attempted to provide a more definite concept and attributed
the sour taste of acids to sharp-edged acid particles. Lemery, another
supporter of the corpuscular theory of chemistry, had similar views and
considered that acid-base reactions were the result of the penetration of
sharp acid particles into porous bases (Walden, 1929; Finston &
Rychtman, 1982). However, the first widely accepted theory was that of
Lavoisier who in 1777 pronounced that oxygen was the universal acidifying
principle (Crosland, 1973; Walden, 1929; Day & Selbin, 1969; Finston &
Rychtman, 1982). An acid was defined as a compound of oxygen with a
non-metal.

13
Theory of acid-base cements
After this theory was disproved, other acidifying principles were
proposed. The most significant was the recognition, by Davy & Dulong
early in the 19th century, of hydrogen as the acidifying principle (Walden,
1929; Finston & Rychtman, 1982). During this period no such search was
made for a basic principle. Bases were merely regarded as a motley
collection of antiacids with little in common apart from the ability to react
with acids.
The first substantial constitutive concept of acid and bases came only in
1887 when Arrhenius applied the theory of electrolytic dissociation to
acids and bases. An acid was defined as a substance that dissociated to
hydrogen ions and anions in water (Day & Selbin, 1969). For thefirsttime,
a base was defined in terms other than that of an antiacid and was regarded
as a substance that dissociated in water into hydroxyl ions and cations. The
reaction between an acid and a base was simply the combination of
hydrogen and hydroxyl ions to form water.
This theory was a milestone in the development of acid-base concepts:
it was the first to define acids and bases in terms other than that of a
reaction between them and the first to give quantitative descriptions.
However, the theory of Arrhenius is far more narrow than both its
predecessors and its successors and, indeed, it is the most restrictive of all
acid-base theories.
Since Arrhenius, definitions have extended the scope of what we mean
by acids and bases. These theories include the proton transfer definition of
Bronsted-Lowry (Bronsted, 1923; Lowry, 1923a,b), the solvent system
concept (Day & Selbin, 1969), the Lux-Flood theory for oxide melts, the
electron pair donor and acceptor definition of Lewis (1923, 1938) and the
broad theory of Usanovich (1939). These theories are described in more
detail below.

2.3.3 Acid—base concepts in AB cement chemistry


We now review the various concepts of acids and bases in order to see how
appropriate and useful they are in the field of AB cements.

The definition of Arrhenius


This definition of acids and bases is of restricted application. The reaction
between acids and bases is seen as the combination of hydrogen and
hydroxyl ions in aqueous solution to form water.

14
Acid-base concepts
An acid is defined as a species that dissociates in aqueous solution to give
hydrogen ions and onions, and a base as a species that dissociates in aqueous
solution to give hydroxyl ions and cations.

Thus, acids and bases are defined as aqueous solutions of substances and
not as the substances themselves. It follows that ionization is a necessary
characteristic of Arrhenius acids and bases. Another restriction of this
definition is that acid-base behaviour is not recognized in non-aqueous
solution.
The Arrhenius definition is not suitable for AB cements for several
reasons. It cannot be applied to zinc oxide eugenol cements, for these are
non-aqueous, nor to the metal oxychloride and oxysulphate cements,
where the acid component is not a protonic acid. Indeed, the theory is,
strictly speaking, not applicable at all to AB cements where the base is not
a water-soluble hydroxide but either an insoluble oxide or a silicate.

The protonic Bronsted-Lowry theory


The theory of Bronsted (1923) and Lowry (1923a, b) is of more general
applicability to AB cements. Their definition of an acid as' a substance that
gives up a proton' differs little from that of Arrhenius. However, the same
is not true of their definition of a base as' a substance capable of accepting
protons' which is far wider than that of Arrhenius, which is limited to
hydroxides yielding hydroxide ions in aqueous solution. These concepts of
Bronsted and Lowry can be defined by the simple equation (Finston &
Rychtman, 1982):
Acid = Base + H+ [2.2]
Thus, the relationship between acid and base is a reciprocal one and an
acid-base reaction involves the transfer of a proton. This concept is not
restricted to aqueous solutions and it discards Arrhenius' prerequisite of
ionization.
This concept covers most situations in the theory of AB cements.
Cements based on aqueous solutions of phosphoric acid and poly(acrylic
acid), and non-aqueous cements based on eugenol, alike fall within this
definition. However, the theory does not, unfortunately, recognize salt
formation as a criterion of an acid-base reaction, and the matrices of AB
cements are conveniently described as salts. It is also uncertain whether it
covers the metal oxide/metal halide or sulphate cements. Bare cations are
not recognized as acids in the Bronsted-Lowry theory, but hydrated

15
Theory of acid-base cements
cations are. Thus, in the case of the group III elements, the octahedral
[M(H2O)6]3+ aquo ions are quite acidic (Cotton & Wilkinson, 1966):
[M(H2O)6]3+ = [M(H2O)5 (OH)]2+ + H+ [2.3]
However, although both zinc and magnesium ions, the cations of the oxy-
cements, are hydrated as [M(H2O)6]2+ ions, these hydrated ions hydrolyse
only slightly (Baes & Mesmer, 1976). Thus, in magnesium chloride
solutions the aquo ions, in contrast to beryllium aquo ions, are not
perceptibly acidic. So there must be some doubt as to whether these
hydrated ions can be regarded as protonic acids. But for this, the
Bronsted-Lowry theory would almost exactly define AB cements.
Aluminosilicate glasses are used in certain AB cement formulations,
and the acid-base balance in them is important. The Bronsted-Lowry
theory cannot be applied to these aluminosilicate glasses; it does not
recognize silica as an acid, because silica is an aprotic acid. However, for
most purposes the Bronsted-Lowry theory is a suitable conceptual
framework although not of universal application in AB cement theory.

The solvent system theory


Although the protonic theory is not confined to aqueous solutions, it does
not cover aprotic solvents. The solvent system theory predates that of
Bronsted-Lowry and represents an extension of the Arrhenius theory to
solvents other than water. It may be represented by the defining equation:
Acid + Base = Salt + Solvent [2.4]
This theory is associated in its early protonic form with Franklin (1905,
1924). Later it was extended by Germann (1925a,b) and then by Cady &
Elsey (1922,1928) to a more general form to include aprotic solvents. Cady
& Elsey describe an acid as a solute that, either by direct dissociation or by
reaction with an ionizing solvent, increases the concentration of the solvent
cation. In a similar fashion, a base increases the concentration of the
solvent anion. Cady & Elsey, in order to emphasize the importance of the
solvent, modified the above defining equation to:
Acidic solution + Basic solution = Salt + Solvent [2.5]
Thus, acids and bases do not react directly but as solvent cations and
anions. Since emphasis is placed upon ionization interactions, inherent
acidity and basicity is neglected, as are interactions in the non-ionic state.
The theory is a simple extension of the Arrhenius theory and suffers from

16
Acid-base concepts
the same drawbacks. The definition cannot be applied directly to the
reaction between a basic solid and acidic liquid characteristic of AB
cements.

The Lux-Flood theory


The Lux-Flood theory relates to oxide melts. Geologists have often used
acid-base concepts for the empirical classification of igneous silicate rocks
(Read, 1948). Silica is implicitly assumed to be responsible for acidity, and
the silica content of a rock is used as a measure of its acid-base balance:
Rock type Silica content (SiO2) %
Acid >66
Intermediate 52-66
Basic 45-52
Ultra-basic <45
Lux (1939) developed an acid-base theory for oxide melts where the oxide
ion plays an analogous but opposite role to that of the hydrogen ion in the
Bronsted theory. A base is an oxide donor and an acid is an oxide acceptor
(Lux, 1939; Flood & Forland, 1947a,b; Flood, Forland & Roald, 1947):
Base = Acid+ O2~ [2.6]
Thus an acid-base reaction involves the transfer of an oxide ion (compared
with the transfer of a proton in the Bronsted theory) and the theory is
particularly applicable in considering acid-base relationships in oxide,
silicate and aluminosilicate glasses. However, we shall find that it is
subsumed within the Lewis definition.

The Lewis theory


This theory was advanced by G. N. Lewis (1916, 1923, 1938) as a more
general concept. In his classic monograph of 1923 he considered and
rejected both the protonic and solvent system theories as too restrictive. An
acid-base reaction in the Lewis sense means the completion of the stable
electronic configuration of the acceptor atom of the acid by an electron
pair from the base. Thus:
A base has the ability to donate a pair of electrons and an acid the ability to
accept a pair of electrons to form a covalent bond. The product of a Lewis
acid—base reaction may be called an adduct, a coordination compound or a
coordination complex (Vander Werf 1961). Neither salt nor conjugate
acid—base formation is a requirement.

17
Theory of acid-base cements
Although Lewis and Bronsted bases comprise the same species, the same is
not true of their acids. Lewis acids include bare metal cations, while
Bronsted-Lowry acids do not. Also, Bell (1973) and Day & Selbin (1969)
have pointed out that Bronsted or protonic acids fit awkwardly into the
Lewis definition. Protonic acids cannot accept an electron pair as is
required in the Lewis definition, and a typical Lewis protonic acid appears
to be an adduct between a base and the acid H+ (Luder, 1940; Kolthoff,
1944). Thus, a protonic acid can only be regarded as a Lewis acid in the
sense that its reaction with a base involves the transient formation of an
unstable hydrogen bond adduct. For this reason, advocates of the Lewis
theory have sometimes termed protonic acids secondary acids (Bell, 1973).
This is an unfortunate term for the traditional acids.
Lewis (1938) was not content with a purely conceptual view of acids and
bases, for he also listed certain phenomenological criteria for an acid-base
reaction. The process of neutralization is a rapid one, an acid or base
displaces a weaker acid or base from its compounds, acids and bases may
be titrated against each other using coloured indicators, and both acids and
bases have catalytic effects.
The Lewis definition covers all AB cements, including the metal
oxide/metal oxysalt systems, because the theory recognizes bare cations as
aprotic acids. It is also particularly appropriate to the chelate cements,
where it is more natural to regard the product of the reaction as a
coordination complex rather than a salt. Its disadvantages are that the
definition is really too broad and that despite this it accommodates
protonic acids only with difficulty.

The Usanovich theory


The Usanovich theory is the most general of all acid-base theories.
According to Usanovich (1939) any process leading to the formation of a
salt is an acid-base reaction. The so-called' positive-negative' definition of
Usanovich runs as follows.

An acid is a species capable of yielding cations, combining with onions or


electrons, or neutralizing a base. Likewise a base is a species capable of
yielding anions or electrons, combining with cations, or neutralizing an acid.

When developed, this theory proved to be more general than the theory of
Lewis, for it includes all the above acid-base definitions and also includes
oxidation-reduction reactions.

18
Acid—base concepts

It is better than the Lewis theory for describing acid-base cements, for
it avoids the awkwardness that the Lewis definition has with protonic
acids. However, as Day & Selbin (1969) have observed, the generality of
the theory is such that it includes nearly all chemical reactions, so that
acid-base reactions could simply be termed 'chemical reactions'.

2.3.4 Relevance of acid-base theories to AB cements


The various acid-base definitions are summarized in the Venn diagram
(Fig. 2.1). From this it can be seen that the Usanovich definition subsumes
the Lewis definition, which in turn subsumes all other definitions (i.e.
Arrhenius, Bronsted-Lowry, Germann-Cady-Elsey, Lux-Flood).
Also shown is how the topic of AB cements relates to these definitions.
An ideal definition for a subject should be one that exactlyfitsit. It should
cover all aspects of the subject while excluding all extraneous topics. Thus,
a theory should be neither too restrictive nor too general. The Arrhenius
and Germann-Cady-Elsey definitions do not relate to the subject at all as

USANOVICH

LEWIS
E l e c t r o n - p a i r acceptor

BR0NSTED ARRHENIUS GERMANN


proton-donor proton-donor sol vent-cation
any solvent in water donor

Figure 2.1 Venn diagram showing the relationship between the various definitions of acids
and bases.

19
Theory of acid-base cements

the basic component of an AB cement is a powdered solid. The


Bronsted-Lowry definition is not broad enough to include all AB cements
and excludes the concept of salt, which is unfortunate since the matrices of
AB cements are salts. Both the Lewis and Usanovich definitions cover all
aspects of AB cement theory at the cost of including topics not relevant to
this subject.
From this discussion it can be seen that there is no ideal acid-base theory
for AB cements and a pragmatic approach has to be adopted. Since the
matrix is a salt, an AB cement can be defined quite simply as the product
of the reaction of a powder and liquid component to yield a salt-like gel.
The Bronsted-Lowry theory suffices to define all the bases and the
protonic acids, and the Lewis theory to define the aprotic acids. The subject
of acid-base balance in aluminosilicate glasses is covered by the Lux-Flood
theory.

2.3.5 Acid-base strength


Ever since the formulation of the Bronsted-Lowry theory, efforts have
been made to develop a general approach to acid-base strength. The
influence of ionic charge and size of the central atom on acidity and
basicity is important. In 1926, Bronsted found that an increase in acidity
corresponded to an increase in positive charge or a decrease in negative
charge on an ion. Cartledge (1928a,b), against the background of the
protonic theory, proposed to correlate acidity or basicity with a function
he called ionic potential, by considering acids and bases to be hydroxides of
non-metals and metals, respectively. He defined ionic potential, (/>, as
</> = Z/r (2.1)
where Z is the charge on the central atom and r its ionic radius. Cartledge
(1928b) then used values of ^ 0 5 to define acidity and basicity of a species.

f5 value Acid-base status


>3-2 acidic
2-2-3-2 amphoteric
<2-2 basic
Thus, highly charged smaller cations are highly acidic. This point is
illustrated for the series Na + , Mg2+, Al3+, Si4+, P5+, S6+ and Cl7+ in Table
2.3a.
Note, however, that Zn(OH)2 is not classified as amphoteric as it should

20
Acid-base concepts
Table 2.3a. Effect of cation on acidity-basicity (Cartledge, 1982a,b)

Cation Ionic potential \ Species Acidity-basicity

Na + 102 NaOH Strong base


Ca 2+ 1-42 Ca(OH) 2 Weak base
Zn 2+ 1-64 Zn(OH) 2 Weak base
Mg 2+ 1-76 Mg(OH) 2 Weak base
Al 3+ 2-45 A1(OH)3 Amphoteric
Si4+ 313 Si(OH) 4 Weak acid
p5+ 3-83 H 3 PO 4 Intermediate acid
S 6 +
4-55 H 2 SO 4 Strong acid
Cl 7+ 5-20 HC1O4 Strongest acid

Table 2.3b. Effect of cation on acidity-basicity

Cation Ionization potential In Species Acidity-basicity

Na + 5-14 NaOH Strong base


Ca 2+ 11-87 Ca(OH) 2 Weak base
Mg 2+ 1503 Mg(OH) 2 Weak base
Cd 2+ 16-84 Cd(OH) 2 Amphoteric
Zn 2+ 17-96 Zn(OH) 2 Amphoteric
Cu 2+ 20-20 Cu(OH) 2 Amphoteric
Bi3+ 25-42 Bi(OH) 3 Amphoteric
Al 3+ 28-45 A1(OH)3 Amphoteric
Si4+ 45-14 Si(OH) 4 Weak acid
p5+ 6502 H 3 PO 4 Intermediate acid
s6+ 88-05 H 2 SO 4 Strong acid

In is the nth ionization potential.

be. Clearly, ionic potential alone is not a sufficient criterion for classi-
fication. As will be shown, unlike other cations in Table 2.3a which are
classified as hard acids, Zn2+ is an intermediate because of the presence of
d orbital electrons. The effect of d electrons in increasing the polarizing
power of the cations, because of ineffective screening, has been demon-
strated by Hodd & Reader (1976). They found that Cd2+ was a more
effective cement-former than Ca2+, because although both have a similar
ionic radius, Ca2+ has no d electrons. For these reasons, ionization
potential is a better criterion than ionic potential. As Table 2.3b shows,
Zn2+ is ranked correctly by this criterion and can be classified as

21
Theory of acid-base cements
amphoteric. Inspection of this table throws some light on the requirements
for cement formation. If judged by strength and hydrolytic stability of
cements formed with orthophosphoric acid, poly(acrylic acid) and poly-
(vinylphosphonic acid), the common cement-forming cations can be
ranked in the following order of decreasing effectiveness.
Al3+> Cu2+ > Zn2+ > Mg2+ > Ca2+
The first three form amphoteric oxides and are distinctly superior, as
cement-formers, to the latter two which form weakly basic oxides. Data
from Table 2.3b indicate that optimum cement formation occurs with
cations that have In values lying between 18 and 29.

2.3.6 Acid-base classification


The strength of a Lewis acid or base depends on the particular reaction,
and for this reason there is no absolute scale for the strengths of Lewis
acids and bases. However, certain qualitative features have been observed.
Ahrland, Chatt & Davies (1958) divided metal ions (which are Lewis
acids), on the basis of the stability of their complexes, into what they
termed class (a) and class (b) acceptors (Table 2.4). They stated that class
(a) acceptors form their most stable complexes with ligands of the lightest
member of a non-metal group. By contrast, class (b) acceptors form their
most stable complexes with heavier members of each group. Thus, complex
stability can be ranked according to the ligand as follows. For class (a)
acceptors O P S and for class (b) acceptors O <^ S. Class (a) metal ions are
small and non-polarizable, whereas class (b) metal ions are large and
polarizable. The class of a given element is not constant and depends on
oxidation state; class (a) character increases with increase in the positive
charge. Chatt (1958) considers that the important feature of class (b) acids
is the presence of loosely held outer d orbital electrons which can form n-
bonds to certain ligands. These ligands would contain empty d orbitals on
the basic atom; examples are P and As.
In the context of AB cements, Al3+, Mg2+, Ca2+ and Zn2+ are in class (a)
while Cu2+ is in the border region. Zn2+ contains a completed 3d shell and
forms stronger complexes with O than with S ligands, as do other class (a)
cations.

22
Table 2.4. Classification of acceptor atoms in their normal valent states (Ahrland, Chatt & Davies, 1958)

Class (a)
H
Li Be a/b border B C N 0 F

Class (a) a/b border


Na Mp Al Si P s Cl

K Ca I >c Ti V Cr Mn Fe Co Ni Cu Zn Ga Ge As Se Br

Rb Sr ^ft Zr Nb Mo Tc Ru Rh Pd Ag Cd In Sn Sb Te I

Cs Ba 1^a Hf Ta W Re Os Ir Pt Au Hg TI Pb Bi Po At

Class (a) a/b border Class (b) a/b border


Theory of acid-base cements

2.3.7 Hard and soft acids and bases (HSAB)


This concept of Chatt and his coworkers was developed further by Pearson
(1963, 1966, 1968a,b) in his theory of hard and soft acids and bases. Hard
acids correspond with class (a) acceptors and soft acids with class (b)
acceptors.

Hard acids prefer to react with hard bases and soft acids prefer to react
with soft bases.

Hard acids are characterized by small size, high positive charge and
absence of outer electrons which are easily excited to higher states; they are
thus of low polarizability. In this class are the common protonic acids, HA,
the hydrogen-bonding molecules in the Lewis scheme and Mg 2+ , which are
all acids of relevance to AB cements. The soft acids have low or zero
positive charge, large size and several easily excited outer electrons (often
d orbital electrons). These properties lead to high polarizability. The
division between these two classes is not sharp; amongst the intermediate
class are Zn 2+ and Cu 2+ .
Pearson (1966) defines a soft base as 'one in which the donor atom is of
high polarizability and low electronegativity and is easily oxidized or
associated with empty, low-lying orbitals'. A hard base has opposite
properties. 'The donor atom is of low polarizability and high electro-
negativity, is hard to reduce, and is associated with empty orbitals of high
energy.'
The underlying theory for hard-hard and soft-soft preferences is
obscure and no one factor is responsible (Pearson, 1966). Pearson (1963,
1968b) advanced several explanations. He stated that the ionic-covalent
theory provides the most obvious explanation. Hard acids are assumed to
bind bases primarily by ionic forces and soft acids by covalent bonds. High
positive charge and small size favour strong ionic bonding, and bases of
large negative charge and small size would be most strongly held. Soft acids
bind to bases by covalent bonding, and the atoms should be of similar size
and electronegativity for good bonding.
The classification of Lewis acids and bases relevant to AB cements is
shown below.

Hard acids: H A , H + , Ca 2 + , Mg 2 + , Al 3 + , Si 4+
Borderline acids: Zn 2+ , Cu2+
Hard bases: H 2 O, O H , F", POJ", SO2", RCOO"

24
Acid-base concepts

Table 2.5. YatsimirskiVs hardness indices {Yatsimirskii, 1970)

Base Indices Acid Indices

OH- 6-3 H+ 9-0


F- 1-7 In 3 + 1-2
HPOJ- 1-7 Cu 2 + 10
CH3COO- 0-8 Zn 2 + 0-2
sor
H2O
0-5
zero
La 3 + 01

Extension of HSAB theory


Yatsimirskii (1970) attempted to quantify HSAB theory and produced
hardness indices (S) for acids and bases. These indices were obtained by
plotting the logarithms of the equilibrium constants for the reactions of
bases with the proton (the hardest acid) against similar values for the
reactions with CH 3 Hg + (one of the softest acids). For acids, the hydroxyl
ion (the hardest base) and the chloride ion (a soft base) were chosen.
These S indices for cations and anions relevant to AB cements are shown
in Table 2.5. Bases which add on through F or O and do not form Tr-bonds
have similar hardness values; they are hard bases. Soft bases form dative
7r-bonds with many cations. They have high-energy-level occupied orbitals
with unshared electron pairs.
Yatsimirskii considered that the hard and soft classification was too
general and proposed instead a more detailed approach. He classified
Lewis acids and bases into six groups, based on the nature of the adduct
bonding.
Group (1) Cations and anions which are incapable of donor-acceptor
interactions. These are the large univalent ions. Bonding is purely
by Coulomb and Madelung electrostatic interactions. From the
Lewis point of view these are not acids or bases. They have no
cement-forming potential.
Group (2) Strong a-acceptor acids and donor bases. Included here are
protonic acids, which are relevant to AB cements. Their adducts
can only contain one coordinate bond.
Group (3) G- and n-acceptor acids and donor bases with o-interactions
predominating. In this group acceptors are capable of adding on
electron pairs of donors in both types of interactions. Includes
cations with stable closed electron shells: Al 3+ , Mg 2+ , Ca2+ and

25
Theory of acid-base cements

Zn 2+ . Donors are ligands coordinated through oxygen atoms or


fluoride ions: RCOO", PO*~, OH", F" and H 2 O. These acceptors
and donors are of relevance to AB cements.
Group (4) Strong a- and n-acceptor acids and donor bases. Bi3+, In 3+
and Sn2+ are of some relevance to AB cements.
Group (5) Acids that are o-acceptors but capable of n-donation in
backbonding. This group includes cations with mobile d electrons
e.g. Cuw+, Co w+ , Fe w+ .
Group (6) Bases that are a-donors but n-acceptors.

According to Yatsimirskii, group (2) and (3) species are equivalent to


Pearson's hard acids and bases, and group (4), (5) and (6) species
correspond to Pearson's soft acids and bases. This classification is of more
value than HSAB theory to our subject. It can be seen that all cement-
forming anions come from group (3) and cations from groups (3), (4) and
(5). Thus, the bonding in cement matrices formed from cation-anion
combinations is not purely a but contains some n character.

References
Ahrland, S., Chatt, J. & Davies, N. R. (1958). The relative affinities of ligand
atoms for acceptor molecules and ions. Quarterly Reviews, 12, 265-76.
Baes, C. F. & Mesmer, R. E. (1976). The Hydrolysis of Cations. New York:
John Wiley.
Bell, R. P. (1947). The use of the terms 'acid' and 'base'. Quarterly Reviews, 1,
113-25.
Bell, R. P. (1973). The Proton in Chemistry. Ithaca, New York: Cornell
University Press.
Bjerrum, J. (1951). Die Entwickhmgsgeschichte des Saure-Basenbegriffes und
iiber die ZweckmaBigkeit der Einfuhrung eines besonderen Antibasenbegriffes
neben dem Saurebegriff. Naturwissenschaften, 38, 461-4.
Boyle, R. (1661). The Sceptical Chymist. Everyman Library Edition, 1911.
Brensted, J. N. (1923). Einige Bemerkungen iiber den Begriff der Sauren und
Basen. Recueil des Travaux chimiques des Pays-Bas et de la Belgique, 42,
718-28.
Bronsted, J. N. (1926). The acid-base function of molecules and its dependency
on the electronic charge type. Journal of Physical Chemistry, 30, 777-90.
Bungenberg de Jong, H. G. (1949). In Kruyt, H. R. (ed.) Colloid Science II, p. 2.
Amsterdam: Elsevier Publishing Co. Inc.
Cady, H. P. & Elsey, H. M. (1922). A general conception of acids, bases and
salts. Science, 56, 27 (Lecture abstract).
Cady, H. P. & Elsey, H. M. (1928). A general definition of acids, bases and
salts. Journal of Chemical Education, 5, 1425-8.

26
References

Cartledge, G. H. (1928a). Studies on the periodic system. I. The ionic potential


as a periodic function. Journal of the American Chemical Society, 50, 2855-63.
Cartledge, G. H. (1928b). Studies on the periodic system. II. The ionic potential
and related properties. Journal of the American Chemical Society, 50, 2863-72.
Chatt, J. (1958). The stabilisation of low valent states of the transition metals.
Journal of Inorganic & Nuclear Chemistry, 8, 515-31.
Cotton, F. A. & Wilkinson, G. (1966). Advanced Inorganic Chemistry, 2nd edn.
New York, London & Sydney: Wiley Inter science.
Crisp, S., O'Neill, I. K., Prosser, H. J., Stuart, B. & Wilson, A. D. (1978).
Infrared spectroscopic studies on the development of crystallinity in dental
zinc phosphate cements. Journal of Dental Research, 57, 245-54.
Crosland, M. P. (1962). Historical Studies in the Language of Chemistry.
London: Heinemann.
Crosland, M. (1973). Lavoisier's theory of acidity. Isis, 64, 306-25.
Day, M. C. & Selbin, J. (1969). Theoretical Inorganic Chemistry. New York:
Reinhold.
Finston, H. L. & Rychtman, A. C. (1982). A New View of Current Acid-Base
Theories. New York: John Wiley & Sons.
Flood, H. & Forland, T. (1947a). The acidic and basic properties of oxides.
Ada Chemica Scandinavica, 1, 592—604.
Flood, H. & Forland, T. (1947b). The acidic and basic properties of oxides. II.
The thermal decomposition of pyrosulphates. Acta Chemica Scandinavica, 1,
781-9.
Flood, H., Forland, T. & Roald, B. (1947). The acidic and basic properties of
oxides. III. Relative acid-base strengths of some polyacids. Acta Chemica
Scandinavica, 1, 790-8.
Flory, P. J. (1953). Principles of Polymer Chemistry, Chapter 11. Ithaca, New
York: Cornell University Press.
Flory, P. J. (1974). Introductory lecture. In Gels and Gelling Processes. Faraday
Discussions of the Chemical Society, No. 57, pp. 7-18.
Franklin, E. C. (1905). Reactions in liquid ammonia. Journal of the American
Chemical Society, 27, 820-51.
Franklin, E. C. (1924). Systems of acids, bases and salts. Journal of the
American Chemical Society, 46, 2137-51.
Germann, A. F. O. (1925a). What is an acid? Science, 61, 71.
Germann, A. F. O. (1925b). A general theory of solvent systems. Journal of the
American Chemical Society, 47, 2461-8.
Hall, N. F. (1940). Systems of acids and bases. Journal of Chemical Education,
17, 124^8.
Hodd, K. A. & Reader, A. L. (1976). The formation and hydrolytic stability of
metal ion-polyacid gels. British Polymer Journal, 8, 131-9.
Jensen, W. B. (1978). The Lewis acid-base definitions: a status report. Chemical
Reviews, 78, 1-22.
Kingery, W. D. (1950a). Fundamental study of phosphate bonding in
refractories. I. Literature review. Journal of the American Ceramic Society, 33,
239-41.

27
Theory of acid-base cements

Kingery, W. D. (1950b). Fundamental study of phosphate bonding in


refractories. II. Cold setting properties. Journal of the American Ceramic
Society, 33, 242-7.
Kolthoff, I. M. (1944). The Lewis and Bronsted-Lowry definitions of acids and
bases. Journal of Physical Chemistry, 48, 51-7.
Lewis, G. N. (1916). The atom and the molecule. Journal of the American
Chemical Society, 38, 762-85.
Lewis, G. N. (1923). Valence and the Structure of Atoms and Molecules. New
York: Chemical Catalog Co.
Lewis, G. N. (1938). Acids and bases. Journal of the Franklin Institute, 226,
293-337.
Lowry, T. M. (1923a). The uniqueness of hydrogen. Chemistry & Industry, 42,
43.
Lowry, T. M. (1923b). Co-ordination and acidity. Chemistry & Industry, 42,
1048-52.
Luder, W. F. (1940). The electronic theory of acids and bases. Chemical
Reviews, 27, 547-83.
Luder, W. F. (1948). Contemporary acid-base theory. Journal of Chemical
Education, 25, 555-8.
Lux, H. (1939). 'Sauren' und 'Basen' im Schelzfluss: Die Bestimmung der
Sauerstoffionen-Konzentration. Zeitschrift fur Elektrochemie, 45, 303-9.
Pattison Muir, M. M. (1883). Heroes of Science-Chemists, Chapter IV, pp.
171-89. London: Society for Promoting Christian Knowledge.
Pauling, L. (1945). The Nature of the Chemical Bond. Ithaca, New York:
Cornell University Press.
Pearson, R. G. (1963). Hard and soft acids and bases. Journal of the American
Chemical Society, 85, 3533-9.
Pearson, R. G. (1966). Acids and bases. Science, 151, 172-7.
Pearson, R. G. (1968a). Hard and soft acids and bases, HSAB. Part I.
Fundamental principles. Journal of Chemical Education, 45, 581-7.
Pearson, R. G. (1968b). Hard and soft acids and bases, HSAB. Part II.
Underlying theories. Journal of Chemical Education, 45, 643-8.
Read, H. H. (1948). Rutle/s Elements of Mineralogy, 24th edn. London:
Thomas Murby & Co.
Smith, D. C. (1968). A new dental cement. British Dental Journal, 125, 381-4.
Steinke, R., Newcomer, P., Komarneni, S. & Roy, R. (1988). Dental cements:
investigation of chemical bonding. Materials Research Bulletin, 23, 13-22.
Usanovich, M. I. (1939). On acids and bases. Journal of General Chemistry
(USSR), 9, 182-92.
Vander Werf, A. (1961). Acids, Bases, and the Chemistry of the Covalent Bond.
New York: Reinhold.
Walden, P. (1929). Salts, Acids and Bases: Electrolytes, Stereochemistry. New
York: McGraw-Hill.
Wilson, A. D. (1968). Dental silicate cements: VII. Alternative liquid cement
formers. Journal of Dental Research, 47, 1133-6.
Wilson, A. D. & Kent, B. E. (1971). The glass-ionomer cement: a new

28
References

translucent cement for dentistry. Journal of Applied Chemistry and


Biotechnology, 21, 313.
Wilson, A. D., Paddon, J. M. & Crisp, S. (1979). The hydration of dental
cements. Journal of Dental Research, 58, 1065-71.
Wygant, J. F. (1958). Cementitious bonding in ceramic fabrication. In Kingery,
W. D. (ed.) Ceramic Fabrication Processes, pp. 171-88. New York: John
Wiley & Sons.
Yatsimirskii, K. (1970). Acid-base and donor-acceptor properties of ions and
molecules. Theoretical and Experimental Chemistry (USSR), 6, 376-80.

29
3 Water and acid-base cements

3.1 Introduction
The setting reaction for the great majority of acid-base cements takes place
in water. (The exceptions based on o-phenols are described in Chapter 9.)
This reaction does not usually proceed with formation of a precipitate but
rather yields a substance which entrains all of the water used to prepare the
original cement paste. Water thus acts as both solvent and component in
the formation of these cements. It is also one of the reaction products,
being formed in the acid-base reaction as the cements set.

3.1.1 Water as a solvent


It is widely recognized that the solvent in which any chemical reaction
takes place is not merely a passive medium in which relevant molecules
perform: the solvent itself makes an essential contribution to the reaction.
The character of the solvent will determine which chemical species are
soluble enough to enter solution and hence to react, and which species are
insoluble, and thus precipitate out of solution, thereby being prevented
from undergoing further chemical change. In the case of water, as will be
seen, polar and ionic species are the ones that most readily dissolve. But
even so, mere polarity or ionic character is not sufficient to ensure
solubility. Solubility depends on a number of subtle energetic factors, and
the possible interactions between water and silver chloride, for example, do
not fulfil the requirements despite the ionic nature of the silver salt. Hence
silver chloride is almost completely insoluble in water.

3.1.2 Water as a component


In AB cements water does not merely act as solvent for the setting reaction.
It also acts as an important component of the set cement. For example,

30
Water
glass-ionomer dental cements as generally formulated include at least
15% by mass of water, all of which becomes incorporated into the
complete cement (Wilson & McLean, 1988). Indeed, great importance is
attached to the retention of water by these cements, since if they are
allowed to dry out by storage under conditions of low humidity, they
shrink significantly, and develop cracks and crazes.
Another class of AB cement, the oxychloride cements of zinc and
magnesium, are also formulated in aqueous solution and retain substantial
amounts of water on setting (Sorrell & Armstrong, 1976; Sorrell, 1977).
Water may have a number of roles in the set versions of these cements.
It is capable of solvating the cement-forming ions, such as Ca2+ or Zn2+,
depending on the cement. It also contributes a sheath of solvating
molecules around polyelectrolytes such as poly(acrylic acid) in glass-
ionomer and zinc polycarboxylate cements. Significant amounts of water
are known to be retained by metal polyacrylate salts at equilibrium and
this water contributes to reducing the glass transition temperature of such
materials by acting as a plasticizer (Yokoyama & Hiraoko, 1979).
These various aspects of water in AB cements are covered in the present
chapter. Its solvent character, structure and hydration behaviour are
described, and the chapter concludes with a more thorough consideration
of the precise role of water in the various AB cements.

3.2 Water
3.2.1 Constitution
Water has a deceptively simple chemical constitution, consisting as it does
of molecules containing two atoms of hydrogen and one of oxygen. It was
viewed by the ancients as one of the four 'elements', following Aristotle's
classification, the others being air, fire and earth. The modern view that it
is a compound composed of hydrogen and oxygen was first established in
1789 by two amateur chemists, Adriaan Paets van Troostwijk (1752-1837),
a merchant, and Jan Rudolph Deiman (1743-1808), a pharmacist (Hall,
1985). They were able to show by synthesis which elements combine to
make water, forming it from reaction of hydrogen gas with oxygen. Their
work was important historically for the part it played in undermining the
phlogiston theory of combustion. It was left to the great Swedish chemist
J. J. Berzelius (1779-1848) to determine that the ratio of hydrogen to
oxygen is 2:1.

31
Water and acid-base cements
Table 3.1. Molecular dimensions of normal and isotopic water in
the vapour phase {Benedict, Gailar & Plyler, 1956)

Bond length, Bond angle,


Molecule pm degrees

D2O 95-75 104-474


H2O 95-718 104-523
HDO 95-71 104-529

As a compound water is remarkable. It is the only inorganic liquid to


occur naturally on earth, and it is the only substance found in nature in all
three physical states, solid, liquid and vapour (Franks, 1983). It is the most
readily available solvent and plays a vital role in the continuation of life on
earth. Water circulates continuously in the environment by evaporation
from the hydrosphere and subsequent precipitation from the atmosphere.
This overall process is known as the hydrologic cycle. Reports estimate
that the atmosphere contains about 6 x 1015 litres of water, and this is
cycled some 37 times a year to give an annual total precipitation of
224 x 1015 litres (Franks, 1983; Nicholson, 1985).
The bond lengths and bond angle for the water molecule are known
very precisely following studies of the rotation-vibration spectra of water
vapour, and also the vapour of the deuterated analogues of water, D2 O
and HDO (Eisenberg & Kauzmann, 1969). The data for these compounds
are shown in Table 3.1. The nuclei of the water molecule, regardless of the
isotopes involved, form an isosceles triangle having a slightly obtuse angle
at the oxygen atom. All of the data in Table 3.1 refer to the equilibrium
state of the water molecules, which is formally acceptable, but is actually
a hypothetical state, since it assumes neither rotation nor vibration in the
molecule.
The equilibrium bond lengths and bond angles can be seen to differ little
between the different isotopic molecules. Such a finding agrees with the
predictions of the Born-Oppenheimer approximation, that the electronic
structure of a molecule is independent of the mass of its nuclei, it being the
electronic structure of a molecule alone which determines the geometry.
The bond angle in water is slightly less than the ideal tetrahedral angle
of 109-5°.This is attributed to the presence of lone pairs of electrons on the
oxygen atom which repel more strongly than the bonding pairs of electrons
between the oxygen and hydrogen nuclei (Speakman, 1975). The valence-

32
Water
Table 3.2. Properties of hydrides of first row elements (Weast, 1985-6)

Gas
phase
dipole
Relative Melting point, Boiling point, moment,
Compound molar mass °C °C debyes

CH 4 16 -182-0 -1640 000


NH 3 17 -11-1 -33-4 1-47
H2O 18 00 1000 1-85
HF 19 -834 19-5 1-82

shell electron-pair repulsion concepts of Gillespie & Nyholm (1957) show


that such increased repulsion by lone pairs closes the angle between the
bonding pairs slightly but significantly for the water molecule.
The O-H bond energy of water is taken as half the energy of formation
of the molecule, since water has two such bonds. This gives a value of
458-55 kJ mol"1 at 0 K (Eisenberg & Kauzmann, 1969). Related to the
bond energy is the dissociation energy, i.e. the energy required to break the
bond at 0 K. Neither of the O-H bonds in water has a dissociation energy
equal to the O-H bond energy. Instead, the first O-H dissociation energy
has been found experimentally to be 424-27 kJ mol"1. From conservation
of energy considerations which lead to the requirement that the sum of the
two dissociation energies must equal the energy of formation, it is found
that the second O-H dissociation energy has to take a value of
492-83 kJ mol"1. This has been explained (Pauling, 1960) by postulating an
electronic rearrangement on the oxygen atom of the O-H fragment left
behind after scission of the first O-H bond, and that breaking the bond
between oxygen in this new electronic configuration and the remaining
hydrogen requires greater energy.

3.2.2 Water compared with other hydrides


Water shows properties that are interestingly different compared with
hydrides of the neighbouring elements of thefirstrow of the periodic table.
Some of these properties are given in Table 3.2. From this table, water can
be seen to have a very high melting point and a very high boiling point for
its relative molar mass. Indeed, it is the only one of the hydrides of the

33
Water and acid-base cements
elements from this portion of the periodic table to be liquid at room
temperature and atmospheric pressure. In the gas phase it has a dipole
moment that, while only slightly greater than that of hydrogenfluoride,is
the highest for this group of hydrides. All of these properties point to water
having a structure in which its constituent molecules are more highly
associated and interact more strongly than the molecules of the closely
related hydrides.

3.3 The structure of water


At first sight the concept of a 'structure' for liquid water appears strange.
In the solid state atoms are relatively fixed in space, albeit with some
vibrational motion about equilibrium positions, and no difficulty is
associated with the idea of locating these equilibrium positions by some
appropriate physical technique, and thereby assigning a structure to the
solid.

3.3.1 The concept of structure in the liquid state


With water or any other liquid, molecules do not occupy even reasonably
fixed locations but have considerably more freedom for movement than in
the solid state. What then do we mean by the term structure applied to a
liquid?
To answer this question we need to consider the kind of physical
techniques that are used to study the solid state. The main ones are based
on diffraction, which may be of electrons, neutrons or X-rays (Moore,
1972; Franks, 1983). In all cases exposure of a crystalline solid to a beam
of the particular type gives rise to a well-defined diffraction pattern, which
by appropriate mathematical techniques can be interpreted to give
information about the structure of the solid. When a liquid such as water
is exposed to X-rays, electrons or neutrons, diffraction patterns are
produced, though they have much less regularity and detail; it is also more
difficult to interpret them than for solids. Such results are taken to show
that liquids do, in fact, have some kind of long-range order which can
justifiably be referred to as a 'structure'.
In considering the structure of a liquid, two possible conceptual
approaches exist. One is to begin from an understanding of the gaseous

34
The structure of water
state, characterized as it is by gross translational movement of the
constituent molecules and substantial disorder. The liquid is then viewed as
a gas that has been condensed and in which translational motion has
become constrained. Alternatively, consideration can start from the solid
state, with its well-characterized structure, having little or no translational
motion, but some vibrational motion of the constituent atoms or
molecules. The liquid state is then viewed as a solid in which some degree
of translational motion has become allowed, but with a structure still
recognizable as being derived from that existing in the solid state (Franks,
1983). With the growth in application of the techniques of X-ray and
neutron diffraction to the study of the liquid state, the latter approach has
become increasingly favoured in recent years.
In this section, rather than give a detailed account of theories of the
liquid state, a more qualitative approach is adopted. What follows includes
first a description of the structure of ice; then from that starting-point,
ideas concerning the structure of liquid water are explained.

3.3.2 The structures of ice


Water is capable of solidifying into a number of different structural states
or polymorphs depending, for example, on the external pressure applied
during solidification. The simplest and most common of these polymorphs
is known as ice I, whose structure was first determined by W. H. Bragg
(1922). In this structure, every oxygen atom occupies the centre of a
tetrahedron formed by four oxygen atoms, each about 0-276 nm away. The
water molecules are connected together by hydrogen bonds, each molecule
being bonded to its four nearest neighbours. The O-H bonds of a given
molecule are oriented towards the lone pairs on two of these neighbouring
molecules, and in turn, each of its lone pairs is directed towards an O-H
bond of one of the other neighbours. This arrangement gives an open
lattice in which intermolecular cohesion is large.
The arrangement of oxygen atoms in ice I is isomorphous with the
wurtzite form of zinc sulphide, and also with the silicon atoms in the
tridymite form of silicon dioxide. Hence, ice I is sometimes referred to as
the wurtzite or tridymite form of ice (Eisenberg & Kauzmann, 1969).
Location of the hydrogen atoms in ice I has caused more problems. This
is because hydrogen is less effective at scattering X-rays or electrons than
oxygen. For a long time, arguments about the position of hydrogen were
based on indirect evidence, such as vibrational spectra or estimates of

35
Water and acid-base cements
residual entropy at 0 K (Eisenberg & Kauzmann, 1969). Since the advent
of neutron diffraction the positions of the hydrogen atoms have become
clearer. These studies have shown that the water molecules have very
similar dimensions in ice I to those in the isolated molecule: the O-H bond
length is 0-101 nm and the bond angle 104*5°.
Ice I is one of at least nine polymorphic forms of ice. Ices II to VII
are crystalline modifications of various types, formed at high pressures;
ice VIII is a low-temperature modification of ice VII. Many of these
polymorphs exist metastably at liquid nitrogen temperature and atmos-
pheric pressure, and hence it has been possible to study their structures
without undue difficulty. In addition to these crystalline polymorphs, so-
called vitreous ice has been found within the low-temperaturefieldof ice I.
It is not a polymorph, however, since it is a glass, i.e. a highly supercooled
liquid. It is formed when water vapour condenses on surfaces cooled to
below -160°C.
It is not appropriate in this chapter to give a detailed review of the solid-
state behaviour of water in its various crystalline modifications. However,
there are some general structures which are relevant and worth high-
lighting. Firstly, water molecules in these various solids have dimensions
and bond angles which do not differ much from those of an isolated water
molecule. Secondly, the number of nearest neighbours to which each
individual molecule is hydrogen-bonded remains four, regardless of the ice
polymorph.
The differences in structure between the polymorphs, particularly the
high-pressure ones, lie in (a) the distances between the non-hydrogen
bonded molecules, and hence the amount of' free volume' in the structure,
(b) the angles of the hydrogen bonds, which may differ markedly from the
180° of ice I, and (c) the distance between nearest neighbouring oxygen
atoms, which may fall to well below the 0-276 nm value in ice I. All of these
are consistent with closer packing of the water molecules, and a closing up
of the cage structure by comparison with that found for ice I.

3.3.3 Liquid water


Before considering the details of the structure of liquid water, it is
important to define precisely what is meant by the term structure as applied
to this liquid. If we start from ice I, in which molecules are vibrating about
mean positions in a lattice, and apply heat, the molecules vibrate with
greater energy. Gradually they become free to move from their original

36
The structure of water
lattice sites and acquire significant translational energy. However, trans-
lational energy is not confined to molecules in the liquid state. There is a
finite possibility of any molecule in ice I moving from its lattice site, thus
acquiring translational energy. In principle, a given molecule can move
through the solid structure in a process that is essentially diffusion.
From this model of ice I we derive three meanings of the term structure
for the solid. We may refer to the positions of the molecules at an instant
of time. We may allow some averaging of the positions, i.e. we may have
a vibrationally averaged structure, considered over a short time-period,
during which molecules have time to undergo only minor vibrational
reorientations. Finally we may have a diffusionally averaged structure,
considered over longer time-periods, in which the minor translational
motion has been allowed to proceed to such an extent as to be significant.
These three possible structures, the instantaneous, the vibrationally
averaged and the diffusionally averaged, are referred to as I-, V- and D-
structures respectively.
Let us now turn our attention to liquid water. Just as in ice I, molecular
motions may be divided into rapid vibrations and slower diffusional
motions. In the liquid, however, vibrations are not centred on essentially
fixed lattice sites, but around temporary equilibrium positions that are
themselves subject to movement. Water at any instant may thus be
considered to have an I-structure. An instant later, this I-structure will be
modified as a result of vibrations, but not by any additional displacements
of the molecules. This, together with the first I-structure, is one of the
structures that may be averaged to allow for vibration, thereby con-
tributing to the V-structure. Lastly, if we consider the structure around an
individual water molecule over a long time-period, and realize that there is
always some order in the arrangement of adjacent molecules in a liquid
even over a reasonable duration, then we have the diffusionally averaged
D-structure.
No experimental technique exists for determining I-structures in either
the liquid or the solid state. Techniques do exist for obtaining information
on both the V- and D-structures of liquid water; the results of applying
these techniques are considered next.
Spectroscopic studies have established that for liquid water, the V-
structure has the following features.

(a) Considerable local variation between the environments of the


individual water molecules, compared with the relatively uniform

37
Water and acid-base cements
molecular environments in a crystal of ice I. The frequency spans
of the uncoupled O-H and O-D spectral bands indicate that some
nearest neighbours are as close as 0-275 nm, while others are
separated by 0.310 nm or more. The most probable equilibrium
separation is about 0.285 nm (Eisenberg & Kauzmann, 1969).
(b) The differences between the various molecular environments are
continuous. In other words, the V-structure does not contain
discrete types of molecular environment.
(c) The frequency of the stretching band indicates that hydrogen
bonds in the V-structure are weaker than those in ice I, though still
distinctly present.

Ideas about the D-structure have come mainly from two sources, namely
a consideration of the underlying reasons for the values of certain physical
properties, such as heat capacity or compressibility, and a study of radial
distribution functions that arise from X-ray diffraction work on liquid
water. The D-structure represents the average arrangement of molecules
around an arbitrary central water molecule. This average is either the
'space average' for several central molecules in different V-structures, or
the 'time average' for a single molecule over very long periods of time.
Near the freezing point, the D-structure is found to have relatively high
concentrations of neighbours at distances 0-29, 0-50 and 0*70 nm from the
central water molecule. This suggests that a substantial hydrogen-bonded
network is discernible, even in the liquid state. As the temperature is raised,
so the distinct concentrations at 0-50 and 0-70 nm disappear. Thermal
agitation thus distorts or destroys the hydrogen-bonded networks, and the
amount of observable long-range order decreases significantly.
Structural studies on liquid water reveal that the majority of molecules
are effectively tetrahedral, since the O-H bonds and the lone pairs are used
in hydrogen-bonding. Questions remain about the nature of these
hydrogen-bonds (Symons, 1989). Specifically: on average, how many such
hydrogen bonds are formed per molecule, how strong and how linear are
they, and what is their lifetime? One recent approach has been to consider
the possibility that, because of their weakness, some of the hydrogen bonds
in liquid water will break. This then gives concentrations of free O-H
bonds, OHfree, and free lone pairs, LPfree, on certain molecules which are
bonded to only three others (Symons, 1989). Symons (1989) also suggests
that the chemical properties of liquid water depend on the relative
concentrations of these species. Fully hydrogen-bonded water can be

38
The structure of water
considered as inert; reactions requiring attack by O-H depend on the
concentration of OHfree molecules; and those requiring nucleophilic attack
by lone pairs depend on the concentration of LPfree molecules. Evidence
for OHfree and LPfree molecules has been obtained spectroscopically using
monomeric deuterated water, HOD, in inert solvents such as dimethyl
sulphoxide, though debate continues over interpretation of the results
obtained in such studies.
An important phenomenon when considering the differences between
ice I and liquid water is that water achieves its maximum density not in the
solid state, but at 4 °C, i.e. in the liquid state. The reasons for this were first
discussed by Bernal & Fowler (1933). They noted that the separation of
molecules in ice I is about 0-28 nm, corresponding to an effective molecular
radius of 014 nm. Close packing of molecules of such radius would yield
a substance of density 1*84 g cm"3. To account for the observed density of
10 g cm"3, it was necessary to postulate that the arrangement of molecules
was very open compared with the disordered, close-packed structures of
simple liquids such as argon and neon.
The increase in density on melting is assumed to arise from two
competing effects that occur as water is heated. First, increasing trans-
lational freedom for the water molecules weakens the hydrogen-bonded
network that exists in ice I. This network thus collapses, and reduces the
volume. Second, increased vibrational energy for the molecules causes an
effective increase in the volume occupied by any one molecule, thus
enlarging the overall volume of the liquid. The first effect is considered to
predominate below 4 °C, the second above 4 °C.
Overall, the main conclusions that are to be drawn concerning the
structure of liquid water are as follows.
(a) Water has a degree of long-range order that is appropriately
described as structure; it is possible to measure detailed para-
meters for either a vibrationally averaged or a diffusionally
averaged structure.
(b) The force between molecules that sustains this order in the liquid
state is the hydrogen bond.
(c) The bond lengths and angles of individual water molecules are
almost independent of whether they occur in ice I or liquid water.

39
Water and acid-base cements

3.4 Water as a solvent


The general criterion for solubility is the rule that 'like dissolves like'. In
other words polar solvents dissolve polar and ionic solutes, non-polar
solvents dissolve non-polar solutes. In the case of water, this means that
ionic compounds such as sodium chloride and polar compounds such as
sucrose are soluble, but non-polar compounds such as paraffin wax are
not.
In general, solubility depends on the relative magnitudes of three pairs
of interactions, namely solute-solute, solvent-solvent and solute-solvent
(Robb, 1983). For a substance to be soluble in a given liquid, the
solute-solvent interactions must be greater than or equal to the other two
interactions.
Insolubility does not only result from the kind of energetic consider-
ations outlined above. It can also be the result of essentially kinetic
barriers. For example, the naturally occurring macromolecule cellulose is
not soluble in water, yet its monomer, D( + )-glucose is extremely water-
soluble (Morrison & Boyd, 1973). This is because cellulose adopts a well-
ordered structure, in which individual hydroxyl groups are aligned via
hydrogen bonds; the overall structure simply has too great an integrity to
allow water molecules to enter and hydrate the individual molecules in
order to carry them off into solution.

3.4.1 Hydrophobic interactions


The qualitative discussion of solubility has focussed so far on the attractive
forces in solute-solvent interactions. However, where water is concerned,
it is also important to consider the forces of repulsion due to the so-called
'hydrophobic' interactions that may arise in certain cases (Franks,
1975). These hydrophobic interactions may be explained in terms of
thermodynamic concepts.
Measuring enthalpy changes for the dissolution of hydrocarbons, such
as alkanes, in water shows that heat is evolved, i.e., AH is negative and
energetically water and alkanes attract each other. However, such
attraction does not make alkanes soluble in water to any appreciable
extent. This is because the free energy change AGsolution opposes the
process and is positive.
From the Gibbs equation,
Absolution — Absolution 1 Absolution

40
Water as a solvent
it follows that the ^ASsolution term (and hence A*Ssolution itself) must be
negative. This means that the proposed solution has lower entropy and is
more ordered than pure water, which is a striking conclusion, since entropy
is usually increased by mixing. It occurs because the relatively ordered
structure of the liquid water, based as it is on a hydrogen-bonded array of
water molecules, actually becomes more ordered when alkane molecules
enter it. This result is attributed to the formation of a 'cage' structure of
water molecules around the non-polar alkane molecule, in which water has
less vibrational and translational freedom than in the pure liquid (Franks,
1983).
In cases where the solvation energies are large, as for example when ionic
compounds dissolve in water, these hydrophobic effects, based on adverse
changes in entropy, are swamped. Dissolving such compounds can be
readily accomplished due to the very large energies released when the ions
become hydrated.

3.4.2 Dissolution of salts


Salts dissolve in water with dissociation of the constituent ions, this
concept having been proposed originally by S. Arrhenius in 1887. His first
idea was that all salts, including those of what would now be regarded as
weak acids or bases, are completely dissociated at extreme dilution (Hall,
1985). It was eventually realized that substances such as NaCl, KC1, etc,
are effectively completely dissociated at all concentrations.
Dissolution of an ionic salt is essentially a separation process carried out
by the interaction of the salt with water molecules. The separation is
relatively easy in water because of its high dielectric constant. Comparison
of the energies needed to separate ions of NaCl from 0-2 nm to infinity
shows that it takes 692-89 kJ mol"1 in vacuum, but only 8-82 kJ mol"1 in
aqueous solution (Moore, 1972). Similar arguments have been used to try
to estimate solvation energies of ions in aqueous solution, but there are
difficulties caused by the variations in dielectric constant in the immediate
vicinity of individual ions.
In order to dissolve ionic solutes so readily, water molecules must
solvate the ions as they enter solution. Consequently, water molecules lose
their translational degrees of freedom as a result of their association with
specific ions. It is possible to estimate the number of water molecules in
clusters of the type M+ (H2O)W using mass spectrometry (Kebarle, 1977).

41
Water and acid-base cements
The number of water molecules in such a cluster, the hydration number,
varies with ionic size; it is four for Li+, three for Na+, but only one for Rb+.
Mass spectrometry has been used to study the energetics of solvation
and has shown that the enthalpies of attachment of successive water
molecules to either alkali metal or halide ions become less exothermic as
the number of water molecules increases (Kebarle, 1977). The Gibbs free
energies of attachment for water molecules have also been found to be
negative.
The different hydration numbers can have important effects on the
solution behaviour of ions. For example, the sodium ion in ionic crystals
has a mean radius of 0-095 nm, whereas the potassium ion has a mean
radius of 0133 nm. In aqueous solution, these relative sizes are reversed,
since the three water molecules clustered around the Na+ ion give it a
radius of 0-24 nm, while the two water molecules around K+ give it a radius
of only 0-17 nm (Moore, 1972). The presence of ions dissolved in water
alters the translational freedom of certain molecules and has the effect of
considerably modifying both the properties and structure of water in these
solutions (Robinson & Stokes, 1955).
The precise orientation of water molecules around cations is not clear,
though two models have been proposed for the possible structures that
occur (Vaslow, 1963). In one, the water molecules are arranged so that the
dipole moments are aligned with the centres of the ions. In the other, water
molecules are arranged so that interaction between the lone-pair orbitals
on the oxygen atoms and orbitals on the cation is maximized. This latter
model is supported by molecular dynamics calculations (Heinzinger &
Palinkas, 1987; Heinje, Luck & Heinzinger, 1987).
Less uncertainty surrounds the structure of hydrated anions: the
hydrogen atoms are almost collinear with the oxygen atoms and the
centres of the ions (Briant & Burton, 1976). Monte-Carlo calculations have
shown that F~ is surrounded by four hydrogen atoms each 017 nm away
(Watts, Clementi & Fromm, 1974).
Neutron scattering has been used for studying the state of solvation of
ions in aqueous solution (Enderby et ai, 1987; Salmon, Neilson &
Enderby, 1988). These studies have shown that a distinct shell of water
molecules of characteristic size surrounds each ion in solution. This
immediate hydration shell was called zone A by Frank & Wen (1957); they
also postulated the existence of a zone B, an outer sphere of molecules, less
firmly attached, but forming part of the hydration layer around a given
ion. The evidence for the existence of zone B lies in the thermodynamics of

42
Water as a solvent

the hydration process, and may be appreciated by considering the


isoelectronic species KCl and two moles of argon.
The standard enthalpy of hydration is significantly more exothermic for
KCl than for the two moles of argon; however, the corresponding entropy
of hydration is less for KCl than for argon. The results for the values of
enthalpy can be readily understood in terms of the greater intensity of the
interactions between water molecules and the ions of KCl than between
those of water molecules and the uncharged argon atoms. At first sight the
greater loss of freedom in the water molecules involved in hydrating the
ions of KCl would be expected to reduce disorder in such solutions. In
other words the entropy of hydration for KCl ought to be greater than the
entropy of hydration for two moles of argon. To explain the fact that the
opposite is found experimentally, Frank & Evans (1945) suggested that
there is a compensating gain in entropy which can be attributed to
disruptions in the water-water interactions within zone B.
As a result of these electrostatic effects aqueous solutions of electrolytes
behave in a way that is non-ideal. This non-ideality has been accounted for
successfully in dilute solutions by application of the Debye-Huckel theory,
which introduces the concept of ionic activity. The Debye-Huckel limiting
law states that the mean ionic activity coefficient y± can be related to the
charges on the ions, z+ and z_, by the equation

Iog 10 y ± =-0-509z + z_

Ionic activity essentially represents the concentration of a particular


type of ion in aqueous solution and is important in the accurate
formulation of thermodynamic equations relating to aqueous solutions of
electrolytes (Barrow, 1979). It replaces concentration because a given ion
tends not to behave as a discrete entity but to gather a diffuse group of
oppositely charged ions around it, a so-called ionic atmosphere. This
means that the effective concentration of the original ion is less than its
actual concentration, a fact which is reflected in the magnitude of the ionic
activity coefficient.
Debye-Hiickel theory assumes complete dissociation of electrolytes into
solvated ions, and attributes ionic atmosphere formation to long-range
physical forces of electrostatic attraction. The theory is adequate for
describing the behaviour of strong 1:1 electrolytes in dilute aqueous
solution but breaks down at higher concentrations. This is due to a
chemical effect, namely that short-range electrostatic attraction occurs

43
Water and acid-base cements
either between ion-pairs or between solvent-separated ion-pairs (Russo &
Hanania, 1989); this effect becomes important in concentrated aqueous
solutions of the type used to form AB cements.

3.4.3 Ion-ion interactions in water


Two ions are particularly important in the chemistry of water, namely H+
and OH~ (Clever, 1963). Hydrogen ions do not exist as discrete entities.
This is because ionization of the hydrogen atom leaves behind a proton,
which is very small compared with a typical ion. Thus the local charge
density developed around the proton is very high. The polarizing effect of
such a high charge density is such that the resulting system is simply too
unstable to form to any detectable extent in aqueous solutions.
When water undergoes self-ionization, a range of cationic species are
formed, the simplest of which is the hydronium ion, H3O+ (Clever, 1963).
This ion has been detected experimentally by a range of techniques
including mass spectrometry (Cunningham, Payzant & Kebarle, 1972), as
have ions of the type H+ (H2O)W with values of n up to 8. Monte-Carlo
calculations show that H3O+ ions exist in hydrated clusters surrounded by
three or four water molecules in the hydration shell (Kochanski, 1985).
These ions have only a short lifetime, since the proton is highly mobile and
may be readily transferred from one water molecule to another. The time
taken for such a transfer is typically of the order of 10~14 s provided that
the receiving molecule of water is correctly oriented.
Several other discrete species have been found to arise from the self-
ionization of water. These include H5Og (Kearley, Pressman & Slade,
1986), H4O2+ (Bollinger et al., 1987), H9O+ (Robinson, Thistlethwaite &
Lee, 1986) and hydrated electrons (Hart & Anbar, 1970).
Intense ion-ion interactions which are characteristic of salt solutions
occur in the concentrated aqueous solutions from which AB cements are
prepared. As we have seen, in such solutions the simple Debye-Huckel
limiting law that describes the strength goes up so the repulsive force
between the ions becomes increasingly important. This is taken account of
in the full Debye-Hiickel equation by the inclusion of a parameter related
to ionic size and hence distance of closest approach (Marcus, 1988).
For concentrated solutions, there are approaches that are more
sophisticated than that of Debye & Hiickel. A particularly successful
method of describing such solutions is that due to McMillan & Mayer
(1945) which has subsequently been developed by Ramanathan &

44
Water as a solvent
Friedman (1971). This approach is described as the hypernetted chain
procedure and in it ion-ion pair potentials are expressed as the sum of four
terms. These are:
COUL U , the charge-charge interactions between two ions, / and
j , as a function of their separation,
COR U , a repulsive core potential term,
CAVU, arising from the dielectric cavity effect, and
GUR U , the so-called Gurney potential (Gurney, 1953), which
describes the effect of co-sphere overlap.
Using this approach, calculations can be made of volumetric, entropic
and energy parameters taking account of the effect of overlapping co-
spheres. Some indication of the organization in the solution is also
possible. The properties of a number of concentrated salt solutions have
been analysed by this procedure, including simple 1:1 salts, alkaline earth
salts and alkylammonium salts.
A number of other attempts have been made to account for the
properties of concentrated aqueous solutions of ionic compounds by
procedures that represent further improvements on the simple Debye-
Hiickel approach. However, they lie outside the scope of the present
chapter. The important point to emphasize is that the concentrated
aqueous solutions that are generally employed in the preparation of AB
cements tend to exhibit significant ion-ion interactions; such interactions
lead to significant deviations from ideality which may be accounted for by
substantial extension of the ideas of simple dilute solution theory.

3.4.4 Dissolution of polymers


AB cements are not only formulated from relatively small ions with well
defined hydration numbers. They may also be prepared from macro-
molecules which dissolve in water to give multiply charged species known
as polyelectrolytes. Cements which fall into this category are the zinc
polycarboxylates and the glass-ionomers, the polyelectrolytes being
poly(acrylic acid) or acrylic acid copolymers. The interaction of such
polymers is a complicated topic, and one which is of wide importance to a
number of scientific disciplines. Molyneux (1975) has highlighted the fact
that these substances form the focal point of 'three complex and
contentious territories of science', namely aqueous systems, ionic systems
and polymeric systems.

45
Water and acid-base cements
In polyelectrolytes, the ionic charges are carried by groups which are
themselves attached covalently to the macromolecular backbone. When all
of the groups are negatively charged, as with polyacrylate, the poly-
electrolyte that results is referred to as a polyanion. Polyelectrolytes are of
high solubility in water, especially when compared to most organic
macromolecules. This increased solubility may be attributed not only to
the more favourable interactions between the charged groups and the
water molecules, but also to the fact that entropy strongly favours
dissolution and dissociation of these molecules (Molyneux, 1975).
The conformations adopted by polyelectrolytes under different condi-
tions in aqueous solution have been the subject of much study. It is known,
for example, that at low charge densities or at high ionic strengths
polyelectrolytes have more or less randomly coiled conformations. As
neutralization proceeds, with concomitant increase in charge density, so
the polyelectrolyte chain uncoils due to electrostatic repulsion. Eventually
at full neutralization such molecules have conformations that are
essentially rod-like (Kitano et ai, 1980). This rod-like conformation for
poly(acrylic acid) neutralized with sodium hydroxide in aqueous solution
is not due to an increase in stiffness of the polymer, but to an increase in the
so-called excluded volume, i.e. that region around an individual polymer
molecule that cannot be entered by another molecule. The excluded
volume itself increases due to an increase in electrostatic charge density
(Kitano et al.9 1980).
In a study of the transition in conformation from random coil to stiff rod
by poly(acrylic acid), it was found that the point of transition depended on
a number of factors, including the nature of the solvent, the temperature,
the particular counterion used and the degree of dissociation (Klooster,
van der Trouw & Mandel, 1984).
Methanol was used in this study, though in terms of Flory-Huggins
solution theory it is not a good solvent for poly(acrylic acid) at room
temperature. In other words, the polymer adopts a tightly coiled
conformation that excludes solvent, thereby approaching the point of
precipitation. This phenomenon may be responsible for the observation
that the addition of methanol to poly(acrylic acid) solutions intended for
use in glass-ionomer cements prevents gelation (Wilson & Crisp, 1974).
This was originally attributed to methylation of the polymer, leading to a
reduction in the stereoregularity of the poly(acrylic acid) and hence to a
lessening of the readiness with which stable hydrogen-bonded links could
be formed. However, there is the alternative possibility that the presence of

46
Hydration in the solid state
methanol altered the conformation of the polymer and that this con-
formational change prevented the development of the hydrogen-bonded
network necessary for gelation to occur.

5.5 Hydration in the solid state


Many ionic compounds contain what used to be referred to as 'water of
crystallization'. For example, magnesium chloride can exist as a fully
hydrated salt which was formerly written MgCl2. 6H2O, but is more
appropriately written Mg(OH2)6Cl2, since the water molecules occupy
coordination sites around the magnesium ions. This is typical. In most
compounds that contain water of crystallization, the water molecules are
bound to the cation in an aquo complex in the manner originally proposed
by Alfred Werner (1866-1919) in 1893 (Kauffman, 1981). Such an
arrangement has been confirmed in numerous cases by X-ray diffraction
techniques.

3.5.1 Coordination of water to ions


The ions that tend to be involved in AB cements include such species as
Al3+, Mg2+, Ca2+ and Zn2+. These are all capable of developing a
coordination number of six, and hexaquo cations are known to be formed
by each of these metal ions (Hiickel, 1950). The typical requirements for an
ion to develop such coordination characteristics are that the ion should
exist in the + 2 or +3 oxidation state, and in this state should be of small
ionic radius (Greenwood & Earnshaw, 1984).
Another feature of the metal ions that are typically involved in
cementitious bonding in AB cements is that most of them fall into the
category of hard in Pearson's Hard and Soft Acids and Bases scheme
(Pearson, 1963). The underlying principle of this classification is that bases
may be divided into two categories, namely those that are polarizable or
soft, and those that are non-polarizable or hard. Lewis acids too may be
essentially divided into hard and soft, depending on polarizability. From
these classifications emerges the useful generalization that hard acids
prefer to associate with hajd bases and soft acids prefer to associate with
soft bases (see Section 2.3.7).
Of the ions most often implicated in cementitious bonding in AB
cements, Ca2+, Mg2+ and Al3+ are classified as clearly hard; Zn2+ by
contrast falls into the category that Pearson designated 'borderline', as

47
Water and acid-base cements
does Cu2+ (Pearson, 1963). This means that most of these ions form
particularly stable complexes with hard bases, i.e. those which are not
readily polarizable. This requirement demands that the bases be those of
first row elements, such as oxygen or nitrogen. Water is thus a hard base,
and the complex that it forms with these ions involves coordination by the
oxygen atoms. As predicted by the Hard and Soft Acids and Bases concept,
the aquo complexes of the cement-forming metal ions are extremely stable
and do not readily lose their coordinated water. Hence, one of the
functions of water in fully set AB cements is coordination to the metal ions.

3.6 The role of water in acid-base cements


Water has three possible roles in acid-base cements. First, it acts as the
medium for the setting reaction of these materials, and second, it is one of
the components of the set cement, actually becoming incorporated into the
cement as it hardens. Third, water may act as plasticizer in these cements.
All of these roles are reviewed here.

3.6.1 Water as solvent in AB cements


Water as the solvent is essential for the acid-base setting reaction to occur.
Indeed, as was shown in Chapter 2, our very understanding of the terms
' acid' and' base', at least as established by the Bronsted-Lowry definition,
requires that water be the medium of reaction. Water is needed so that the
acids may dissociate, in principle to yield protons, thereby enabling the
property of acidity to be manifested. The polarity of water enables the
various metal ions to enter the liquid phase and thus react. The solubility
and extent of hydration of the various species change as the reaction
proceeds, and these changes contribute to the setting of the cement.

3.6.2 Water as a component of AB cements


Water is also a component of set AB cements. In glass-ionomer cements,
for example, it may serve to coordinate to certain sites around the metal
ions. It also hydrates the siliceous hydrogel that is formed from the glass
after acid attack has liberated the various metal ions (Wilson & McLean,
1988). Such reactions continue long after the initial hardening of the
cement is complete, and for this reason water must be retained as far as
possible during the first hours and days after formation of the cement. If
water is lost from the cement and desiccation occurs, these post-hardening

48
The role of water in acid-base cements
reactions stop, and the cement will not achieve maximum possible strength.
Moreover, if desiccation is excessive, the material will also shrink and
crack.
Water occurs in glass-ionomer and related cements in at least two
different states (Wilson & McLean, 1988; Prosser & Wilson, 1979).
These states have been classified as evaporable and non-evaporable,
depending on whether the water can be removed by vacuum desiccation
over silica gel or whether it remains firmly bound in the cement when
subjected to such treatment (Wilson & Crisp, 1975). The alternative
descriptions 'loosely bound' and 'tightly bound' have also been applied to
these different states of water combination. In the glass-poly(acrylic acid)
system the evaporable water is up to 5 % by weight of the total cement,
while the bound water is 18-28 % (Prosser & Wilson, 1979). This amount
of tightly bound water is equivalent to five or six molecules of water for
each acid group and associated metal cation. Hence at least ten molecules
of water are involved in the hydration of each coordinated metal ion at a
carboxylate site.
It has been suggested by Ikegami (1968) that the carboxylate groups of
a polyacrylate chain are each surrounded by a primary local sphere of
oriented water molecules, and that the polyacrylate chain itself is
surrounded by a secondary sheath of water molecules. This secondary
sheath is maintained as a result of the cooperative action of the charged
functional groups on the backbone of the molecule. The monovalent ions
Li+, Na+ and K+ are able to penetrate only this secondary hydration
sheath, and thereby form a solvent-separated ion-pair, rather than a
contact ion-pair. Divalent ions, such as Mg2+ or Ba2+, cause a much greater
disruption to the secondary hydration sheath.
The effectiveness with which divalent ions cause gelation of poly(acrylic
acid) has been found to follow the order Ba2+ > Sr2+ > Ca2+ (Wall &
Drenan, 1951) and this has been attributed to the formation of salt-like
crosslinks. Gelation has been assumed to arise in part from dehydration of
the ion-pairs (Ikegami & Imai, 1962), and certainly correlates with
precipitation in fairly dilute systems. Indeed, the term precipitation has
sometimes been applied to the setting of AB cements derived from
poly(acrylic acid) as they undergo the transition from soft manipulable
paste to hard brittle solid.
At the molecular level, a number of features are associated with the
phenomenon of gelation or precipitation. In particular the disruption of
the secondary hydration sheaths around the polyacrylate chains appears

49
Water and acid-base cements
important (Prosser & Wilson, 1979). In glass-ionomer cements the bound
water has been assumed to be associated with the intrinsic water spheres
around the carboxylate anion-metal cation units, while evaporable water
is associated with the secondary hydration sheath around the polyacrylate
chain. As these cements age, the ratio of tightly bound to loosely bound
water increases. This is accompanied by an increase in strength and
modulus of the cement and by a decrease in plasticity (Paddon & Wilson,
1976; Wilson, Crisp & Paddon, 1981).
The loosely bound water in glass-ionomer cements is labile, and is easily
lost or gained. Indeed, such cements are stable only in an atmosphere of
80% relative humidity (Hornsby, 1980). In higher humidities the cements
absorb water and the resulting hydroscopic expansion can exceed the
shrinkage that usually accompanies setting, which is a distinct clinical
advantage for the use of these cements in dentistry. By contrast, the cement
can lose water under drying conditions leading to shrinking, crazing and
failure to develop full strength.
Glass-ionomer cements become less susceptible to desiccation as they
age, because a greater proportion of the water in older cements has become
'tightly bound'. Early contact with moisture is also damaging, and this
problem is overcome clinically to some extent by using some sort of
protection such as clear nail varnish to seal the cement during its early life
(Wilson & McLean, 1988). However, this does not give perfect results, and
as yet there is no ideal barrier material for this purpose (Earl, Hume &
Mount, 1985).
The role of water in dental silicate cements was studied by Wilson et al.
(1970), and they found that the properties of these materials including
setting time, compressive strength and resistance to attack by water and
acids were markedly affected by the amount of water in the original cement
paste. Water in these materials was also found to fall into the two
categories of evaporable and non-evaporable. In this case non-evaporable
water was defined as that water remaining in the cement after heating at
105 °C for 24 hours. With increasing acid concentration (i.e. decreasing
amounts of water in the initial cement paste), the amount of non-
evaporable water went down, until at 75 % phosphoric acid concentration
nearly all of the water in the cement was found to be evaporable. Moreover
the cements containing almost no non-evaporable water were found to be
extremely weak. Hence the non-evaporable (bound) water could be
equated with bonding water. Infrared analysis had previously shown that
non-bonding water was associated with the water-soluble hydrated salt

50
The role of water in acid-base cements

sodium dihydrogen phosphate, NaH 2 PO 4 .H 2 O (Wilson & Mesley, 1968).


The presence of this compound was known to have a deleterious effect on
cement properties and the water associated with it was known to be readily
removed. Hence in cements prepared from aqueous solutions having high
concentrations of phosphoric acid this salt was assumed to be present in
quantity and to be responsible for the relatively high levels of evaporable
water (Wilson et aL, 1970).
A series of AB cements can be prepared from aqueous solutions of
oxides and halides (or sulphates) of magnesium or zinc. These cements are
described in detail in Chapter 7. For the moment we will confine our
discussion to a consideration of the role of water in these cements.
In the cements of this type a number of phases are known to be present.
For example, in the zinc oxychloride cement two discrete phases,
corresponding to the composition ZnO. ZnCl 2 . H 2 O in the ratios 4:1:5
and 1:1:2 respectively, are known to occur (Sorrell, 1977). Similarly,
in the magnesium oxychloride cement, phases corresponding to
Mg(OH) 2 . MgCl 2 . H 2 O in the ratios 5:1:8 and 3:1:8 have been shown to
exist and have been studied by X-ray diffractometry (Sorrell & Armstrong,
1976).
The precise structural role played by the water molecules in these
cements is not clear. In the zinc oxychloride cement, water is known to be
thermally labile. The 1:1:2 phase will lose half of its constituent water at
about 230 °C, and the 4:1:5 phase will lose water at approximately 160 °C
to yield a mixture of zinc oxide and the 1:1:2 phase. Water clearly occurs
in these cements as discrete molecules, which presumably coordinate to the
metal ions in the cements in the way described previously. However, the
possible complexities of structure for these systems, which may include
chlorine atoms in bridging positions between pairs of metal atoms, make
it impossible to suggest with any degree of confidence which chemical
species or what structural units are likely to be present in such cements.
One is left with the rather inadequate chemical descriptions of the phases
used in even the relatively recent original literature on these materials, from
which no clear information on the role of water can be deduced.

3.6.3 Water as plasticizer


An additional possible role for water in AB cements is plasticization.
Water is known to act as plasticizer for a number of polymeric materials,
whether synthetic or natural, and whether or not they are predominantly

51
Water and acid-base cements
polar. Thus water has been found to affect the properties of poly(methyl
methacrylate) (Turner, 1982) and of alkyd resins used in surface coatings
(Mayne & Mills, 1982). As plasticizer, the principal effect of water in these
systems is to reduce the glass transition temperature, Tg, and this in turn
affects a number of other properties of the materials, including rigidity,
dimensional stability and diffusion coefficients within the bulk. Given the
polar nature of the components of AB cements, the known water content
of set cements, and the fact that water has been shown to act as plasticizer
in pure metal poly(acrylate) salts (Yokoyama & Hiraoko, 1979) it seems
probable that one of the roles of water in the solid state of AB cements is
that of plasticizer.

References
Barrow, G. M. (1979). Physical Chemistry, 4th edn. Tokyo: McGraw-Hill
Kogakusha Ltd.
Benedict, W. S., Gailar, N. & Plyler, E. K. (1956). Rotation-vibration spectra of
deuterated water vapour. Journal of Chemical Physics, 24, 1139-65.
Bernal, J. D. & Fowler, R. H. (1933). A theory of water and ionic solutions with
particular reference to hydrogen and hydroxyl ions. Journal of Chemical
Physics, 1, 515-48.
Bollinger, J. C , Faure, R., Yvernault, T. & Stahl, D. (1987). On the existence of
the protonated dication H4 O2+ in sulfolane solution. Chemical Physics Letters,
140, 579-81.
Bragg, W. H. (1922). The crystal structure of ice. Proceedings of the Physical
Society of London, 34, 98-103.
Briant, C. L. & Burton, J. J. (1976). Molecular dynamics study of the effects of
ions on water microclusters. Journal of Chemical Physics, 64, 2888-95.
Clever, H. L. (1963). The hydrated hydronium ion. Journal of Chemical
Education, 40, 637^41.
Cunningham, A. J. C , Payzant, J. D. & Kebarle, P. (1972). A kinetic study of
the proton hydrate H+(H2O) and equilibria in the gas phase. Journal of the
American Chemical Society, 94, 7627-32.
Earl, M. S. A., Hume, W. R. & Mount, G. J. (1985). Effect of varnishes and
other surface treatments on water movement across the glass-ionomer cement
surface. Australian Dental Journal, 30, 298-301.
Eisenberg, D. & Kauzmann, W. (1969). The Structure and Properties of Water.
Oxford: Oxford University Press.
Enderby, J. E., Cummings, S., Herdman, G. H., Neilson, G. W., Salmon, P. S.
& Skipper, N. (1987). Diffraction and study of aqua ions. Journal of Physical
Chemistry, 91, 5851-8.
Frank, H. S. & Evans, M. W. (1945). Entropy in binary liquid mixtures; partial
molal entropy in dilute solutions; structure and thermodynamics in aqueous
electrolytes. Journal of Chemical Physics, 13, 507-32.

52
References

Frank, H. S. & Wen, W-Y. (1957). Structural aspects of ion-solvent interactions


in aqueous solutions-water structure. Discussions of the Faraday Society, 24,
133^0.
Franks, F. (1975). The hydrophobic interaction. In Franks, F. Water. A
Comprehensive Treatise, vol. 4, Chapter 1. London and New York: Plenum
Press.
Franks, F. (1983). Water. London: Royal Society of Chemistry.
Gillespie, R. J. & Nyholm, R. S. (1957). Inorganic stereochemistry. Quarterly
Reviews of the Chemical Society, 11, 339-80.
Greenwood, N. N. & Earnshaw, A. (1984). The Chemistry of the Elements.
Oxford: Pergamon Press.
Gurney, R. W. (1953). Ionic Processes in Solution. New York: McGraw-Hill.
Hall, V. M. D. (1985). In Russell, C. A. (ed.) Recent Developments in the History
of Chemistry. London: Royal Society of Chemistry.
Hart, E. J. & Anbar, M. (1970). The HydratedElectron. New York: Wiley.
Heinzinger, K. & Palinkas, G. (1987). In Kleeberg, H. (ed.) Interactions of
Water in Non-ionic Hydrates. Berlin: Springer-Verlag.
Heinje, G., Luck, W. A. P. & Heinzinger, K. (1987). Molecular dynamics
simulation of an aqueous sodium perchlorate solution. Journal of Physical
Chemistry, 91, 331-8.
Hornsby, P. R. (1980). Dimensional stability of glass-ionomer cements. Journal
of Chemical Technology and Biotechnology, 30, 595-601.
Hiickel, W. (1950). Structural Chemistry of Inorganic Compounds, vol. 1. New
York: Elsevier.
Ikegami, A. (1968). Hydration of polyacids. Biopolymers, 6, 431-40.
Ikegami, A. & Imai, N. (1962). Precipitation of polyelectrolytes by salts. Journal
of Polymer Science, 56, 133-52.
Kauffman, G. B. (1981). Coordination Chemistry. New York: Heyden.
Kearley, G. J., Pressman, H. A. & Slade, R. C. T. (1986). The geometry of the
H5OJ ion in dodecatungstophosphoric acid hexahydrate, (H5O£)3 (PW^O^),
studied by inelastic neutron scattering vibrational spectroscopy. Journal of the
Chemical Society Chemical Communications, 1801-2.
Kebarle, P. (1977). Ion thermochemistry and solvation from gas phase ion
equilibria. Annual Reviews in Physical Chemistry, 28, 445-76.
Kitano, T., Taguchi, A., Noda, I. & Nagasawa, M. (1980). Conformation of
polyelectrolytes in aqueous solution. Macromolecules, 13, 57-63.
Klooster, N. Th. M., van der Trouw, F. & Mandel, M. (1984). Solvent effects in
polyelectrolyte solutions. 3. Spectropho tome trie results with (partially)
neutralised poly(acrylic acid) in methanol and general conclusions regarding
these systems. Macromolecules, 17, 2087-93.
Kochanski, E. (1985). Theoretical studies of the system H3O+(H2O)n for
n = 1—9. Journal of the American Chemical Society, 107, 7869-73.
Marcus, Y. (1988). Ionic radii in aqueous solution. Chemical Reviews, 88,
1475-98.
Mayne, J. E. O. & Mills, D. J. (1982). Structural changes in polymer films.
Part 1. The influence of the transition temperature on the electrolytic resistance

53
Water and acid-base cements

and water uptake. Journal of the Oil and Colour Chemists' Association, 65,
138^2.
McMillan, W. G. & Mayer, J. E. (1945). The statistical thermodynamics of
multicomponent systems. Journal of Chemical Physics, 13, 276-305.
Molyneux, P. (1975). Synthetic polymers. In Franks, F. Water. A Comprehensive
Treatise, vol. 4, Chapter 7. London and New York: Plenum Press.
Moore, W. J. (1972). Physical Chemistry, 5th edn. London: Longman Group
Ltd.
Morrison, R. T. & Boyd, R. T. (1973). Organic Chemistry, 3rd edn. New York:
Allyn and Bacon.
Neilson, G. W., Schioberg, D. & Luck, W. A. P. (1985). The structure around
the perchlorate ion in concentrated aqueous solutions. Chemical Physics
Letters, 111, 475-9.
Nicholson, J. W. (1985). Waterborne Coatings. OCCA Monograph No. 3.
London: Oil and Colour Chemists' Association.
Paddon, J. M. & Wilson, A. D. (1976). Stress relaxation studies on dental
materials. 1. Dental cements. Journal of Dentistry, 4, 183-9.
Pauling, L. (1960). The Nature of the Chemical Bond, 3rd edn. Ithaca, New
York: Cornell University Press.
Pearson, R. G. (1963). Hard and soft acids and bases. Journal of the American
Chemical Society, 85, 3533-9.
Prosser, H. J. & Wilson, A. D. (1979). Litho-ionomer cements and their
technological applications. Journal of Chemical Technology and Biotechnology,
29, 69-87.
Ramanathan, P. S. & Friedman, H. L. (1971). Refined model for aqueous 1-1
electrolytes. Journal of Chemical Physics, 54, 1086-99.
Robinson, R. A. & Stokes, R. H. (1955). Electrolyte Solutions. London:
Butterworth Scientific Publications.
Robinson, G. W., Thistlethwaite, P. J. & Lee, J. (1986). Molecular aspects of
ionic hydration reactions. Journal of Physical Chemistry, 90, 4224-33.
Robb, I. D. (1983). Polymer-small molecule interactions. In Finch, C. A. (ed.)
Chemistry and Technology of Water Soluble Polymers. New York: Plenum
Press.
Russo, S. O. & Hanania, G. I. H. (1989). Ion association solubilities and
reduction potentials in aqueous solution. Journal of Chemical Education, 66,
148-53.
Salmon, P. S., Neilson, G. W. & Enderby, J. E. (1988). The structure of Cu2+
aqueous solutions. Journal of Physics C, 21, 1335-49.
Sorrell, C. A. & Armstrong, C. R. (1976). Reactions and equilibria in
magnesium oxychloride cements. Journal of the American Ceramic Society, 59,
51-4.
Sorrell, C. A. (1977). Suggested chemistry of zinc oxychloride cements. Journal
of the American Ceramic Society, 60, 217-20.
Speakman, J. C. (1975). The Hydrogen Bond. London: Chemical Society.
Symons, M. C. R. (1989). Liquid water- the story unfolds. Chemistry in Britain,
25, 491-4.

54
References

Turner, D. T. (1982). Poly(methyl methacrylate) plus water. Sorption kinetics


and volumetric changes. Polymer, 23, 197-202.
Vaslow, F. (1963). The orientation of water molecules in the field of an alkali
ion. Journal of Physical Chemistry, 67, 2773-6.
Wall, F. T. & Drenan, J. W. (1951). Gelation of polyacrylic acid by divalent
ions. Journal of Polymer Science, 1, 83-8.
Watts, R. O., Clementi, E. & Fromm, J. (1974). Theoretical study of the lithium
fluoride molecule in water. Journal of Chemical Physics, 61, 2550-5.
Weast, R. C. (ed.) (1985-6). Handbook of Physics and Chemistry. Ohio:
Chemical Rubber Company.
Wilson, A. D. & Crisp, S. (1974). Unpublished data cited in Wilson & McLean,
1988.
Wilson, A. D. & Crisp, S. (1975). Ionomer cements. British Polymer Journal, 1,
279-96.
Wilson, A. D., Crisp, S. & Paddon, J. M. (1981). The hydration of a
glass-ionomer (ASPA) cement. British Polymer Journal, 13, 66-70.
Wilson, A. D., Kent, B. E., Batchelor, R. F., Scott, B. G. & Lewis, B. G. (1970).
Dental silicate cements. XII. The role of water. Journal of Dental Research,
49, 307-14.
Wilson, A. D. & McLean, J. W. (1988). Glass-ionomer Cement. Chicago,
London, etc.: Quintessence Publishers.
Wilson, A. D. & Mesley, R. J. (1968). Dental silicate cements. VI. Infrared
studies. Journal of Dental Research, 47, 644—52.
Yokoyama, T. & Hiraoko, K. (1979). Hydration and thermal transition of
poly(acrylic acid) salts. Polymer Preprints of the American Chemical Society,
Division of Polymer Chemistry, 20, 511-13.

55
4 Polyelectrolytes, ion binding
and gelation

4.1 Polyelectrolytes
4.1.1 General
The setting of AB cements is an example of gelation, and gelation is related
to ion binding. A theoretical examination of the various phenomena
associated with ion binding and gelation finds its clearest exposition in the
field of polyelectrolytes. Moreover, this field may be wider than it seems at
first.
Polyelectrolytes form the basis of those modern cements which are
distinguished by their ability to adhere to reactive surfaces. At present the
main use of such cements lies in the medical field, principally in dental
surgery. They adhere permanently to biological surfaces where they have
to withstand adverse conditions of wetness, chemical attack, the stress of
biological activity, and chemical and biological changes within the
substrate. Nevertheless, adhesive bonds are maintained.
Polyelectrolytes are polymers having a multiplicity of ionizable groups.
In solution, they dissociate into polyions (or macroions) and small ions
of the opposite charge, known as counterions. The polyelectrolytes of
interest in this book are those where the polyion is an anion and the
counterions are cations. Some typical anionic polyelectrolytes are depicted
in Figure 4.1. Of principal interest are the homopolymers of acrylic acid
and its copolymers with e.g. itaconic and maleic acids. These are used in the
zinc polycarboxylate cement of Smith (1968) and the glass-ionomer
cement of Wilson & Kent (1971). More recently, Wilson & Ellis (1989) and
Ellis & Wilson (1990) have described cements based on polyphosphonic
acids.
There is also the question of whether there are inorganic polyelectrolytes
within the field of AB cements. A number of cements are based on
orthophosphoric acid, and in the two most important ones aluminium is

56
Polyelectrolytes
known to be essential for cement formation. Aluminium forms complex
aluminophosphoric acids with orthophosphoric acid. The solutions of
these complexes are markedly viscous and there is some NMR spec-
troscopic evidence that these aluminophosphoric acids form linear
polymers based on the Al-O-P linkage (Sveshnikova & Zaitseva, 1964;
Akitt, Greenwood & Lester, 1971; O'Neill et al., 1982). Callis, Van Wazer
& Arvan (1954), Salmon & Wall (1958) and Wilson et al. (1972) consider
that aluminophosphate polymers are formed in which [POJ tetrahedra are
linked by aluminium atoms.
Polyanion chains containing many linked charged groups exert a
considerable electrostatic effect on the orientation of dipolar solvent
molecules and on the counterions. The counterions are constrained to
remain in the neighbourhood of the charged polymer chains, a phenom-
enon known as ion binding. This phenomenon is supported by a wealth of
experimental evidence (Morawetz, 1975; Wilson & Crisp, 1977; Rymden
& Stilbs, 1985a, b) and an early illustration of it is found in the work of
Huizenga, Grieger & Wall (1950a, b) who observed that, in an electric field,
cations were sometimes transported with the polyanion.

r
CH—COOH
CH2

C H — COO"
1
2
T
CH—COOH
CH2

C H — COO~

Poly (acrylic acid)


1
Polyacrylate

CH2
1
CH 2
1 1
CH—PO(OH) 2

1
CH—SO2OH
CH2

CH—SO 2 OH
T2
CH—PO(OH) 2

Poly (vinyl sulphonic acid) Poly (vinyl phosphonic acid)


Figure 4.1 Some typical anionic polyelectrolytes.

57
Polyelectrolytes, ion binding and gelation

4.1.2 Polyion conformation


The shape, configuration or morphology of a polyion is usually known as
its conformation. There are very many possible conformations available to
a polyion because of the flexibility of the main chain due to the free
rotation of bonds. The particular conformation adopted will be the one
with the lowest free energy. This free energy has two components, one
arising from chainflexibilityand the other from electrical interactions. The
intrinsic free energy of rotation is a function of the relative position of
neighbouring bonds. There are energy minima, at the trans position, which
corresponds to the stretched form of the chain, and at the two gauche
positions, corresponding to the contracted form. The difference in energy
between the trans and gauche positions is one important factor determining
theflexibilityof the chain. The other component of free energy arises from
interactions between charged groups on the polyion, counterions and
solvent molecules.
There are two broad kinds of polyion conformation; the random coil
and the ordered helix. In a helix there are regularly repeated structures
along the coil; there are none in the case of a random coil. In this book we
are concerned with the latter where there are often several conformations
with approximately equal free energies and, thus, conformational changes
occur readily.
Random coil conformations can range from the spherical contracted
state to the fully extended cylindrical or rod-like form. The conformation
adopted depends on the charge on the polyion and the effect of the
counterions. When the charge is low the conformation is that of a
contracted random coil. As the charge increases the chains extend under
the influence of mutually repulsive forces to a rod-like form (Jacobsen,
1962). Thus, as a weak polyelectrolyte acid is neutralized, its conformation
changes from that of a compact random coil to an extended chain. For
example poly(acrylic acid), degree of polymerization 1000, adopts a
spherical form with a radius of 20 nm at low pH. As neutralization
proceeds the polyion first extends spherically and then becomes rod-like
with a maximum extension of 250 nm (Oosawa, 1971). These pH-
dependent conformational changes are important to the chemistry of
polyelectrolyte cements.
The situation is more complex in the reactions found in AB cements
because neutralization is accompanied by ion binding. Although a polyion
chain extends as the number of ionized groups increases, the binding of

58
Ion binding

counterions has the reverse effect because intrachain repulsive forces are
decreased. An increase in the concentration of polyions in solution has the
same effect, for an increase in interchain repulsion will inhibit the
unwinding of polymer chains. Thus, the effects predicted by dilute solution
theory will be much less in the concentrated conditions found in AB
cements. From this it can be seen that the effect of ion binding on
conformation change is complex.
Conversely, conformation affects the binding of counterions to polyions
(Jacobsen, 1962). In the compact spherical conformation some ionized
groups on polymer chains will be inaccessible for ion binding.

4.2 Ion binding


4.2.1 Counterion binding
Oppositely charged ions are attracted to each other by electrostatic forces
and so will not be distributed uniformly in solution. Around each ion or
polyion there is a predominance of ions of the opposite charge, the
counterions. This cloud of counterions is the ionic atmosphere of the
polyion. In a dynamic situation, the distribution of counterions depends
on competition between the electrostatic binding forces and the opposing,
disruptive effects of thermal agitation.
The phenomenon has been studied by a number of techniques: titration
(Gregor & Frederick, 1957; Kagawa & Gregor, 1957); viscosity and
electrical conductance measurements (Gregor, Gold & Frederick, 1957;
Bratko et al., 1983); determination of counterion activity (Kagawa &
Katsuura, 1955); measurement of transference (Ferry & Gill, 1962);
dilatometry (Strauss & Leung, 1965; Begala & Strauss, 1972); and NMR
spectroscopy (Rymden & Stilbs, 1985a, b).
Ion binding is affected by the size and charge of the counterion, the
charge and conformation of the polyion, and states of hydration. We will
examine these effects in some detail.

4.2.2 The distribution of counterions


The potential distribution around the polyion is important to any
discussion of counterion binding and hydration effects. Oosawa (1971) has
distinguished four regions of potential about a polyion (Figure 4.2): (1) a
localized potential hole around each charged group, (2) a cylindrical
potential valley or tube along the polyion chain, (3) a spherical trough in

59
Poly electrolytes, ion binding and gelation

the apparent volume occupied by the whole of the coiled chain, and (4) the
region outside the polyion. Counterions are distributed between these four
potential regions and may be classified as free, bound but mobile
(atmospheric) and localized {site-bound). Free ions remain outside the
volume of the polyion (in region 4); the remaining ions are bound to the
polyion. Of the bound ions, the mobile atmospheric ions occupy the
potential trough or valley around each polyion (regions 2 and 3). Localized
binding occurs in the potential holes at the sites of the individual charged
groups of the polyion, and ion-pairs are formed.
Oosawa (1971) used a simple calculation to illustrate the effect of a
highly charged polyion on the binding counterions. The distribution
between free ions and bound ions depends on the ratio between potential
energy and kinetic energy. In the case of a random coil, containing n
ionized groups of charge — e0, and of spherical conformation, radius/?, the
potential drop, Si//, for a counterion of charge + e0 at the edge of the
polyion is given by
neo/eop (4.1)

\
\

r
Region of
free counterions

Figure 4.2 The four regions of potential about a polyion. Based on Oosawa (1971).

60
Ion binding

where the dielectric constant of the solvent is equal to e0. Hence the
potential energy is
nel/eop (4.2)
The ratio of the potential energy to the kinetic energy, kT, is
ne*/eopkT (4.3)
This ratio is related linearly to the degree of polymerization n. In the case
of a poly(acrylic acid) where n = 1000 and p = 20 nm, this ratio works out
at 35. Thus, many of the counterions must enter the region of the polyion.
Even when 90 % of the counterions are within the polyion this ratio is still
high with a value of 3-5. A similar calculation for the rod-like random coil
gives an energy ratio of 26 and similar arguments apply.
Oosawa (1971) developed a simple mathematical model, using an
approximate treatment, to describe the distribution of counterions. We
shall use it here as it offers a clear qualitative description of the
phenomenon, uncluttered by heavy mathematics associated with the
Poisson-Boltzmann equation. Oosawa assumed that there were two
phases, one occupied by the polyions, and the other external to them. He
also assumed that each contained a uniform distribution of counterions.
This is an approximation to the situation where distribution is governed by
the Poisson distribution (Atkins, 1978). If the proportion of site-bound
ions is negligible, the distribution of counterions between these phases is
then given by the Boltzmann distribution, which relates the population
ratio of two groups of atoms or ions to the energy difference between them.
Thus, for monovalent counterions

nJK = (nJV^xpi-e^/kT) (4.4)


where Sy/ is the average potential difference between the two phases, nh is
the number of bound ions contained in a total volume Vh9 nt is the number
of free ions contained in a total volume Vt and n is the total number of
counterions in a total volume V.
This case can be rewritten

In {(1-/?)/# = ln{0/(l-0)}-e o <ty/£r (4.5)


where p is the apparent dissociation constant, i.e. the ratio of free to total
counterions, nt: n, and 6 is the volume concentration of the polyion, Vh: V.
For a rod-like or cylindrical polyion, the potential difference Sy/ between

61
Polyelectrolytes, ion binding and gelation
the inside and outside of a cylindrical polyion of length / and radius r, with
an average distance between the polyions of 2R, is given by
dy, = -2(nie0/s0)\n(R/r) (4.6)
If v is the mean effective volume occupied by a single polyion and
N is the number of polyions then v = nrH and V/N = nR2l; thus
r*/R2 = Nv/V=0 and
(4.7)

Hence equation (4.5) becomes


= ln{0/(l-0)}-/fein0 (4.8)
where Q is the charge density along the polyion and equals nel/sokTl.
The equation for a spherical polyion conformation of radius a is similar:
ln{(l -fi)/fi = In{0/(1 -0)}-/?P(l -0 1/3 ) (4.9)
where P is the charge density along the polyion and equals nel/skTa.
Thus /?, and hence the extent of the ion binding, depends both on the
volume concentration of the polyion, 9, and the charge density, Q. The
consequences of these equations, are not easy to see, because they cannot

1.0

0 =1

0.5 •
Q=2

Q=3

1 I I
0.1 0.2 0.3
Figure 4.3 Variation of the apparent degree of dissociation, /?, with v and Q. Based on
Oosawa (1971).

62
Ion binding

be solved. However, when 9 is sufficiently small the equations can be


simplified and rendered easy to discuss. Although the conditions in very
dilute solutions are far removed from practical reality, the simplified
situation can be used to illustrate certain basic points. Thus, for a rod-like
configuration where 9^0, equation (4.8) reduces to
In {(1-/?)//?} = ( l - / ? 0 In 0 (4.10)
or
(\-P)/P=8a-pQ) (4.11)
The apparent degree of dissociation, /?, varies in a complex way with 9
and depends on the value of Q (Figure 4.3). Apart from the case Q = 1,
where /? decreases with 9, fi shows little variation with 9. It slowly decreases
when Q = 2, and when Q = 3 or 4 it increases slightly to a plateau.
Consequently, in practical cases, /? is unaffected by increases in the charge
on the polyion associated with ionization. This conclusion is supported by
the results of Nagasawa & Kagawa (1957).

4.2.3 Counterion condensation


The theory of counterion condensation is implicit in Oosawa (1957) but the
term was coined later (Imai, 1961). The phenomenon was demonstrated by
Ikegami (1964), using refractive index measurements of the interaction
between sodium and polyacrylate ions. It has since been confirmed for
many mono-, di- and trivalent counterions and polyionic species
(Manning, 1979).
Manning (1969) suggested that there is a critical charge density above
which counterions condense on the surface of the polyion. This phenom-
enon is most clearly illustrated by the simple case of infinite dilution. As
9^0 in equation (4.11), the graph of j$ against Q falls into two parts
about the critical point Q = 1:
P^\ andy->l for Q ^ 1 (4.12)
£-> \/Q and y-> \/Q for Q ^ 1 (4.13)
where y is the activity coefficient and equals /?/(l — 0).
The consequences of these solutions are shown in Figure 4.4. The
abscissa is n, the total number of counterions or charged groups on the
polyion, and is proportional to Q. Along the ordinate are the numbers of
counterions bound, nh, and free, nt, equal to n(\—fi) and n^respectively.
The increase of counterion binding with the charge on the polyion has

63
Polyelectrolytes, ion binding and gelation
been termed counterion condensation as it is analogous to the con-
densation of a vapour. This point is illustrated by Figure 4.4. As the charge
on the polyion increases from zero there is a proportional increase in the
number of counterions. At first, all the counterions are free and none
are bound. This continues until a plateau is reached at a critical value of
Q = 1. Above this point, all additional counterions are bound to the
polyion and the number of free ions remains constant. Thus, PQ remains
constant and /? decreases as Q increases.
The simple situation depicted in Figure 4.4 is a limiting one, and the
discontinuity does not occur when 6 > 0; however, for large values of g,
PQ increases only slowly. Nor does the discontinuity appear in the case of
the spherical conformation, but again, for large values of P, PP increases
only logarithmically. Thus, the situation is similar to that for the rod-like
configuration although there is no specific critical value for P.
The above treatment is based on the assumption that 9 is small.
However, as Figure 4.3 shows, /? does not greatly change with con-
centration so that counterion condensation is probably insensitive to
concentration. The delayed binding of counterions is of some importance
to the onset of gelation.

bound /
counterions, /
n
a b 7
H

/
C free /
counterions, /
r—t
n
f /
a /
/
/ :Q = i /
c /
/
/
/
/ :
/ •
/ ;

Total number of counterions, n[Q]


Figure 4.4 The abscissa is n, the total number of counterions or charged groups on the
polyion, and is proportional to Q. Along the ordinate are the number of counterions bound,
nh, and free, «f, equal to n(\ —f$) and «/? respectively. Based on Oosawa (1971).

64
Ion binding

Theories of counterion condensation have been reviewed by Manning


(1979, 1981); and Satoh, Komiyama & Iijima (1984) have extended the
theory.

4.2.4 Effect of valence and size on counterion binding


Cations of small ionic radius and high charge are more firmly bound than
monovalent ions of large ionic radius (Ikegami & Imai, 1962; Strauss &
Leung, 1965; Begala & Strauss, 1972). Divalent ions are more strongly
bound than monovalent ions and the interaction is often localized. This
can be examined theoretically by applying Oosawa's two-phase model to
counterions with a valence z and charge of + e0, and a polyion with a total
charge — ne.
Equations (4.5) and (4.8), which were developed for univalent ions, can
be rewritten, thus:

In {(1-/?)//?} = \n{e/(\-e)}-zeQ5¥/kT (4.14)


(4.15)

The critical value for Q is 1/z. There is a proportional increase in the


number of free counterions, njz, as Q increases from zero, reaching a
plateau when Q = 1/z. Also, below this value the degree of dissociation, /?,
increases as the concentration decreases, and tends to unity as v tends to
zero. When Q > 1/z, f5 decreases with 9 and tends to 1/zQ as 6 tends to
zero. The number of free ions cannot exceed n/z2Q. Note that this number
is inversely proportional to the square of the valence. The condensation of
ions is thus very sensitive to valence; for multivalent counterions it takes
place at a lower value of Q and the number of free ions is much smaller
(l/* a ).
Imai (1961) has observed that multivalent counterions are more strongly
bound than are monovalent ones. This phenomenon can be demonstrated
theoretically by considering equilibrium conditions for two counterions
with valencies zx and z2 (z2 > zx) and degrees of dissociation fix and /?2.
For a cylindrical model the equilibrium equations are
^-^/.A)^ (4.16)
fflwJi+fMW (4.17)

where fx and/ 2 represent the proportions of the total charge carried by the
counterions, i.e./i+/ 2 = 1.

65
Polyelectrolytes, ion binding and gelation

As 9^0 then the solutions to these equations fall into four groups
depending on the value of Q.
A->1, &"•! forg<l/z 2 (4.18)
z for l z l z
&^1, A^l//2 2 2 - / i / / 2 / 2 < Q < /fi 2 (4.19)
Px-> 1, p2->0 for I/ft z2 ^ Q ^ I/ftz x (4.20)
P1->l/f1z1Q, 02^O for I / f t z ^ e (4.21)
These equations are represented graphically in Figure 4.5. Increase in the
binding of counterions as Q increases is reflected as a decrease inftvalues.
No binding occurs until Q reaches l/z 2 , when the binding of the higher-
valence ions begins. This process is complete when Q attains a value of
l/ftz 2 . There is no further binding of counterions until Q reaches l// 1 z 1
when the binding of the lower-valence ions commences. Figure 4.5 shows
that when there is a mixture of counterions then those of the higher valence
are preferentially bound. Lower-valence ions can completely suppress the
dissociation of those of higher valence.

Figure 4.5 The effect of Q on the dissociation {fix /?2) of ions of two valencies. Note the
suppression of the dissociation ( / y of ions of higher valence zx by those of lower valence z2.
Based on Oosawa (1957).

66
Ion binding

The selective binding of cations is not as sensitive to size as to valence.


The value of Q for the condensation of counterions of the same valence is
unaffected. In the case of monovalent cations, the dissociation of all
counterions is complete at infinite dilution, when Q ^ 1. When Q ^ 1 the
dissociation of the smaller counterion is always greater than that of the
larger one and increases relatively as Q increases.
A number of workers have observed that the strength of binding of
monovalent counterions depends on ionic radius. However, the effect of
ionic radius is somewhat obscure as it depends on hydration phenomena
and whether the size of the bare ion or that of the hydrated ion is the
significant parameter (Wilson & Crisp, 1977).

4.2.5 Site binding - general considerations


Not all ions are mobile within the ionic atmosphere of the polyion. A
proportion are localized and site-bound-a concept apparently first
suggested by Harris & Rice (1954). Localized ion binding is equivalent to
the formation of an ion-pair in simple electrolytes. Experimental evidence
comes mainly from studies on monovalent counterions.
This concept is due to Bjerrum, who in 1926 suggested that in simple
electrolytes ions of the opposite charge could associate to form ion-pairs
(Szwarc, 1965; Robinson & Stokes, 1959). This concept of Bjerrum arose
from problems with the Debye-Hiickel theory, when the assumption that
the electrostatic interaction was small compared with kT was not justified.
Bjerrum considered the case of spherical ions in a solvent of dielectric
constant e. The probability of finding two ions of opposite charge at a
distance A from each other is calculated from the number of ions
surrounding a central ion of opposite charge in a spherical shell of
thickness dA and radius A. This probability, W{A), is given by
W(A) dA = (4nA2 dA/v) exp (e2/sAkT) (4.22)
for monovalent ions. This distribution has a minimum, Am, at
Am = e'/2ekT (4.23)
For a cation of charge z+ and anion of charge z_, this minimum becomes
Am = z+z_e2/2ekT (4.24)
When A > Am the ions are free and the Debye-Hxickel theory applies.
When A < Am the two ions tend to approach each other and form an ion-
pair, and there is no contribution to the electrostatic energy from the
interaction between an ion and its atmosphere.

67
Poly'electrolytes, ion binding and gelation

The high dielectric constant of water normally militates against the


formation of ion-pairs for simple salts because a high dielectric constant
reduces the strength of the electrostatic forces. The phenomenon is more
readily observed in solvents of low dielectric constant; for a typical mono-
monovalent salt, ion-pair formation takes place only when the dielectric
constant is less than 41 (Fuoss & Kraus, 1933).
The fraction of all ions forming ion-pairs is

W(A)dA (4.25)
J2c

where a is the radius of the central ion.


This distribution has some inconsistencies - for example it diverges
when R is large - and was modified by Fuoss (1934); see Figure 4.6.
These arguments for simple electrolytes can be extended to the
relationship between the two types of bound counterion in poly-
electrolytes: the bound but mobile (atmospheric) and the localized
(site-bound). Under equilibrium conditions, the relationship between site-
bound and atmospheric ions is
(4.26)

2
I

Con tact Ion-pair


distance range
Inter-ion distance
Figure 4.6 The Fuoss (1934) distribution function.

68
Ion binding
where ns is the number of site-bound ions, «a is the number of atmospheric
ions and K is the equilibrium constant. For monovalent cations in dilute
solution (0-1 M) the degree of localized ion binding is negligible; in more
concentrated solutions some site binding does occur. In general, localized
ion binding can be expected only with multivalent cations.
When site binding occurs, the equations which relate the numbers of free
and bound ions require some modification. The relationship is then
between free ions and those bound ions that are mobile. The equations are
similar to equations (4-8), (4-9), (4-14), and (4-15), but the number of site-
bound ions has to be discounted in all calculations for P, Q, /? etc.

4.2.6 Effect of complex formation


In a discussion of papers by Rice & Harris (1954) and Harris & Rice (1954),
Van Wazer (1954) suggested that there could be covalent binding as well as
electrostatic interaction and that cations could be held at specific sites by
complex formation.
This is a reasonable inference, because site binding is significant only
with multivalent cations and strong electrostatic interactions. Under these
conditions ion polarization occurs and bonds have some covalent character
(Cotton & Wilkinson, 1966). This is illustrated by the data of Gregor,
Luttinger & Loebl (1955a,b). They measured the complexation constants
of poly(acrylic acid), 0-06 N in aqueous solution, with various divalent
metals, which, as it so happens, are of interest to AB cements (Table 4.1).
The order of stability was found to be

Mg < Ca < Co < Zn < Mn < Cu


Mandel & Leyte (1964) found a similar order for the complexes of
poly(methacrylic acid):
Mg < Co < Ni < Zn < Cd < Cu
Some of these divalent cations form part of the Irving-Williams series:
Mn, Fe, Co, Ni, Cu and Zn. Irving & Williams (1953) examined the
stability constants of complexes of a number of divalent ions and found
that the order
Mn < Fe < Co < Ni < Cu > Zn
held for the stability of most complexes irrespective of the nature of the
coordinated ligand. The stability constants of metal-poly(alkenoic acid)

69
Polyelectrolytes, ion binding and gelation
Table 4.1. Metal PAA complexes (Gregor, Luttinger & Loebl,
195 5 a,b)

Metal ion Crystal ionic radius A Complexation constant

Cu 2+ 0-72 6-0 xlO 3


Mn 2+ 0-80 2-3 x 103
Zn 2+ 0-80 2-1 x 103
Co 2+ 0-72 4-0 x 102
Ca 2+ 0-99 1-OxlO2
Mg 2+ 0-66 6-0 x 101

complexes, for the most part, follow the Irving-Williams series as do the
stabilities and strengths of poly(alkenoic acid) cements.
The complexation constant for copper(II) is particularly high and Wall
& Gill (1954) have suggested that chelate formation takes place with two
carboxyl groups:

From all of this discussion it is apparent that, as Manning (1979) said, the
binding between counterion and polyion can range from atmospheric to
covalent site binding.

4.2.7 Effect of the polymer characteristics on ion binding


The extent of ion binding depends on a number of characteristics of the
polyion: degree of dissociation, acid strength, conformation, distribution
of ionizable groups and cooperative action between these groups (Wilson
& Crisp, 1977; Oosawa, 1971; Harris & Rice, 1954, 1957). The hydration
state of the macromolecule, which is in turn dependent on conformation,
also affects ion binding (Begala, 1971).

70
Ion binding

There are differences in ion binding between different polyacids. Thus,


alkali metal ions are bound more strongly to poly(acrylic acid) than to the
weaker poly(methacrylic acid) (Wilson & Crisp, 1977). Again, the ranking
order for the binding strength of alkali metal ions depends on the nature of
the polyanion, and the order is different for poly(acrylic acid) than for
poly(maleic acid) or poly(itaconic acid). Thus, for poly(acrylic acid) the
binding strength increases in descending order of the ionic radius of the
bare cation:

K+ < Na+ < Li+

For poly(maleic acid) and poly(itaconic acid), the binding strength


increases in descending order of the size of the hydrated metal ion, which
is the reverse of that for the bare ion (Muto, Komatsu & Nakagawa, 1973;
Muto, 1974). This observation has been explained by postulating the
formation of a stable ring structure with a hydrogen bridge between
ionized and non-ionized carboxylate groups.
The strength of ion binding is enhanced when the arrangements of the
functional groups permit chelate formation (Begala & Strauss, 1972).
Thus, magnesium is more firmly bound to poly(vinyl methyl ether-maleic
acid) than to either poly(acrylic acid) or poly(ethylene maleic acid).
The charge or number of dissociated groups on a poly acid chain depends
on the degree of neutralization and is reflected by the pH of the solution.
Behaviour is determined by the site binding of hydrogen ions; in the case
of a weak polyacid the number of free hydrogen ions may be neglected.
It follows that decrease of site binding of hydrogen ions is directly
proportional to the amount of added alkali. In the case of poly(acrylic
acid) its polymer chain can be regarded as a copolymer containing pendant
COOH and COO" groups, the relative amounts of each depending on the
degree of neutralization.
When the degree of neutralization is small the charge on the polyion
and the number of counterions will also be small and the majority of
counterions will be free. As the degree of neutralization, a, increases, the
polyion charge, Q, will increase. This observation follows from the
following equations:

Q = nel/eokTl (4.27)

where n = number of ionized groups on the polyion. It follows that

Q = anoel/sokTl (4.28)

71
Poly electrolytes, ion binding and gelation

where n0 = the potential number of ionizable groups on the polyion. When


Q is low, most of the counterions are free, but as neutralization increases
a point is reached at which the counterions condense; above this point,
additional counterions are bound. This follows from the discussion in
Section 4.2.3.

4.2.8 Solvation (hydration) effects


The solvation (hydration) and desolvation of ions is important to the
gelation process in AB cement chemistry. The large dipole moment of ion-
pairs causes them to interact with polar molecules, including those of the
solvent. This interaction can be appreciable. Much depends on whether the
solvent molecule or molecules can intrude themselves between the two ions
of the ion-pair. Thus, hydration states can affect the magnitude of the
interaction. The process leading to separation of ions by solvent molecules
was perceived by Winstein et al. (1954) and Grunwald (1954).
Consider two ions in contact. As they are pulled apart the potential
energy of the two ions increases. At some critical point the separation
becomes sufficient for a polar solvent molecule to occupy the space
between them, which reduces the energy of the system. Further separation
increases the energy of the system again. These changes demonstrate that
two types of ion-pair exist: contact and solvent-separated.
This distinction is meaningful if the resultant distribution function is of
the type shown in Figure 4.7 (Szwarc, 1965). This figure shows that there
is a high probability that the cation and anion are either in contact,
separated by a solvent molecule or far apart (Szwarc, 1965). Intermediate
positions are improbable. The structure of solvated ion-pairs has been
studied by Grunwald (1979) using dipole measurements.
Winstein & Robinson (1958) used this concept to account for the
kinetics of the salt effects on solvolysis reactions. They considered that
carbonium ions (cations) and carbanions could exist as contact ion-pairs,
solvated ion-pairs and as free ions and that all these forms participated in
the reactions and were in equilibrium with each other. These equilibria can
be represented, thus:

X : Y = X+Y" = X+SY~ = X+
contact solvent-separated free ions
ion-pair ion-pair

72
Ion binding

where X is a carbonium radical and X+ the carbonium ion, Y~ the


carbanion and Y the associated radical. S represents a solvent molecule.
Eigen & Tamm (1962a,b) and Atkinson & Kor (1965, 1967) envisage a
more complex situation and consider that there are two kinds of solvent-
separated ion-pairs: those with one intervening molecule of solvent and
others where the ion-pair is fully solvated (Wilson & Crisp, 1977).

4.2.9 Hydration of the polyion


The electric potential around a polyion can aflfect the structure of water.
There are three regions of potential about a polyion to consider: the
potential holes at the site of the individual charged groups, cylindrical
regions along the polymer chain, and the outlying region. In the outlying
region the potential is small and the water molecules have a normal
structure. In the other two regions there are strong electricfields,and water
molecules are oriented and have special structures. Oriented water is
denser and has a higher refractive index than normal water (Begala &
Strauss, 1972; Ikegami, 1964, 1968).
The structure of this water can be affected by ion binding. If the
counterions are tightly bound at the sites of individual charged groups, the

Solvent
ion-pair
Inter-ion distance
Figure 4.7 Distribution function for contact and solvent-separated ion-pairs.

73
Poly electrolytes, ion binding and gelation

structure of the water around them will be profoundly modified. If the


counterions are not localized but mobile, the influence on water structure
may be small. Thus, the state of the binding of a counterion will be reflected
in changes in water structure which in turn can be measured by changes in
refractive index or density.
The effect has been studied experimentally by Ikegami (1964, 1968) who
measured changes in refractive index, Asai (1961) employing an ultrasonic
method, Begala & Strauss (1972) who measured changes in molar volume,
and Grunwald (1979) using dipole moment measurements. When an acid
is neutralized by a base the refractive index of the salt solution formed is
less than the weighted mean of the refractive indices of the acid and base
solutions from which it is formed. Likewise, the density increases. By these
means, the progress of neutralization may be followed.
At low degrees of neutralization, the average distance between ionized
groups is great, so that the rearrangement of neighbouring water molecules
induced by the ionization of a carboxyl group is solely due to the charge on
that individual group. Individual hydration spheres of oriented water,
intrinsic water, are formed at each charged site. In the case of poly(acrylic
acid) when the degree of neutralization, a, is 0-3 the radius of these spheres
is 031 nm (Ikegami, 1964).
As a increases, the average distance between ionized groups decreases so
that these neighbouring groups begin to have an effect. When a exceeds 0-3,
individual water spheres begin to overlap and eventually coalesce into a
cylindrical form. With further increases in a, a second outer cylindrical
sheath of water appears in which water molecules are oriented by the
cooperative effect of two or more carboxyl groups.
When neutralization is complete, the inner layer of intrinsic water
assumes a cylindrical form along the length of the polyion with a diameter
of 0-5-0-7 nm (Ikegami, 1964). The outer second cylindrical hydration
region has a diameter of 0-9-1-3 nm (Figure 4.8).
The explanation for this volume increase is as follows. For a cylindrical
model of uniform charge density the electric field around the cylinder is
2cm0e0/e0rl (4.29)
where a is the degree of neutralization, r the radius of the cylinder and /its
length. When the magnitude of the electric field exceeds a certain value,
water molecules are reoriented; the above expression shows that the
radius, r, of the cylinder increases as the degree of neutralization, a,
increases.

74
Ion binding

According to Ikegami (1968) the presence of hydrophobic groups, for


example the methyl group in poly(methacrylic acid), can induce an
additional hydration region around neighbouring charged groups.
The arrangement of carboxyl groups on the polyacid is also important.
Thus, poly(ethylene maleic acid), PEMA, which is an 'isomer' of
poly(acrylic acid), PAA, has a different hydration structure. Whereas in
PAA the COOH groups are pendant on alternate chain C atoms, those in
PEMA are paired on adjacent chain C atoms. These structural differences
affect hydration (Begala, 1971). The separation of the hydrophilic carboxyl
groups by a pair of hydrophobic chain C atoms effectively prevents the
cooperative effect between ionizable groups. Thus, by contrast with PAA,
as the degree of ionization increases, the hydration regions around PEMA
never coalesce to form a cylindrical sheath. In the fully ionized state there
is a spherical region of intrinsic water around each carboxyl group and an
outer spherical region of water which encloses each pair of carboxyls.
The formation of a stable hydrogen-bonded ring structure as in
poly(itaconic acid) and in poly(maleic acid) has also been shown to affect
hydration states (Muto, Komatsu & Nakagawa, 1973; Muto, 1974).

0.9-1.3nm

0.5-0.7 nm

a<0.3 a=0.3 a=1.0


Figure 4.8 Cylindrical and spherical hydration regions around poly(acrylic acid) at various
degrees of neutralization (or charge densities). Based on Ikegami (1964).

75
Poly electrolytes^ ion binding and gelation

4.2.10 Hydration and ion binding


Counterions can affect the structure of hydration regions, and conversely
hydration regions can affect ion binding. We have already touched on this
subject in discussing contact and solvent-separated ion pairs in Section
4.2.8.
Large bound monovalent cations, e.g. tetrabutylammonium ions, are
too large to penetrate any of the hydration regions. However, the smaller
lithium, sodium and potassium ions are able to penetrate the outermost
hydration region of the neutralized polyacid and this is accompanied by
volume increases (Figure 4.9). These cations are probably not site-bound
but are mobile in the outer cylindrical region of hydration (Figure 4.10).
Divalent cations cause a much greater disruption of the hydration

0.5 1.0

Figure 4.9 Volume increases associated with the binding of various counterions to
poly(acrylic acid). Based on Ikegami (1964).

76
Ion binding

regions. These ions completely penetrate the outer hydration region and
partly penetrate the inner one (Figure 4.10). Such effects manifest
themselves in much greater changes in molar volume than are the case for
monovalent ions (Figure 4.9). Divalent ions may be considered to be partly
mobile and partly site-bound.
Even greater disruption is encountered in the case of trivalent cations
(Figures 4.9, 4.10). They completely penetrate both hydration regions and
destroy the structure of water around the polyion. This amounts to
complete desolvation. The same is true of bound hydrogen ions which are
localized.

4.2.77 Desolvation and precipitation


Divalent and trivalent ions can precipitate PAA, and this phenomenon is
related to the loss of a hydration region. Such precipitation is to be
distinguished from salting-out effects which occur with high concentrations
of monovalent ions.

Figure 4.10 The effect of monovalent, divalent and trivalent counterions on the hydration
state of neutralized poly (acrylic acid). Based on Ikegami (1964).

77
Poly electrolytes, ion binding and gelation

Ikegami & Imai (1962) made a study of precipitation and hydration


using turbidity, conductance, refractive index and viscosity measurements.
The following account is based on their description.
Although all divalent ions precipitate PAA when the degree of
dissociation, a, approaches 1-0, there are differences when a = 0-25 (Figure
4.11). Small amounts of barium and calcium ions precipitate PAA at this
low a value, whereas magnesium ions do not. These differences are not to
be attributed to differences in the amounts of counterions bound, for
condensation theory (Section 4.2.3) predicts that all divalent counterions
are bound to polyanions to the same extent (Imai, 1961). Therefore,
differences must arise from differences in solubility between the various
polyacrylates. At low degrees of neutralization barium polyacrylate has
low solubility, while magnesium polyacrylate is very soluble. This is related
to the extent of disruption of hydration regions as cations are bound to
polyions.
Ikegami & Imai (1962) explained their results by assuming that divalent
ions can be bound to PAA in two forms, which they represented as
(I) COO-Me+ + COO"
(II) COO-Me-OOC

1.2 -

0.9 -

0.6 •

0.3

0.25 0.50 0.75

Figure 4.11 The effect of a on the precipitation of PAA by divalent ions. Cs is the critical salt
concentration. Based on Ikegami & Imai (1962).

78
Ion binding
According to these workers the formation of COO-Me+ causes a small
degree of dehydration, while that of COO-Me-OOC is accompanied by
considerable dehydration. The experimental results showed that, when
a = 1*0, divalent ions are bound as COO-Me-OOC, a form which favours
precipitation. However, as a decreases, the COO-Me+ form becomes more
apparent.
The ratio of the two forms depends on the cation as well as on a. Ba2+
has a greater tendency to make linkages of the COO-Me-OOC type than
Mg2+ and this difference is accentuated when the density of COO~ in
the polyanion is low. Thus, at a = 0*25 more Ba2+ ions are in the
COO-Ba-OOC form than in the COO-Ba+ form, while the reverse is true
for Mg2+ ions. Moreover, the structure COO-Mg+ is more stable and
soluble than COO-Ba+ because Mg2+ is more hydrophilic than Ba2+. For
these reasons, Ba2+ is precipitated at a = 0-25 while Mg2+ is not. This
interpretation is supported by titration experiments in the presence of
divalent cations (Jacobsen, 1962). Magnesium forms very stable hydrates
and would be expected to be more difficult to desolvate.
It is, perhaps, more in line with other thinking to represent form (I) as a
solvent-separated ion-pair: COO" H2O Me2+ H2O OOC, and form (II) as
a contact ion-pair: COO~ Me2+ OOC. Thus, precipitation occurs when a
solvent-separated ion-pair is desolvated.

4.2.12 Conformational changes in polyions


The conformation of macro- or polyions has been defined and discussed
briefly in Section 4.1.1. The conformation of a polyion is determined by a
balance between contractile forces, which depend on conformation free
energy, and extension forces, which arise from electrical free energy. The
extent of conformational change is determined by several factors. Changes
are facilitated by the degree of flexibility of the polyion, and conform-
ational change is greatest at low concentration of polyions.
Conformation depends on the degree of ionization and concentration of
the polyion, the type and concentration of the counterion and the
interaction between counterion and polyion. Extension is favoured by low
concentrations of counterion and polyion. Conformational change is also
affected by the extent of the charge on the polyion. As the charge on a
polyion increases, the chain uncoils and expands under the influence of
repulsive forces. Thus, the neutralization of a polyacid is accompanied by

79
Polyelectrolytes, ion binding and gelation

chain expansion as carboxyl groups ionize. The distribution of ionized


groups is labile and depends on conformation; an extended polyion has a
larger number of ionized groups than a contracted one.
Ionization of the carboxyl groups is accompanied by binding of the
cations. But if counterions are site-bound the charge on the carboxyl
groups is neutralized and chain contraction results. A special case is that of
the polyacid which adopts a contracted form because the close association
of hydrogen ions with carboxyl groups results in a neutral chain.
Extensive forces arise from the electrical interaction between counterions
and polyions. There are two repulsive forces which act to extend a polyion.
One results from coulombic repulsion between the charged groups on the
polyion and the other from osmotic pressure of the counterions within,
which seek to increase the space in which they can move.
The coulombic force is proportional to the square of the effective charge
on the polyion, i.e. n\. (The effective charge is equivalent to the number of
free counterions, nv) When the charge along the polyion, Q, is small the
extensive forces involved are those of purely coulombic repulsion.
The most important factor determining the sensitivity of the con-
formation to the concentration of polyions is the change in ion activity or
osmotic pressure with conformation. If the activity coefficient of the
counterions is sensitive to conformation then conformational change
resulting from concentration changes of polyions becomes large.
Osmotic pressure results from the difference in concentration between
the bound but mobile counterions within the polyion and the free
counterions outside it. The concentration of counterions is greater within
the polyion so that solvent molecules tend to enter this region. The osmotic
force is proportional to the difference n — nt, where n equals the total
number of counterions or the number of ionizable groups on the polyion.
The predominant force is that of osmotic pressure, unless both the
charge density and the concentration are low. These statements may be
substantiated by a simple and approximate mathematical treatment which
applies for very dilute solutions.
Coulombic forces oc n\ oc (nflf oc (/?g)2
where ji is the degree of dissociation;
Osmotic force oc (n-nt) oc n{\~P) oc Q(\-p)
These relationships apply for both forces along and perpendicular to the
chain, although the proportionality constants differ. These simple

80
Ion binding

expressions are represented graphically in Figure 4.12. As the figure


shows, when Q is low the extensive force depends solely on coulombic
repulsion, thus when Q ^ 1, p = 1
Coulombic force oc Q2, Osmotic force = 0
At higher Q values, when Q>l,PQ=l
Osmotic force oc 1 — 1 /Q
and the Coulombic force is constant.
At high Q values the contribution from osmotic pressure predominates.
This is shown by the ratio of the two forces which is given by
Coulombic force: Osmotic force oc /?2/(l —/?)
These considerations apply to dilute solutions. In concentrated solutions
the extensive forces will be diminished. Also if the bound counterions
become site-bound then both extensive forces are diminished. These are
important factors to consider in the theory of acid-base gelation in AB
cements, where solutions are concentrated and many counterions are site-
bound.

Osmotic
force

Figure 4.12 The effect of Q on the extensive forces, coulombic and osmotic, acting on a
cylindrical and a coiled polyion. Based on Oosawa (1971).

81
Polyelectrolytes, ion binding and gelation

4.2.13 Interactions between poly ions


Repulsive coulombic forces exist between charged polyions. These are
attenuated by the bound counterions; conversely they are stronger for
polyions having a higher concentration of free counterions. When the
charge along the polyion, Q, is small the forces involved are purely
coulombic repulsion forces. However, when Q exceeds a certain value,
counterions condense on the polyions and reduce the repulsive forces.
Attractive forces arise from dipole interaction, a result of the fluctuations
in the cloud of counterions. Although the mean distribution of counterions
is uniform along the length of the polyion, there are fluctuations in the
cloud of counterions which induce transient dipoles. When two polyions
approach each other counterionfluctuationsbecome coupled and enhance
the attractive force. Since polyions have a high polarizability these
attractive forces can be considerable.
The repulsive force between polyions, calculated for the mean
equilibrium distribution of the counterions, is
ell/eQz* (4.30)

where / is the length of the polyion. The attractive forces are


kTl/D2 (4.31)
where D is the average distance between polyions. If
D<z2e20/s0kT (4.32)
then the attractive force predominates over the repulsive one. This occurs
for monovalent ions when D is 0*7 nm and for divalent ions when D is
2-8 nm. This means that attraction before direct contact between two rod-
like macroions will occur only in the case of multivalent counter ions. The
attractive force is important at high charge densities because it continues to
increase with charge density whereas the repulsive forces become constant.

4.2.14 Polyion extensions, interactions and precipitation


The precipitation of polyelectrolytes by the addition of multivalent
counterions may be explained in these terms. When there are no
multivalent ions in solution there is a strong repulsive force between
polyions and the osmotic pressure is large. The solubility of polyions is a
result of these repulsive forces.

82
Gelation
The binding of multivalent counterions decreases the repulsion and
causes attraction between polyions. This attraction is the result of the
fluctuation of the counterion distribution and is equivalent to a multivalent
counterion bridge between polyions.

4.3 Gelation
The theory of gelation (Flory, 1953,1974) has been summarized in Section
2.2.3. This theory regards gelation as the consequence of the random
crosslinking of linear polymer chains to form an infinite three-dimensional
network. The phenomenon is, of course, well illustrated by examples
drawn from the gelation of polycarboxylic acids by metal ions.
Since chemical gelation occurs only when the cations have a valency
greater than one, an early view was that it resulted from the formation of
ionic crosslinks, a concept which is useful when applying Flory's theory of
gelation. Thus, in early studies, the gelation of alginates and pectinates by
Ca2+ ions was attributed to the crosslinking of COO~ groups by Ca2+
bridge formation. Wall & Drenan (1951) had a similar view in their study
of the gelation of poly(acrylic acid) with various divalent alkaline earth
ions. However, they noted that the concentration of cations required to
produce gelation differed widely between cations and so concluded that the
phenomenon could not be explained in terms of simple ionic equilibrium.
Implicitly their mechanism assumes that chain extension occurs during ion
binding.
The concept of ionic crosslinking is in accord with the idea that a gel
must possess a coherent structure. However, although crosslinking may be
essential to gel formation it does not necessarily have to be a simple ionic
salt bridge.
Michaeli (1960) opposed these views. He concluded that whatever the
exact mechanism was, the binding of divalent cations caused contraction
and coiling of the polyelectrolyte as was the case with acids. He disagreed
with the concept of ionic crosslinking. The phenomenon of precipitation
could be explained simply in terms of reduced solubility. From this he
concluded that precipitation took place in an already coiled molecule and
the matrix consisted of spherical macromolecules containing embedded
cations.
These early views are, perhaps, too simplistic to explain in full the
rheological changes that occur in polyelectrolyte cement pastes before and
at gelation. There are several physicochemical processes that underlie

83
Poly electrolytes^ ion binding and gelation

such changes: ionization, ion binding, desolvation of the ion-pair,


conformational changes in the polymer chain and interpolyion attraction.
The extent and rate of interaction between hydrated counterion and
polyanion depends on polymer structure and conformation, acid strength,
degree of dissociation, and distribution and density of ionic charge on the
polymer chain.
The underlying physicochemical process leading to gelation in AB
cements may be summarized as follows. The interaction between metal
oxide or silicate and the polyacid solution involves a neutralization
process. As neutralization proceeds and the charge on the polymer chain
grows, the polymer chain, which is initially in random coil form, unwinds,
a process which causes the cement paste to thicken. The forces which cause
this unwinding are osmotic pressure and coulombic repulsion between the
charged groups on the polymer chain.
The cations released become bound by electrostatic forces to the
polyanionic chain. These counterions can be either mobile (atmospheric)
or site-bound at specific centres.
Ion binding reduces the repulsive forces between the charged groups on
the polyanion but, unless the counterions are site-bound, the repulsive
osmotic forces are not affected. At full neutralization the coulombic forces
along the polymer chain become zero. However, the polymer does not
contract, because the osmotic forces remain; unless, of course, all the
cations become site-bound. (Of course, in the case of a free weak acid the
concentration of mobile hydrogen ions is very small and the polymer
adopts a compact form.)
Ion binding by reduction of repulsive forces also causes the attractive
forces between polyions to increase, and the cement paste thickens. This
interaction between polyions may be regarded as a kind of bridge formed
by multivalent ions located between the polyions. At this stage the cement
paste has the characteristic of a lyophilic sol - high viscosity.
It is well known that lyophilic sols are coagulated by the removal of a
stabilizing hydration region. In this case, conversion of a sol to a gel occurs
when bound cations destroy the hydration regions about the polyanion,
and solvated ion-pairs are converted into contact ion-pairs. Desolvation
depends on the degree of ionization, a, of the polyacid, and the nature of
the cation. Ba2+ ions form contact ion-pairs and precipitate PAA when a
is low (0-25), whereas the strongly hydrated Mg2+ ion disrupts the
hydration region only when a > 0*60.
More than one type of site binding is possible. There is the simple

84
References
Bjerrum ion-pair formation based on purely coulombic attraction. There is
also complex formation and, if the ligand is bidentate, chelate formation
enhances this effect, as in the case of Cu2+ (Wall & Gill, 1954). In cement
gels Crisp, Prosser & Wilson (1976) found that the binding of Na+, Mg2+
and Ca2+ was purely ionic, whereas Al3+ binding had some covalent
character. There was a suggestion, too, that the binding of Zn2+ might not
be purely ionic. Nicholson et al. (1988a,b) found positive evidence that the
binding of Zn2+ and Al3+ involved some covalent character.
It is difficult to avoid the view, which is consistent with gelation theory,
that crosslinking is involved in gelation. Simple ionic bridgesfitin with this
view, but there are alternatives. Networks of type 3 (see Section 2.2.3) can
be formed by crosslinks consisting of bundles of chains or multistranded
helices (Flory, 1974). In gelatin, triple helices are involved, i.e. three chains
are joined at a point (Peniche-Covas et al., 1974). In alginates, gelation is
believed to result from the formation of a junction zone where there is local
chain dimerization with cavities formed capable of holding calcium ions
(Reid, 1983). This complicated junction of chain association and ion
binding is known as the 'egg-box' model.
Although this account of gelation is made with reference to organic
polyelectrolytes, it is of wider application and may be applied to
phosphoric acid cements. Orthophosphoric acid solutions used in these
cements contain aluminium, and soluble aluminophosphate complexes are
formed. Some appear to be multinuclear and there is evidence for polymers
based on the bridging Al-O-P unit. These could be termed polyelectrolytes
(Akitt, Greenwood & Lester, 1971; Wilson et al, 1972; O'Neill et al.,
1982).

References
Akitt, J. W., Greenwood, N. N. & Lester, G. D. (1971). Nuclear magnetic
resonance and Raman studies of aluminium complexes formed in aqueous
solutions of aluminium salts containing phosphoric acid and fluoride ions.
Journal of the Chemical Society, A, 2450-7.
Asai, H. (1961). Study of the hydration-dehydration in polyelectrolyte solutions
by the ultrasonic technique. Journal of the Physical Society of Japan, 16,
761-6.
Atkins, P. W. (1978). Physical Chemistry, p. 338. Oxford: Oxford University
Press.
Atkinson, G. & Kor, S. K. (1965). The kinetics of ion association in manganese
sulphate solutions. I. Results in water, dioxane-water mixtures, and
methanol-water mixtures at 25 °C. Journal of Physical Chemistry, 69, 128-33.

85
Poly electrolytes, ion binding and gelation

Atkinson, G. & Kor, S. K. (1967). The kinetics of ion association in manganese


sulphate solutions. II. Thermodynamics of stepwise association in water.
Journal of Physical Chemistry, 71, 673-7.
Begala, A. J. (1971). Interactions of cations with polycarboxylic acids. PhD
Dissertation. Rutgers University, The State University of New Jersey.
Begala, A. J. & Strauss, U. P. (1972). Dilatometric studies of counterion binding
by polycarboxylates. Journal of Physical Chemistry, 76, 254-60.
Bratko, D., Dolar, D., Godec, A. & Span, J. (1983). Electric transport in
poly(styrenesulfonate) solutions. Makromolekulare Chemie Rapid
Communications, 4, 697-701.
Bungenberg de Jong, H. G. (1949). In Kruyt, H. R. (ed.) Colloid Science II, p. 2.
Amsterdam: Elsevier Publishing Co. Inc.
Callis, C. F., Van Wazer, J. R. & Arvan, P. G. (1954). The inorganic phosphates
as polyelectrolytes. Chemical Reviews, 54, 777-96.
Cotton, F. A. & Wilkinson, G. (1966). Advanced Inorganic Chemistry, 2nd edn,
Chapter 2. New York: Wiley Inter science.
Crisp, S., Prosser, H. J. & Wilson, A. D. (1976). An infra-red spectroscopic
study of cement formation between metal oxides and aqueous solutions of
poly (acrylic acid). Journal of Materials Science, 11, 36-48.
Eigen, M. & Tamm, K. (1962a). Schallabsorption in Elektrolytlosungen als
Folge chemischer Relaxation. 1. Relaxationtheorie der mehrstufigen
Dissoziation. Zeitschrift fur Elektrochemie, 66, 93-107.
Eigen, M. & Tamm, K. (1962b). Schallabsorption in Elektrolytlosungen als
Folge chemischer Relaxation. 2. Messergebnisse und Relaxationmechanismen
fur 2-2-wertige Elektrolyte. Zeitschrift fur Elektrochemie, 66, 107-21.
Ellis, J. & Wilson, A. D. (1990). Polyphosphonate cements: a new class of
dental materials. Journal of Materials Science Letters, 9, 1058-60.
Ferry, G. V. & Gill, S. (1962). Transference studies of sodium polyacrylate
under steady state electrolysis. Journal of Physical Chemistry, 66, 999-1003.
Flory, P. J. (1953). Principles of Polymer Chemistry. Ithaca, New York: Cornell
University Press.
Flory, P. J. (1974). Introductory lecture. In Gels and Gelling Processes. Faraday
Discussions of the Chemical Society, No. 57, pp. 7-18.
Fuoss, R. M. (1934). Distribution of ions in electrolyte solutions. Transactions
of the Faraday Society, 30, 967-80.
Fuoss, R. M. & Kraus, C. A. (1933). Properties of electrolytic solutions. IV. The
conductance minimum and the formation of triple ions due to the action of
Coulomb forces. Journal of the American Chemical Society, 55, 2387-99.
Gregor, H. P. & Frederick, M. (1957). Titration studies of polyacrylic acid and
polymethacrylic acids with alkali metals and quaternary ammonium bases.
Journal of Polymer Science, 23, 451-65.
Gregor, H. P., Gold, D. H. & Frederick, M. (1957). Viscometric and
conductometric titrations of polymethacrylic acids with alkali metals and
quaternary ammonium bases. Journal of Polymer Science, 23, 467-75.
Gregor, H. P., Luttinger, L. B. & Loebl, E. M. (1955a). Metal-polyelectrolyte

86
References

complexes. I. The polyacrylic acid-copper complex. Journal of Physical


Chemistry, 59, 34-9.
Gregor, H. P., Luttinger, L. B. & Loebl, E. M. (1955b). Metal-polyelectrolyte
complexes. IV. Complexes of polyacrylic acid with magnesium, calcium,
cobalt and zinc. Journal of Physical Chemistry, 59, 990-1.
Grunwald, E. (1954). Interpretation of data obtained in nonaqueous media.
Analytical Chemistry, 26, 1696-701.
Grunwald, E. (1979). Structure of solvated ion pairs from electric dipole
moments. Journal of Pure and Applied Chemistry, 51, 53-61.
Harris, F. E. & Rice, S. A. (1954). A chain model for polyelectrolytes. I. Journal
of Physical Chemistry, 58, 725-32.
Harris, F. E. & Rice, S. A. (1957). A model for ion binding and exchange in
polyelectrolyte solutions and gels. Journal of Physical Chemistry, 58, 725-32.
Huizenga, J. R., Grieger, P. F. & Wall, F. T. (1950a). Electrolytic properties of
aqueous solutions of polyacrylic acid and sodium hydroxide. I. Transference
experiments using radioactive sodium. Journal of the American Chemical
Society, 72, 2636-42.
Huizenga, J. R., Grieger, P. F. & Wall, F. T. (1950b). Electrolytic properties of
aqueous solutions of polyacrylic acid and sodium hydroxide. II. Diffusion
experiments using radioactive sodium. Journal of the American Chemical
Society, 72, 4228-32.
Ikegami, A. (1964). Hydration and ion binding of polyelectrolytes. Journal of
the Polymer Society, A2, 907-21.
Ikegami, A. (1968). Hydration of polyacids. Biopolymers, 6, 431-40.
Ikegami, A. & Imai, N. (1962). Precipitation of polyelectrolytes by salts. Journal
of Polymer Science, 56, 133-52.
Imai, N. (1961). Interaction between polyions and low molecular weight ions.
Journal of the Physical Society of Japan, 16, 746-60.
Irving, H. & Williams, R. J. P. (1953). The stability of transition-metal
complexes. Journal of the Chemical Society, 3192-210.
Jacobsen, A. (1962). Configurational effects of binding of magnesium to
polyacrylic acids. Journal of Polymer Science, 57, 321-36.
Kagawa, I. & Gregor, H. P. (1957). Theory of the effect of counter ion size
upon titration behavior of polycarboxylie acids. Journal of Polymer Science,
23, 477-84.
Kagawa, I. & Katsuura, K. (1955). Activity of counterions in polyelectrolyte
solutions. Journal of Polymer Science, 17, 365-74.
Mandel, M. & Leyte, J. C. (1964). Interactions of poly(methacrylic acid) and
bivalent counterions. Journal of Polymer Science, A2, 2883-99.
Manning, G. S. (1969). Limiting laws and counterion condensation in
polyelectrolyte solutions. 1. Colligative properties. Journal of Chemical
Physics, 51, 924-33.
Manning, G. S. (1979). Counterion binding in polyelectrolyte theory. Accounts
of Chemical Research, 12, 443-9.
Manning, G. S. (1981). Limiting laws and counterion condensation in

87
Poly electrolytes, ion binding and gelation

polyelectrolyte solutions. 6. Theory of the titration curve. Journal of Physical


Chemistry, 85, 870-7.
Michaeli, I. (1960). Ion-binding and the formation of insoluble polymethacrylic
salts. Journal of Polymer Science, 48, 291-9.
Morawetz, H. (1975). Macromolecules in Solution, 2nd edn, Chapter 7. New
York: Wiley.
Muto, N., Komatsu, T. & Nakagawa, T. (1973). Counterion effect on the
titration behaviour of poly(maleic acid). Bulletin of the Chemical Society of
Japan, 46, 2711-15.
Muto, N. (1974). Counterion effect on the titration behaviour of poly(itaconic
acid). Bulletin of the Chemical Society of Japan, 47, 1122-8.
Nagasawa, M. & Kagawa, I. (1957). Colligative properties of polyelectrolyte
solutions. IV. Activity coefficient of sodium ion. Journal of Polymer Science,
25, 61-76.
Nicholson, J. W., Brookman, P. J., Lacy, O. M., Sayers, G. S. & Wilson, A. D.
(1988a). A study of the nature and formation of zinc polyacrylate cement
using Fourier transform infrared spectroscopy. Journal of Biomedical
Materials Research, 22, 623-31.
Nicholson, J. W., Brookman, P. J., Lacy, O. M. & Wilson, A. D. (1988b).
Fourier transform infrared spectroscopic study of the role of tartaric acid in
glass-ionomer cements. Journal of Dental Research, 67, 1450-4.
O'Neill, I. K., Prosser, H. J., Richards, C. P. & Wilson, A. D. (1982). NMR
spectroscopy of dental materials. 1.31P studies on phosphate-bonded cement
liquids. Journal of Biomedical Materials Research, 16, 39-49.
Oosawa, F. (1957). A simple theory of thermodynamic properties of
polyelectrolyte solutions. Journal of Polymer Science, 23, 421-30.
Oosawa, F. (1971). Polyelectrolytes. New York: Marcel Dekker.
Peniche-Covas, C. A. L., Dev, S. B., Gordon, M., Judd, M. & Kajiwara, K.
(1974). The critically branched state in covalent synthetic systems and the
reversible gelation of gelatin. In Gels and Gelling Processes. Faraday
Discussions of the Chemical Society, No. 57, pp. 165-80.
Reid, D. S. (1983). Ionic polysaccharides. In Wilson, A. D. & Prosser, H. J.
(eds.) Developments in Ionic Polymers-1, Chapter 6. London and New York:
Applied Science Publishers.
Rice, S. A. & Harris, F. E. (1954). A chain model for polyelectrolytes. II.
Journal of Physical Chemistry, 58, 733-9.
Robinson, R. A. & Stokes, R. H. (1959). Electrolyte Solutions, 2nd edn, Chapter
14. London: Butterworths.
Rymden, R. & Stilbs, P. (1985a). Counterion self-diffusion in aqueous solutions
of poly (aery lie acid) and poly (methacry lie acid). Journal of Physical
Chemistry, 89, 2425-8.
Rymden, R. & Stilbs, P. (1985b). Concentration and molecular weight
dependence of counterion self-diffusion in aqueous poly(acrylic acid)
solutions. Journal of Physical Chemistry, 89, 3502-5.
Salmon, J. E. & Wall, J. G. L. (1958). Aluminium phosphates. Part II. Ion-
References

exchange and pH-titration studies of aluminium phosphate complexes in


solution. Journal of the Chemical Society, 1128-34.
Satoh, M., Komiyama, J. & Iijima, T. (1984). Counterion condensation in
polyelectrolyte solutions: a theoretical prediction of the dependences on the
ionic strength and degree of polymerization. Macromolecules, 18, 1195-2000.
Szwarc, M. (1965). Ions, ion-pairs, and their agglomerates. Die
Makromolekulare Chemie, 89, 44-80 (in English).
Smith, D. C. (1968). A new dental cement. British Dental Journal, 125, 381-4.
Strauss, U. P. & Leung, Y. P. (1965). Volume changes as a criterion for site
binding of counterions by poly electrolytes. Journal of the American Chemical
Society, 87, 1476-80.
Sveshnikova, V. N. & Zaitseva, S. N. (1964). Aluminophosphates as
poly electrolytes. Russian Journal of Inorganic Chemistry, 9, 672-5.
Van Wazer, J. R. (1954). In Rice, S. A. & Harris, F. E. (1954). A chain model
for polyelectrolytes. II. Journal of Physical Chemistry, 58, 739.
Wall, F. T. & Drenan, J. W. (1951). Gelation of polyacrylic acids by divalent
cations. Journal of Polymer Science, 7, 83-8.
Wall, F. T. & Gill, S. J. (1954). Interaction of cupric ions with polyacrylic acid.
Journal of Physical Chemistry, 58, 1128-30.
Wilson, A. D. & Crisp, S. (1977). Organolithic Macromolecular Materials,
Chapters 2 & 4. London: Applied Science Publishers.
Wilson, A. D. & Ellis, J. (1989). Poly-vinylphosphonic acid and metal oxide or
cermet or glass-ionomer cements. British Patent Application 2, 219, 289A.
Wilson, A. D. & Kent, B. E. (1971). The glass-ionomer cement: a new
translucent cement for dentistry. Journal of Applied Chemistry and
Biotechnology, 21, 313.
Wilson, A. D., Kent, B. E., Clinton, D. & Miller, R. P. (1972). The formation
and microstructure of dental silicate cement. Journal of Materials Science, 1,
220-38.
Winstein, S., Clippinger, E., Fainberg, A. H. & Robinson, G. C. (1954). Salt
effects and ion-pairs in solvolysis. Journal of the American Chemical Society,
76, 2597-8.
Winstein, S. & Robinson, G. C. (1958). Salt effects and ion-pairs in solvolysis
and related reactions. IX. The //zre0-3-/>-anisyl-2-butyl system. Journal of the
American Chemical Society, 80, 169-81.

89
5 Polyalkenoate cements

5.1 Introduction
Poly(acrylic acid) and its salts have been known to have useful binding
properties for some thirty years; they have been used for soil consolidation
(Lambe & Michaels, 1954; Hopkins, 1955; Wilson & Crisp, 1977) and as
aflocculant(Woodberry, 1961). The most interesting of these applications
is the in situ polymerization of calcium acrylate added to soil (de Mello,
Hauser & Lambe, 1953). But here we are concerned with cements formed
from these poly acids.
The polyelectrolyte cements are modern materials that have adhesive
properties and are formed by the cement-forming reaction between a
poly(alkenoic acid), typically poly(acrylic acid), PAA, in concentrated
aqueous solution, and a cation-releasing base. The base may be a metal
oxide, in particular zinc oxide, a silicate mineral or an aluminosilicate
glass. The presence of a polyacid in these cements gives them the valuable
property of adhesion. The structures of some poly(alkenoic acid)s are
shown in Figure 5.1.
The polyelectrolyte cements may be classified by the type of basic
powder used to form the cement.
(1) The metal oxide cements (Section 5.6)
(2) The zinc polycarboxylate cement (Section 5.7)
(3) The mineral ionomer cements (Section 5.8)
(4) The glass-ionomer or glass polyalkenoate cement (Section 5.9)
Only two of these materials are of practical importance: the zinc
polycarboxylate cement of Smith (1968) and the glass-ionomer cement of
Wilson & Kent (1971). Both are used in dental applications and both have
been used as bone cements. The glass-ionomer cement is, perhaps, the
most versatile of all AB cements. It has many applications in dentistry: a

90
Introduction
filling material for the restoration of anterior (front) teeth, a cementing
agent for the attachment of crowns and bridges, a cavity liner and a base
under amalgams and composite resins, and a general repair material.
Outside dentistry it is marketed as a splint bandage material and as a bone
cement. It has also been considered as an underwater cement for North Sea
pipelines, as a replacement for plaster of Paris in slip casting, and as a
model material.
The invention and development of the zinc polycarboxylate and
glass-ionomer cements was brought about by a change in basic attitudes in
materials science in dentistry. This largely revolved around the necessity of
inventing materials which would adhere to tooth enamel and dentine.

Acrylic acid unit


I
CH—COOH

CH2
.COOH Itaconic acid unit
-CH 2 COOH

T
Maleic acid unit
CH—COOH

CH—COOH

7
CH 9
.COOH 3-Butene 1,2,3-tricarboxylic
XH—COOH acid unit

CH2COOH

Figure 5.1 The structure of poly(alkenoic acid)s containing acrylic, itaconic, maleic and 3-
butene 1,2,3-tricarboxylic acid units.

91
Polyalkenoate cements

5.2 Adhesion
5.2.1 New attitudes
Up to the 1950s the quality of a dental material was judged entirely by its
physical and mechanical properties (Wilson, 1991). This proved to be a
concept which hampered development. We may take the amalgam as
representing the traditional dental restorative material with all its
advantages and disadvantages. The amalgam is strong and resistant to
abrasion, but it is essentially a foreign body in the tooth, an unattractive
black mass of metal that does not bond to tooth structure. In order to
ensure its mechanical retention, cavities have to be cut which are wasteful
of sound tooth material. It does nothing for the tooth and, despite its
excellent mechanical properties, is little more than a mechanical plug.
In the late 1940s a reaction against this idea of a dental material took
place. Increasing attention was paid to problems of compatibility between
the restoration and the tooth. We now believe that a restorative should be
at one with the tooth material in all respects. It should possess identical
properties. Its thermal characteristics should be the same as those of the
tooth and its appearance should match that of the enamel. It should
provide some therapeutic action. In fact, a restorative material should no
longer be regarded as a 'filling' but as an 'enamel or dentine substitute'.

5.2.2 The need for adhesive materials


To achieve such compatibility the primary requisite is that the restorative
adheres to tooth material. This concept of adhesion is hardly to be found
in the literature of the 1920s and 1930s. For that reason wefindno attempt
at developing tooth adhesives in that period. Adhesion was, apparently,
only recognized as a desirable property in the 1950s. It seems for some
reason to be associated with the introduction of simple resins as dental
restorative materials. Although they were not a great success, attempts
were made to bond them to tooth material.
The Conference on Adhesive Restorative Dental Materials held in
Indianapolis in 1961 (Phillips & Ryge, 1961) may be considered as ushering
in the era when dental adhesives were actively sought. Buonocore (1961)
summed up the new thinking.
The lack of adhesion of available filling materials to tooth structure is
considered as one of their shortcomings. A solution to this problem
would indeed represent a milestone in dentistry.

92
Adhesion
Thus, thought became directed towards developing adhesive dental
materials, an approach that has led to considerable successes and has
revolutionized restorative dentistry.

5.2.3 Acid-etching
The first experimental study on adhesion appears in a paper by Kramer &
McLean (1952). They reported on the use of glycerol phosphoric acid
dimethacrylate as a dentine bonding agent. They achieved some success
but, unfortunately, the bond deteriorated with time (Buonocore, 1961).
More significant was thefindingreported by Buonocore in 1955. This was
his innovative technique for the acid-etching of enamel for the micro-
mechanical attachment of dental resins. Resins are capable of penetrating
an etched surface and, when polymerized, are bonded to the enamel by
resin tags. Buonocore's significant innovation proved to be far ahead of its
time because the simple restorative resins then available were not a clinical
success. Buonocore's invention remained unnoticed until the arrival of the
composite resin some ten years later. This technique has ensured the lasting
success of the composite resin. The dental surgeon now has the means to
achieve the aesthetic restoration of damaged incisal edges on anterior
teeth. Formerly such damaged teeth would have had to be crowned.
Although the importance of Buonocore's discovery cannot be over-
emphasized, micromechanical attachment cannot be regarded as true
adhesion. True adhesion must be on the molecular level and must involve
chemical or physicochemical bonds.

5.2.4 Obstacles to adhesion


There are many obstacles to permanent adhesion under oral conditions.
The substrate is a biological tissue and subject to change, and the presence
of moisture represents the worst kind of situation for adhesion. Water is
the great barrier to adhesion. It competes for the polar surface of tooth
material against any potential polymer adhesive. It also tends to hydrolyse
any adhesive bond formed. These twin obstacles gave rise to considerable
doubt as to whether materials adhesive to tooth material could be
developed at all (Cornell, 1961).
Nevertheless adhesive materials were developed, for in 1968 Dennis
Smith announced the zinc polycarboxylate cement (Smith, 1968,1969) and

93
Polyalkenoate cements
this material was followed by the glass-ionomer cement of Wilson and
Kent in 1969 (Wilson & Kent, 1971, 1972, 1973, 1974). These inventions
demonstrated that materials based on poly(acrylic acid) or similar
poly acids are effective dental adhesives. Even today, these materials are the
only ones known with certainty to form a permanent bond to tooth
material - that is a bond that does not deteriorate with time; if anything
the bond strength increases (Tyas et al., 1988).

5.2.5 The nature of the adhesion of polyalkenoates to tooth material


Dentine forms the bulk of a tooth and is covered by a harder material,
enamel. Enamel is almost entirely inorganic, with only 0-25-0-45 % protein
and 0-60% lipids (Hess, 1961). Enamel is composed of rod-shaped
structural units known as enamel prisms c. 5 jam in diameter (Silverstone,
1982). It is generally accepted that the mineral is hydroxyapatite (Posner,
1961; Silverstone, 1982), although the evidence is not entirely conclusive.
Dentine is composed of 70 % inorganic material, 20 % organic material
and 10 % water (Ten Cate & Torneck, 1982). The mineral portion is largely
a hydroxyapatite-like mineral and the organic portion is largely the protein
collagen.
The precise nature of the adhesion of the polyelectrolyte cements to
untreated dental enamel and dentine has yet to be established. The earliest
theory was due to Smith (1968) who speculated that the polyacrylate
chains of the cement formed a chelate with calcium ions contained in the
hydroxyapatite-like mineral in enamel and dentine. Beech (1973) con-
sidered this unlikely since it involved the formation of an eight-membered
ring. Beech studied the interaction between PAA and hydroxyapatite,
identified the formation of polyacrylate and so considered that adsorption
was due to ionic attraction.
Wilson (1974) emphasized the importance of wetting the substrate
surface. Later, as the reaction proceeded, these hydrogen bonds would be
replaced by ionic salt bridges. Wilson stressed the importance of the
polymeric nature of these cements in adhesion. Their polymeric nature
allowed interfacial gaps between cement and substrate to be bridged and
also provided a multiplicity of bonds. Under oral conditions, where the
substrate is subject to change, adhesive bonds will be broken, but if there
are a multiplicity of these, attachment of the cement to the substrate will
endure and allow broken bonds to be re-established. It is significant that

94
Adhesion

the related phosphate cements based on monomeric [POJ units do not


have this adhesive property.
Wilson, Prosser & Powis (1983) studied the adsorption of polyacrylate
on hydroxyapatite using infrared and chemical methods. They observed an
exchange of ions and concluded that polyacrylate displaced surface
phosphate and calcium, and entered the hydroxyapatite structure itself
(Figure 5.2). They postulated that an intermediate layer of calcium and
aluminium phosphates and polyacrylates must be formed at the cement-

0" 0 0 " o-

Hydroxy apatite
Surface

0" O 0" O
Ca 2+ Ca 2+
Ca 2+

COO"

POf

o- oo- o
Ca2+ Ca2+
Figure 5.2 The adsorption of polyacrylate on hydroxyapatite.

95
Polyalkenoate cements
hydroxyapatite interface. This layer has actually been observed by Mount
(1990) who observed debonding at the interface between this intermediate
layer and the body of the cement when the cement was dehydrated.
Adhesion in vivo appears to be dynamic. Bonding to bone was observed
to be disrupted as extensive bone remodelling took place, and then re-
established once damage had been repaired (Brook, Craig & Lamb,
1991b).

Adsorption studies
In order to elucidate the mechanism of adhesion of ionomer-carboxylate
cements, Wilson and his coworkers have carried out several studies on the
adsorption of carboxylates - aliphatic, aromatic and polymeric-on
hydroxyapatite (Skinner et ai, 1986; Scott, Jackson & Wilson, 1990; Ellis
et al9 1990).
Aliphatic monocarboxylates are not adsorbed at all (Skinner et al,
1986). The extent of the adsorption of aliphatic dicarboxylates depends on
the spacing between the carboxyl groups, and is greatest when the number
of carbon atoms in the molecule is three or four (malonate and succinate).
Adipate (six carbon atoms) is not adsorbed at all. Adsorption is, therefore,
dependent on a cooperative effect between pairs of carboxyl groups.
Adsorption cannot occur at OH sites in hydroxyapatite for these are 0-69
or 0-94 nm apart and, remembering that the length of the C-C bond lies
between 014 to 016 nm, could only be bridged by the pairs of carboxyls in
adipate, which is not adsorbed. However, attachment to H2PC>4 sites is
possible via hydrogen bonds.
Similar results were found in a study of aromatic carboxylates with one
to six carboxyl groups (Scott, Jackson & Wilson, 1990). Adsorption
increased with the number of carboxyl groups and was also dependent on
the spacing between the carboxyl groups. With the benzene dicarboxylates,
maximum permanent adsorption was obtained with the 1,3-dicarboxylate,
while the 1,4-dicarboxylates was not adsorbed at all. This is again evidence
of the cooperative effect between carboxyl groups.
Polymeric aliphatic carboxylates, the poly(alkenoic acid)s, were very
much more strongly adsorbed than the difunctional carboxylates (Ellis et
al., 1990). Results showed that adsorption depended on the conformation
of the polyanion. When extended, as in dilute solutions, a polyanion is
adsorbed onto a relatively large number of sites and further adsorption is
hindered. Thus, increases in acidity (and concentration) were found to
result in greater adsorption because the polyanion adopted a more compact

96
Preparation of poly(alkenoic acid)s
conformation. Above a certain concentration, further adsorption had a
reversible rather than an irreversible (permanent) character. High levels of
adsorption were achieved under conditions of high chain entanglement,
that is with polyanions of high molecular weight.
In all studies it was noted that calcium and phosphate ions were
displaced, the amount generally increasing with the degree of carboxylate
adsorption. It would appear that negatively charged carboxylate groups
disrupt the hydroxyapatite surface, upsetting the equilibrium between
phosphate in solid phase and solution phase thus allowing phosphate to be
exchanged by carboxylate.

5.3 Preparation ofpoly(alkenoic acid)s


The most common poly(alkenoic acid) used in polyalkenoate, ionomer
or polycarboxylate cements is poly(acrylic acid), PAA. In addition, co-
polymers of acrylic acid with other alkenoic acids - maleic and itaconic
and 3-butene 1,2,3-tricarboxylic acid - may be employed (Crisp & Wilson,
1974c, 1977; Crisp et al, 1980). These polyacids are prepared by free-
radical polymerization in aqueous solution using ammonium persulphate
as the initiator and propan-2-ol (isopropyl alcohol) as the chain transfer
agent (Smith, 1969). The concentration of poly(alkenoic acid) is kept
below 25 % to avoid the danger of explosion. After polymerization the
solution is concentrated to 40-50% for use.
Poly(alkenoic acid)s may be prepared as follows. 200 cm3 of a solution
containing between 0-5 and 2-5 g of ammonium persulphate contained in a
flask is heated to a controlled temperature, lying between 80 and 95 °C,
while purging with nitrogen to displace dissolved oxygen. Two solutions,
Solution (I) and Solution (II) are added, in the ratio 3-4:1-0, to the flask
charge, with continuous stirring, over a period of two hours. These
solutions are:

Solution (I) 100 g redistilled inhibitor-free alkenoic acid, 20 g propan-


2-ol in 100 cm3 water
Solution (II) 0-5-2-5 g ammonium persulphate in 60 cm3 water.

After the addition is completed the contents of the flask are heated for
a further two hours. The reaction mixture is then concentrated by vacuum
distillation at 40-45 °C until the desired concentration is attained.

97
Polyalkenoate cements
The molecular mass of the poly acid obtained lies between 10000 and
55000. Increasing the temperature of polymerization and the concen-
tration of ammonium persulphate serves to decrease the molecular mass of
the poly(alkenoic acid).
Poly(acrylic acid) is very soluble in water as are its copolymers with
maleic and itaconic acids. Solutions of 50% by mass are easily obtained.
The 'isomer' of PAA, poly(ethylene maleic acid), is not so soluble.
However, solutions of PAA tend over a period of time to gel when their
concentration in water approaches 50 % by mass (Crisp, Lewis & Wilson,
1975); this is attributed to a slow increase in the number of intermolecular
hydrogen bonds. Copolymers of acrylic acid and itaconic acid are more
stable in solution and their use has been advocated by Crisp et al. (1975,
1980).

5.4 Setting reactions


The cement-forming reaction of the polyelectrolyte cements may be
considered to take place in a number of overlapping stages. These are the
attack by the acid on the oxide or glass, the migration of the liberated ions
from the oxide or glass into the aqueous phase, the ionization of the
polyacid with consequent unwinding of the polymer chain, the interaction
between the charged chains and oxide or glass cations leading to ion
binding and gelation, and lastly the hardening phase represented by the
continuation of ion binding.
Setting results from the gelation of the poly(alkenoic acid) by metal ions
liberated from the metal oxide or silicate by acid attack. The gelation of
polysalts, which has been discussed in Sections 2.2.3 and 4.3, occurs as the
pH of the cement increases. As pointed out in those chapters there are
several physicochemical processes that underlie these rheological changes.
Amongst these are conformational changes in the polymer chain, binding
of the cations to the polymer chains, and hydration changes.
As reaction proceeds, the polymer chain (which is in random coil form)
unwinds as the charge on it grows as a result of neutralization and
ionization. This contributes to thickening of the cement paste. Cations
released become bound to the polymer chain. Countercations can either be
bound to a polyanionic chain by general electrostatic forces or be site-
bound at specific centres. More than one type of site binding is possible.
Complex formation and, if the ligand is bidentate, chelate formation
enhance the effect.

98
Molecular structures

The extent and rate of interaction between hydrated counterion and


polyanion depend on polymer structure, acid strength, conformation,
degree of dissociation, and distribution and density of ionic charge on the
polymer chain. This interaction between the cations - the counterions -
and the polyanion chain disrupts the hydration regions surrounding both.
Desolvation of the ion-pair, which depends on the nature of the cation and
the degree of neutralization, results in gelation. Gelation itself occurs
suddenly when the critical condition for the formation of an infinite
random network is met, that is when there are more than two crosslinks per
polymer chain (Flory, 1953, 1974).

5.5 Molecular structures


The molecular structure of the polyelectrolyte cements has been examined
by a number of workers using infrared spectroscopy (Crisp et ah, 1914;
Crisp, Prosser & Wilson, 1976; Wilson, 1982; Nicholson et al., 1988a,b).
The asymmetrical COO" stretching modes in particular can be used to

0 0 \ ^

2 +
C«0Zn 0iC ZrC C, Zn

0 0
^—•-" P I ' c ———^
(a) IONIC \ \ / / ( c ) CHELATING BIDENTATE

(b) BRIDGING BIDENTATE

\
CH-

\^ CH, Zn

CHC ^
(d) ASYMMETRIC UNIDENTATE / \
0
(e) CHELATE BIDENTATE
8-membered ring

Figure 5.3 Metal polyacrylate molecular structures: (a) purely ionic, (b) bridging bidentate,
(c) chelating bidentate, (d) asymmetric unidentate, (e) chelate bidentate.

99
Polyalkenoate cements
obtain structural information; if the metal-carboxylate bond is not purely
ionic and coordination complexes are formed then there are frequency
shifts. The types of structure are given in Figure 5.3. These structures are
(a) purely ionic, (b) bridging bidentate, (c) chelating bidentate, (d)
asymmetric unidentate, (e) chelate bidentate (Nakamoto, 1963; Mehrotra,
Bohra & Gauer, 1978; Mehrotra & Bohra, 1983). Infrared spectroscopy is
not able to distinguish between all these structures but asymmetric
stretching bands can be used to distinguish between COOH (c. 1700 cm"1),
ionic COO" (c. 1540 cm"1) and certain coordination complexes. Infrared
spectroscopy shows complexes with asymmetric bands both above and
below the ionic COO" band.
The bonds between PAA and Na+, Mg2+, Ca2+ are purely ionic, but the
absorption bands of other cation-PAA interactions - Zn2+ (1540-
1560 cm"1), Cu2+ (1605 cm-1) and Al3+ (1600 cm" 1 )-show evidence of
complex formation. The strength and stability of cements parallel the

I
CH 2 H20 H20 CH2 CH 2 H20 H20

/ | 0 \ /
I
|
o \ / CH — C — 0 Ca A
CH—C — 0"—Ca2—"0—C CH

I
CH2 H20
/ \ H20
o CH2 CH2 H20 H20

CH2

CH—C H20
CH 2 A' H20 CH 2
0"| / H20
I 8 \ / I CH 2 Ca2
CH—C—0'—Al 3 — 0 — C — CH
I / \ 5 ICH CH — C H20

CH
I
2 H0
2 H20 2
I °
CH2

Figure 5.4 Hypothetical molecular structures in polyalkenoate cements, where A represents


OH" or F-.

100
Molecular structures
magnitude of the complexation constants of the cations and are in the
order
Al3+ > Cu2+ > Zn2+ > Ca2+ > Mg2+.
The detailed molecular structure of the polyelectrolyte cements remains
a subject for conjecture. The structure is determined basically by the charge
and coordination number of the cation. Firstly, we must consider the
question of coordination and examine it in respect of the three most
important cations in these cements: Zn2+, Ca2+ and Al3+. Of the divalent
cations, Zn2+ can assume a coordination number of 4, 5 or 6 and Ca2+ of
6 or 8. If we assume a coordination number of 6, then an electrically neutral
coordination complex would have to contain two ligands with a single
negative charge and four neutral ligands. The coordination of Al3+ in
aqueous solution is 6 and for an electrically neutral complex this requires
that there should be three single-charged ligands and three neutral ligands.
In AB cements the ligands available are COO~, F", OH~ and H2 O; there
is a possibility of chelate formation and there are a number of possible
complexes. Chelate formation and bridging between chains cannot be
excluded.
Cations can be seen as acting as ionic crosslinks between polyanion
chains. Although this may appear a naive concept, crosslinking can be seen
as equivalent to attractions between polyions resulting from the fluctuation
of the counterion distribution (Section 4.2.13). Moreover, it relates to the
classical theory of gelation associated with Flory (1953). Divalent cations
(Zn2+ and Ca2+) have the potential to link two polyanion chains. Of course,
unlike covalent crosslinks, ionic links are easily broken and re-formed;
under stress there could therefore be chain slipping and this may explain
the plastic nature of zinc polycarboxylate cement.
A trivalent cation, for example Al3+, has the potential to link three
chains. Sterically, this is improbable; Mehrotra & Bohra (1983) assert that
simple aluminium tricarboxylates are not known in solution. Nevertheless
we consider it probable that a small proportion of Al3+ ions link three
chains, in which all three charged ligands are COO". More probable
molecular structures would contain one or two F~ ions; with a single F",
an [A1F(H2O)3]2+ unit could bridge two polyanion chains, while an
[A1F2(H2O)2]+ would have no crosslinking ability.
Some possible molecular structures are depicted in Figure 5.4.

101
Polyalkenoate cements
Table 5.1. Compressive strength of metal oxide-poly{acrylic acid) cements
{Elliott, Holliday & Hornsby, 1975; Hornsby, 1977)

Wet compressive
Strain at Trup
liquid, Strength, Modulus, failure, porosity,
Oxide gem" 3 MPa GPa % %
ZnO 1-4 76 1-40 5-4 18
CuO 20 83 102 81 24
HgO 10 29 0-66 4.4 25
PbO 20 26 0-61 4-3 26
MgO 10 58 0-45 12-9 —
Bi2O3 2-0 32 0-97 3-3 12

5.6 Metal oxide poly electrolyte cements


Many divalent and trivalent oxides form cements with PAA (Crisp,
Prosser & Wilson, 1976; Hodd & Reader, 1976; Hornsby, 1977). Cement
formation was observed using infrared spectroscopy and physical and
chemical tests. Of these cements that of ZnO (Smith, 1968) was thefirstand
remains by far the most important; it is given detailed treatment in Section
5.7.
Certain oxides of divalent metals, those of ZnO, CuO, SnO, HgO, and
PbO, form cements that are hydrolytically stable; in addition MgO, CaO,
BaO and SrO form cements that are softened when exposed to water.
Compressive strengths of these materials range from 26 to 83 MPa, the
strongest being the copper(II) and zinc polyacrylate cements (Table 5.1).
Crisp, Prosser & Wilson (1976) found that for divalent oxides the rate of
reaction increased in the order
CuO < ZnO < CaO < MgO.
This is in descending order of the stability constants of the cations.
Trivalent oxides A12O3, La2O3, Bi2O3 and Y2O3 are also capable of
cement formation but the reaction is only partial (Hornsby, 1977).
Hornsby also made the interesting observation that B2O3 forms a cement
with poly(acrylic acid), but, since B2O3 is acidic, an acid-base reaction
does not take place. Although the cement is hydrolytically unstable it is of
theoretical interest; it is to be presumed that cement formation takes place
by the formation of hydrogen-bonded complexes rather than by salt
formation.

102
Zinc polycarboxylate cement

The nature of the poly(alkenoic acid) can affect the hydrolytical


stability of metal oxide cements (Hodd & Reader, 1976). For example the
B2O3-poly(ethylene maleic acid) cement, unlike its poly(acrylic acid)
counterpart, is not hydrolytically stable.

5.7 Zinc poly carboxy late cement


5.7.1 Historical
The zinc polycarboxylate cement was the first of a new generation of dental
cements. It is based on the gelation of concentrated solutions of a
poly(alkenoic acid) by zinc ions provided by a zinc oxide powder (Wilson,
1975a,b, 1978a). It was invented as a result of a search by Smith (1968,
1969) for a luting cement that would, unlike the traditional zinc phosphate
dental cement, adhere to tooth material. It was the first adhesive dental
cement discovered and represented a considerable advance in dental
cement technology. It combined the strength of the zinc phosphate cement
with the bland qualities of the zinc oxide eugenol cement.
It is used for luting, lining and as a periodontal pack. Indeed, it can be
used to replace the zinc phosphate dental cement in all applications with
the possible exception of post crowns (crowns which are placed on a metal
post placed in the tooth root) and cantilever bridges (Smith, 1982a).
There are a number of brands on the market and, as far as can be
ascertained, development of this cement has virtually ceased since the mid
1970s.

5.7.2 Composition
In their original form these cements came as a zinc oxide powder and a
concentrated solution of poly(acrylic acid) (Wilson, 1975b). Since then
they have been subject to a number of chemical modifications.
The liquid is usually a 30-43 % solution of a poly(alkenoic acid) which
is a homopolymer of acrylic acid or a copolymer with itaconic acid, maleic
acid, or 3-butene 1,2,3-tricarboxylic acid (Smith, 1969; Bertenshaw &
Combe, 1972a; Jurecic, 1973; ESPE, 1975; Wilson, 1975b; Suzaki, 1976;
Crisp, Lewis & Wilson, 1976a; Crisp & Wilson, 1974c, 1977; Crisp et aL,
1980). The method of preparation has already been given in Section 5.3,
and the structures of these alkenoic acid units are shown in Figure 5.1.
The molecular mass of these polyacids varies from 22000 to 49000

103
Polyalkenoate cements
Table 5.2. Composition of zinc polycarboxylate cements
(Bertenshaw & Combe, 1972a, b, 1976)

Zinc oxide powder: 85-2-96-8% ZnO; 4-73-10-06% MgO


Poly(acrylic acid) solution: 32-4-42-9 %
Molecular weight: 15000-50000

Copolymers of acrylic acid with maleic or itaconic acid are sometimes


substituted for poly (acrylic acid).

(Smith, 1969; Bertenshaw & Combe, 1976). There is an optimum molecular


mass. A high molecular mass gives high-strength cements but leads to
difficulties in manipulation of the cement paste.
Two methods are available for the preparation of the powder (Smith,
1969). In one, zinc oxide is ignited at 900 to 1000 °C for 12 to 24 hours until
activity is reduced to the desired level. This oxide powder is yellow,
presumably because zinc is in excess of that required for stoichiometry.
Alternatively, a blend of zinc oxide and magnesium oxide in the ratio of
9:1 is heated for 8 to 12 hours to form a sintered mass. This mass is ground
and reheated for another 8 to 12 hours. The powder is white. Altogether
the powder is similar to that used in zinc phosphate cements.
Commercial powders are composed chiefly of a deactivated zinc oxide
containing up to 10% magnesium oxide (Bertenshaw & Combe, 1972b;
Kohmura & Ida, 1979). In addition they may contain silica, alumina or
bismuth salts. The most important additive is stannous fluoride (4-5 %),
which strengthens the cement although it was originally added as a fluoride
release agent (Foster & Dovey, 1974).
In some brands the polyacid is in dry form and blended with the zinc
oxide powder (Baumann & Gerhard, 1970; Jurecic, 1973; Bertenshaw &
Combe, 1972a). The cement is formed by mixing this powder blend with
water. In early examples, sodium dihydrogen phosphate was added to
the liquid (Bertenshaw & Combe, 1972a); as a result the viscosity of the
cement paste was lowered and setting was retarded, possibly because of the
slow dissolution of the solid polyacid (Bertenshaw, Combe & Grant, 1979).
Typical compositions of zinc polycarboxylate cements are given in Table
5.2.

104
Zinc polycarboxylate cement

5.7.3 Setting and structure


The cement sets as the result of an acid-base reaction between a zinc oxide
dental powder and a poly(alkenoic acid). The pH increases and an
insoluble amorphous salt is formed which acts as the cement matrix. A
general account of the gelation processes is given in Section 5.4.
Wilson (1982) studied the setting reaction of a stoichiometric cement
and a cement containing 100% excess of zinc oxide, using infrared
spectroscopy and physical and chemical methods. The setting of the
cement, measured by an oscillating rheometer, was paralleled by the loss of
bands associated with COOH groups (1700 cm"1 asymmetric C-O stretch)
and the appearance and growth of carboxylate bands at 1540-1560 cm"1
(asymmetric C-O stretch). Using deuterated cements, two bands were
observed in the young cement paste (1550 and 1560 cm"1), but only one in
the set cement (1550 cm"1). Nicholson et al (1988a) clarified this picture
using Fourier transform infrared spectroscopy. An initial fast ionic
reaction, associated with a band at 1562 cm"1, was attributed to a purely
ionic structure (Figure 5.3a). Later, as the cement matured, bands at 1554
and 1548 cm"1 became predominant; these were tentatively assigned to
chelated structure (Figure 5.3c). Finally, when the cement had set there was
one band at 1537 cm"1 (Figure 5.3d). This was attributed to a change in
bond type during setting and hardening. Of course, assignment of bands to
bond type is rendered more difficult by hydration and dehydration
processes. Wilson, Paddon & Crisp (1979) and Wilson, Crisp & Paddon
(1981) noted that as the cement matured the proportion of bound water to
total water increased.
X-ray diffraction shows that both the cement matrix and the salt are
amorphous (Wilson, 1982; Smith, 1971; Steinke et al, 1988). On the basis
of chemical analysis, Wilson (1982) assigned the following empirical
formula to the zinc polyacrylate salt:

Zn0.98H0.004(CH2. CH. COO)2 ,{H,O\ 61

He compared the infrared spectra of cements with that of zinc polyacrylate


salt and found differences. Inspection of his data shows that, unlike the
cements, the salt was purely ionic, so that it seems here that cement
formation is associated with the formation of coordination complexes.
There are no ligand field stabilization effects with the Zn2+ ion because it
has a completed d shell (Cotton & Wilkinson, 1966). For this reason the

105
Polyalkenoate cements
stereochemistry of Zn2+ compounds is determined solely by size, elec-
trostatic forces and covalent bonding forces. Zinc can be four-,five-or six-
coordinate. Most commonly it is four-coordinate, although six-coordinate
compounds are known. Five-coordination is rare.
Scanning electron microscopy shows the cement to consist of zinc oxide
particles embedded in an amorphous matrix (Smith, 1982a). As with the
zinc phosphate cement, a separate globular water phase exists since the
cement becomes uniformly porous on dehydration. Porosity diminishes as
the water content is decreased. Wilson, Paddon & Crisp (1979) distinguish
between two types of water in dental cements: non-evaporable (tightly
bound) and evaporable (loosely bound). They found, in the example they
examined, that the ratio of tightly bound to loosely bound water was
0-22:1-0, the lowest for all dental cements. They considered that loosely
bound water acted as a plasticizer and weakened the cement.
Most practical cements contain Mg2+ which is less strongly bound to the
polyacrylate than Zn2+ (Gregor, Luttinger & Loebl, 1955a). Magnesium
oxide forms a paste with PAA which sets to a plastic mass; this is not
hydrolytically stable, for when placed in water it swells and softens
(Hornsby, 1977; Smith, 1982a). Moreover, if ZnO powder contains more
than 10% MgO, the resultant cement deteriorates under oral conditions.
Evidence for the firm binding of Zn2+ comes from studies using labelled
zinc polyacrylate containing 65Zn and 14C. Only small amounts of these
ions were lost to a saline solution over a three-month period, even in the
presence of calcium (Peters et al., 1972; Peters, Jackson & Smith, 1974).
There is some evidence, from leaching studies, that Zn2+ is more firmly
bound to a copolymer of acrylic and itaconic acids than to poly(acrylic
acid), and less firmly bound to a copolymer of maleic and acrylic acids.

5.7.4 Properties
Setting
The zinc polycarboxylate cement sets within a few minutes of mixing and
hardens rapidly. Strength is substantially developed within an hour.
However, even when fully hardened the cement exhibits marked plastic
behaviour. Its most important property is its ability to bond permanently
to untreated dentine and enamel.
The early zinc polycarboxylate cement did not possess the ease of mixing
characteristic of the zinc phosphate and zinc eugenolate cements. It
suffered because it was expected to mix exactly as a traditional zinc

106
Zinc polycarboxylate cement

Table 5.3. Properties of zinc polycarboxylate cements (Jendresen


& Trowbridge, 1972; Plant, Jones & Wilson, 1972; Paddon &
Wilson, 1976; Powers, Johnson & Craig, 1974; Powers, Farah &
Craig, 1976; Chamberlain & Powers, 1976; Levine, Beech &
Garton, 1977; 0ilo & Espevik, 1978; Bertenshaw, Combe &
Grant, 1979; Peddy, 1981; Hinoura, Moore & Phillips, 1986)

Working time, 23 °C 2-5 minutes


Setting time, 37 °C 3-12 minutes
Compressive strength (wet), 24 h 48-80 MPaa
Tensile strength (wet), 24 h 4-8-15-5 MPab
Compressive modulus (wet), 24 h 3-2-6-2 GPa
Adhesion to enamel, 24 h (tensile) 4-1-6-9 MPa
Adhesion to dentine, 24 h (tensile) 2-2-5-1 MPa

a
Omitting atypical high and low values of 10 and 100 MPa
6
Measured by the diametral compression method, omitting an
atypical low value of 1-5 MPa

phosphate cement and the viscous nature of the polyacid liquid could
deceive the operator as to the actual fluidity of the cement paste (McLean,
1972). Frequently, the cement was mixed too thinly in a misguided attempt
to make it appear as fluid as a zinc phosphate cement paste. This led to
poor properties. In fact, the fluidity of the cement is greater than the
apparent consistency of the cement paste would suggest, because it is
pseudo-plastic. Thus, it exhibits shear thinning when a restoration is seated
on it, and it flows as readily as a zinc phosphate cement (Mortimer &
Tranter, 1969; McLean, 1972).
An unfortunate characteristic of early zinc polycarboxylate cements was
the early development of elastomeric characteristics - 'cobwebbing' - in
the cement pastes as they aged, thus shortening working time (McLean,
1972). Improvements in cement formulation, the addition of stannous
fluoride to the oxide powder (Foster & Dovey, 1974, 1976) and modi-
fications in the polyacid have eliminated this defect. However, the cements
have to be mixed at quite a low powder/liquid ratio, 1*5:1-0 by mass, when
used for luting.
The properties of these cements - the fluidity of the mix, the working
and setting times of the cement paste, and the strength of the set cement
- are affected by a number of factors. These include the composition of the
powder, the concentration, molecular mass and type of the polyacid, the

107
Polyalkenoate cements
powder/liquid ratio and the presence or absence of metalfluorides(Smith,
1971; Foster & Dovey, 1974, 1976).
Working time varies from 2 to 5 minutes (at 23 °C) and setting time from
3 to 12 minutes (at 37 °C) (Plant, Jones & Wilson, 1972; Jendresen &
Trowbridge, 1972; Chamberlain & Powers, 1976; Powers, Johnson &
Craig, 1974) (Table 5.3). These ranges are suitable, at the lower end, for the
cementation of single crowns and, at the upper end, for bridges. As with
other cements, working time can be prolonged by refrigerating the mixing
slab (McLean, 1972; Chamberlain & Powers, 1976).
Bertenshaw, Combe & Grant (1979) found thefilmthickness of cements
to vary widely from 20 |am to 110 |im, but this property depends on the
plasticity of the paste which changes rapidly with time. Thus, 0ilo & Eyje
(1986) found for one cement that film thickness increased from 15 \mi at
1 min, to 25 jim at 3*5 min, and to 60 \xm at 5 min. In practice,filmthick-
nesses lower than 25 jim, the specification upper limit for luting agents,
can be obtained (Jendresen & Trowbridge, 1972; 0ilo & Evje, 1986).
There is a hardening stage after set when the cement rapidly becomes
stronger and less plastic (Plant & Wilson, 1970; Bertenshaw, Combe &
Grant, 1979; Paddon & Wilson, 1976; Wilson, Paddon & Crisp, 1979).

Mechanical properties
All properties are time-dependent. Smith (1982a) reported one example
that developed 80% of its ultimate tensile strength in one hour and
maximum strength in 24 hours. Watts, Combe & Greener (1979) noted
little change in strength in seven days while Smith (1971) reported a slight
decline which he attributed to water sorption. Small increases in strength
have been recorded after 30 days (Osborne et al.9 1978; Smith, 1971) and
228 days (Smith, 1977). Paddon & Wilson (1976) found little increase in
either strength or modulus after 24 hours.
When cements are mixed to a luting consistency, compressive strength
varies, typically, from 48 to 80 MPa, compressive modulus from 3-2 to
6-2 GPa and tensile strength from 4-8 to 15-5 MPa (Table 5.3), all
measurements being made on 24-hour-old cements. Thesefiguresare not
absolute as they depend on test conditions. Thus, 0ilo & Espevik (1978)
found a temperature dependency; an increase in temperature from the
usual 23 °C to 37 °C reducing the compressive strength of a cement from
48 MPa to 36 MPa and modulus from 3-2 GPa to 1-9 GPa.
The cement shows marked viscoelastic properties. Thus, measured
strength is affected by the crosshead speed of the testing machine and this

108
Zinc polycarboxylate cement

effect is particularly noticeable if the crosshead speed is less than


O^mmmnT 1 (0ilo & Espevik, 1978; Wilson & Lewis, 1980). Wilson &
Lewis (1980) recorded a compressive strength of 65 MPa with crosshead
speed of 0-05 mm min"1 which increased to 100 MPa when the crosshead
speed was increased to 20 mm min"1.
Unlike other aqueous dental cements, the zinc polycarboxylate retains
plastic characteristics even when aged and shows significant stress
relaxation after four weeks (Paddon & Wilson, 1976). It creeps under static
load. Wilson & Lewis (1980) found that the 24-hour creep value for one
cement, under a load of 4-6 MPa, was 0-7 % in 24 hours, which was more
than that of a zinc phosphate cement (0-13 %) and a glass-ionomer cement
(0-32%), but far less than that of the zinc oxide eugenol cement (2-2%).
Plastic deformation is observed when the freshly set cement is subjected
to a slowly increasing load at 37 °C (Plant & Wilson, 1970; Hertet et al.,
1975; Paddon & Wilson, 1976; 0ilo & Espevik, 1978; Wilson, Paddon &
Crisp, 1979). 0ilo & Espevik (1978) recorded strain at failure of 1-7%, at
23 °C, and 4-3%, at 37 °C, values which are greater than that of a zinc
phosphate cement and far less than that of ZOE and EBA cements
(Chapter 9).
The plastic strain at fracture decreases markedly with time as the cement
ages; also the elastic modulus increases (Wilson, Paddon & Crisp, 1979;
Barton et al., 1975). There is an increase in dynamic modulus with time
(Barton et al., 1975).
Properties are affected by temperature. Compressive strength is reduced
from 48 MPa at 23 °C to 36 MPa at 37 °C. Strain at failure increases from
1-7% at 23 °C to 4-3% at 37 °C. But these are nothing like the massive
changes encountered with the ZOE and EBA cements. Although these
appear to be about as strong as the zinc polycarboxylate cement when
measurements are made at 23 °C, they are far weaker when tested at 37 °C,
the temperature of the mouth.

Erosion
Erosion depends on the solubility of the powder (the filler) and the matrix
in the aqueous medium. Here, acidity and complexing power of the
solution for metal ions compared with the stability of the metal PAA
complexes are important.
The solubility of these cements in water (when aged from one to 24
hours) is small and ranges from 0-1 to 0-6% (Gourley & Rose, 1972;
Bertenshaw, Combe & Grant, 1979; Crisp, Lewis & Wilson, 1976a; Smith,

109
Polyalkenoate cements
1971; Chamberlain & Powers, 1976; Jendresen & Trowbridge, 1972). The
addition of stannous fluoride to the cement increases dissolution, but this
is an advantage rather than a disadvantage, for the fluoride released is
taken up by neighbouring enamel (Bitner & Weir, 1973).
Once the cement has aged, dissolution occurs mainly at the site of the
oxide particles rather than at the matrix (Crisp, Lewis & Wilson, 1976a;
Anzai et al., 1977). The addition of high levels of magnesium oxide to zinc
oxide (for the purpose of densification) is undesirable. Early commercial
examples, containing 10% or more MgO added to the ZnO powder,
absorbed water, swelled and showed high dissolution. Crisp, Lewis &
Wilson (1976a) found that both zinc and magnesium are steadily eluted
from these cements with magnesium predominating. This observation led
them to recommend the omission of magnesium oxide from cement
formulations. The nature of the poly(alkenoic acid) was found to affect the
rate of elution of zinc. This elution was highest for cements based on a
copolymer of maleic and acrylic acids and lowest for those based on a
copolymer of acrylic and itaconic acids. Values for cements based on
poly(acrylic acid) lay between these two extremes. Water absorption
varied, according to brand, from 1-2 to 3-4 % and appeared to increase with
the ratio of COO to total C in the poly(alkenoic acid). It was also affected
by powder/liquid ratio.
More important is the behaviour of these cements in solutions
approximating to conditions in the mouth. Calcium does not affect the
stability, but phosphate, also a constituent of saliva, increases dissolution
(Peters et al, 1972; Peters, Jackson & Smith, 1974).
Acidic conditions greatly increase the erosion of the cement, to an extent
depending on the nature of the acid. Using the impinging jet method with
lactic acid/lactate solutions, Wilson et al. (1986b) found, for one cement,
erosion rates of 0-4% per hour at pH = 5-0, 5-3 % at pH = 4-0 and 16-2%
at pH = 2-7. For a group of different cements, Wilson et al. (1986a) found
erosion rates in solutions of pH = 2-7 in the range 8-5 to 19-8 %, and in one
exceptional case 0-1 %. These workers found that the zinc polycarboxylate
cement was markedly less resistant to acid erosion than the aluminosilicate
glass cements, the glass-ionomer and dental silicate cements. They also
found that, with one exception, zinc polycarboxylate cements were
somewhat less resistant to acid erosion than the zinc phosphate cement.
These results have been confirmed by Smink & Arends (1980) and Beech &
Bandyopadhyay (1983) using a similar method.
In vivo studies show that zinc polycarboxylate cements are much less

110
Zinc polycarboxylate cement
resistant to erosion than aluminosilicate cements, but there is no consensus
on their durability vis-a-vis the zinc phosphate cement (Ritcher & Ueno,
1975; Mitchem & Gronas, 1978, 1981; Osborne et al9 1978; Pluim &
Arends, 1981, 1987; Mesu & Reedijk, 1983; Theuniers, 1984; Pluim et al.,
1984; Pluim, Arends & Havinga, 1985).

Adhesion
An important property of the zinc polycarboxylate cement is its ability to
bond to untreated dentine and enamel. It also adheres to bone. This
adhesion can be observed visually (Mizrahi & Smith, 1969b; Smith, 1975;
Abramovich et al., 1977) and there is, of course, mechanical and chemical
evidence as well. After fracture, areas of cement remained attached to the
substrate (Smith, 1975; Eick et al., 1972). The principles and mechanism of
adhesion have already been discussed in Section 5-2. The bond strength to
tooth material develops rapidly in a matter of hours (Mizrahi & Smith,
1969a) and is maintained over many months (Mizrahi & Smith, 1969b,
1971). The tensile bond strength of the zinc polycarboxylate cement to
untreated dentine is 2-2 to 5-1 MPa (Bertenshaw, Combe & Grant, 1979;
Peddy, 1981; Levine, Beech & Garton, 1977; 0ilo, 1981; Hinoura, Moore
& Phillips, 1986). The bond strength to enamel is somewhat higher at 4-1
to 6-4 MPa (Peddy, 1981; Levine, Beech & Garton, 1977). Jemt, Stalblad
& 0ilo (1986)findthat in vivo bond strengths are much lower and reported
tensile values as low as 1-7 and 2-8 MPa. The lower bond strength to
dentine is significant and emphasizes that the bonding is to apatite. Thus,
demineralization of tooth material by acids reduces bond strength (Smith,
1975).
Both fluoride and calcium ions lead to an increase in bond strength.
Fluoride-containing cements bond more strongly to dentine, with a shear
bond strength of 7-0 MPa against 5-2 MPa for others (Causton, 1982);
bond strength can be increased by increasing the calcium concentrations at
the dentine surface (Beech, 1973). Such observations led to successful
attempts to improve bond strength by pre-treating dentine with specially
formulated solutions. Calcifying solutions were developed to pre-treat the
tooth surface and so improve bond strength. Levine, Beech & Garton
(1977) used a solution containing calcium hydrogen phosphate, sodium
fluoride and disodium hydrogen phosphate to pre-treat dentine and so
raised tensile bond strength from 2-4 to 5-5 MPa. Similarly, Causton &
Johnson (1982) used their so-called IT-S solution, a calcifying isotonic
solution buffered with carbonate and phosphates to pH of 7-4, and

111
Polyalkenoate cements
improved the shear bond strength of two examples of cement from 5-2 and
7-0 MPa to 7-2 and 12-9 MPa respectively.
The cement also bonds to metals. Saito et al. (1976) found that the bond
strength of zinc polycarboxylate cement to alloys decreased in the following
order of substrate: copper alloy, nickel-chromium alloy, silver-palladium
alloy, type III gold alloy. The bond strength to stainless steel has been
reported as varying from 6 to 9 MPa (Moser, Brown & Greener, 1974;
Jendresen & Trowbridge, 1972; Mizrahi & Smith, 1969a,b) with similar
values for the cobalt-chromium alloy (Moser, Brown & Greener, 1974).
The cement does not bond to the inert surfaces of porcelains.

Biological properties
The biocompatibility of these materials is in general excellent (Beagrie,
Main & Smith, 1972; Peters et al, 1972; Peters, Jackson & Smith, 1974;
Beagrie et al, 1974; Lawrence, Beagrie & Smith, 1975; Eames, Hendrix &
Mohler, 1979; Main et al., 1975). Results from in vitro experiments are
conflicting, but it would appear that cytotoxic effects (inhibition of cell
growth or cell death) are low unless test conditions permit the release of
toxic concentrations of zinc ions, fluoride ions and free poly(acrylic acid)
(Peters et al., 1972; Leirskar & Helgeland, 1977; Spangberg, Rodrigues &
Langeland, 1974; Welker & Neupert, 1974). It is probable that these
cements do not cause cytotoxic effects in use. However, the possible release
of zinc rules them out as bone cements.
The effects of zinc polycarboxylate cements on calcified tissue appear
minimal and they are as bland as the zinc oxide eugenol cement is towards
dental pulp (Smith, 1969; Plant, 1970; von Klotzer et al., 1970; Barnes &
Turner, 1971; Beagrie, Main & Smith, 1972; Jendresen & Trowbridge,
1972); the traditional zinc oxide eugenol is recognized as exceptional in this
respect. Reactions were mild in the teeth of monkeys even in deep cavities
(El-Kafrawy et al., 1974). In clinical practice these cements give little pain
and there are no long-term adverse effects (McLean, 1972). They are less
irritating than the zinc oxide eugenol cement when used in implants in soft
tissue and bone (Beagrie, Main & Smith, 1972; Lawrence, Beagrie &
Smith, 1975).
There are several reasons why these cements are bland. Acid irritation is
probably minimal. Poly(acrylic acid) is a weak acid and, in addition,
because of its high molecular weight will not readily diffuse along dentinal
tubules and is also immobilized by phosphatic material in these tubules.
Moreover, once set these cements rapidly become neutral.

112
Zinc polycarboxylate cement
Another cause of inflammation is leakage of bacteria from the mouth at
the interface between the cement and tooth material. Adhesion at the
interface reduces this effect.
Full accounts of the biological responses of these cements are to be
found in reviews by Smith (1982b), Helgeland (1982) and Granath (1982).

5.7.5 Modified materials


The most important modification of these materials was the discovery of
the effect of adding stannous fluoride (Foster & Dovey, 1974, 1976).
Originally added to provide fluoride release, it was found to improve the
mixing qualities of the cement and to increase strength by about 50 %. This
is reflected also in improved adhesion to enamel and dentine (Section
5.7.4).
Attempts have been made to improve the mechanical properties of these
cements by adding reinforcing fillers (Lawrence & Smith, 1973; Brown &
Combe, 1973; Barton et al, 1975). Lawrence & Smith (1973) examined
alumina, stainless steel fibre, zinc silicate and zinc phosphate. The most
effective filler was found to be alumina powder. When added to zinc oxide
powder in a 3:2 ratio, compressive strength was increased by 80% and
tensile strength by 100 % (cements were mixed at a powder/liquid ratio of
2:1). Because of the dilution of the zinc oxide, setting time (at 37 °C) was
increased by about 100 %. As far as is known, this invention has not been
exploited commercially.

5.7.6 Conclusions
The cement is to be regarded as a replacement for the zinc phosphate
cement. Its advantages over that traditional cement are its adhesion to
tooth substance and a more bland reaction towards living tissues. Its
disadvantages are that it is less user-friendly than the zinc phosphate
cement and requires greater care in preparation. Unlike other aqueous-
based dental cements, it retains distinct plastoelastic properties even after
ageing for months. Thus, it is less rigid and more liable to creep. Whether
this constitutes an advantage or disadvantage is a matter of opinion. The
current view is that cement properties should match those of dentine, in
which case it is too plastic. On the other hand, the zinc phosphate cement
is too rigid.

113
Polyalkenoate cements

5.8 Mineral ionomer cements


Cements can be formed by reacting silicate minerals with solutions of a
poly(alkenoic acid) but they are weaker than zinc polycarboxylate and
glass polyalkenoate cements and so far have found little practical use. They
are, however, of theoretical interest for they give some insight into the
silicate structures required for cement formation. Broadly, silicates that are
capable of cement formation fall into the class of naturally occurring
silicates known as gelatinizing minerals. These are minerals decomposed
by acids with the formation of silica gel; this behaviour is related to silicate
structure. Larsen & Berman (1934) recorded the effect of acids on a large
number of minerals and their work has been used by others to classify
gelatinizing minerals (Murata, 1943; Mase, 1961; Petzold, 1966).
It is helpful in the discussion to describe silicate structures using the Qn
nomenclature, where Q represents [SiOJ tetrahedra and the superscript n
the number of Q units in the second coordination sphere. Thus, isolated
[SiO4]4~ are represented as Q° and those fully connected to other Q units as
Q4. In general, minerals based on Q°, Q1, and Q2 units are decomposed by
acids. Such minerals are: those containing isolated silicate ions, the
orthosilicates, SiO4^ (Q°); the pyrosilicates, Si2C>7~ (Q1); ring and chain
silicates, (SiO3)2w~ (Q2). Certain sheet and three-dimensional silicates can
also yield gels with acids if they contain sites vulnerable to acid attack. This
occurs with aluminosilicates provided the Al/Si ratio is at least 2:3 when
attack occurs at Al sites, with scission of the network (Murata, 1943).
Mase (1961) explains the reactivity of silicate minerals towards acids in
terms of the polarizing ability of the cation and the bond energy of the
mineral. Clearly, too, the reactivity of minerals towards acids is connected
with their basicity, and one notes that the orthosilicates are basic minerals.
According to the ideas of Flood & Forland (1947a,b) (Section 2.3.3)
basicity is related to the residual polarizability of the oxygen atom. If this
is large, as will be the case if the associated cation has little polarizing
power, then the oxygen atom is basic. Thus, sodium silicates are the most
basic of the silicates, and silica itself is acidic and so resistant to acid attack.
Basicity is also related to the silicate structure: orthosilicates are more
basic than pyrosilicates which, in turn, are more basic than ring and chain
silicates.
Using this information, Crisp et al. (1977, 1979) and Hornsby et al.
(1982) selected candidate minerals for cement formation with poly (acrylic
acid) and found a number of minerals that formed cements (Table 5.4).

114
Table 5.4. Properties of mineral ionomer cements {Crisp et al, 1979; Hornsby et aL, 1982)

Compressive
strength
Powder: Setting (24 h), MPa
liquid, time, Sohibil
gem" 3 min Humid Water

Orthosilicates
Willemite Zn 2 [SiOJ 2-0 26 19 21 0-35
Gadolinite Be 3 Fe(YO) 2 [SiOJ 2 2-1 140 40 40 002
Pyrosilicates
Gehlenite Ca2Al[AlSiO7] 1-0 48 1 6 2-4
Hardystonite Ca 2Zn[Si 2O7] 10 6 9 12 0-4
Chain silicate
Wollastonite Ca3[(SiO3)3] 1-0 14 18 3 3-4
Sheet silicate
Thuringite (Fe(II), Fe(III), Mg, Al)12[(Si, Al)8O20](O, OH, F ) r 1-5 29 28 35 10
Zeolite
Scolecite Ca[Al2Si3O10]3H2O 0-5 20 160 30 1-35
Ultramarine
Hackmanite Na 8 [Al 6 Si 6 O 2 J(Cl 2 ,S) 10 13 89 37 2-4
Feldspar
Labradorite (Ca,Na)[Al1_2Si2_3O8] 1-0 59 134 2 3-6
Polyalkenoate cements
While some formed hard, rigid cements that were stable in water, others
yielded rubbery or plastic masses that were hydrolytically unstable.
Minerals with cement-forming capability were found in the following
classes:

(1) Island silicates containing discrete ions. Orthosilicates,


SiO4~ (Q°); pyrosilicates, Si2O*~ (Q 1 ); and ring silicates, Si3OJ-,
S&.O"- (Q2).
Most orthosilicates reacted completely with poly(acrylic acid)
solution; an exception was andradite, Ca 3 Fe 2 [SiO4]3. Even so, the
cements of gehlenite and hardystonite were very weak and affected
by water. Only gadolinite and willemite formed cements of some
strength which were unaffected by water, probably because one
contained beryllium and iron and the other zinc.
(2) Chain silicates, consisting of connected metasilicate units,
(SiO3)2w~ (Q2), and of an open structure. Wollastonite Ca(SiO3)
reacted completely with poly(acrylic acid), but the cement was
much affected by water.
(3) Sheet silicates (Q3) with significant isomorphic replacement of Si4+
by Al3+ or Fe 3+ . These were decomposed by poly (aery lie acid) to
silica gel. The chlorite, thuringite, formed a strong cement but was
much affected by water.
(4) Aluminosilicates with a three-dimensional network (Q 3 and Q4)
where the Al/Si ratio was 2:3. These reacted with poly(acrylic
acid), but none reacted completely.
The zeolite, scolecite, the feldspar, labradorite, and the ultra-
marine, hackmanite, gave high-strength cements but all were
much affected by water - the strength of the labradorite cement
disappeared almost entirely - possibly because of the presence of
free acid.

5.9 Glass polyalkenoate (glass-ionomer) cement


5.9.1 Introduction
The glass polyalkenoate cement, formerly known as the glass-ionomer
cement, was invented by Wilson and Kent in 1969 (Wilson & Kent, 1973)
and is now well established as a material that has an important role in
clinical dentistry. It has proved to have considerable development potential
and has been subjected to continuous development, improvement and

116
Glass polyalkenoate (glass-ionomer) cement
diversification. It is the most versatile of all dental cements and currently
accounts for most of the research and development on them. There are
other applications of the cement as a biomaterial. It is used as a splint
bandage material and as a bone cement.
Glass polyalkenoate cement has a unique combination of properties. It
adheres to tooth material and base metals. It releases fluoride over a long
period and is a cariostat. In addition it is translucent and so can be colour-
matched to enamel. New clinical techniques have been devised to exploit
the unique characteristics of the material.
The material originated from the general dissatisfaction with the clinical
performance of the dental silicate cement. Wilson and his coworkers made
extensive studies on the dental silicate cement (Section 6.5) and drew the
conclusion that this cement could not be further improved. Wilson (1968)
examined several alternatives to orthophosphoric acid, including organic
chelating agents, as a liquid cement-former, but none of these were
successful. Finally, after considerable research, the glass polyalkenoate
cement was developed (Wilson & Kent, 1971, 1972, 1973). The cement is
formed by mixing an ion-leachable glass powder with an aqueous solution
of a poly(alkenoic acid). The glass is generally a fluoride-containing
calcium aluminosilicate but calcium may be replaced by strontium or
lanthanum.
The cement was originally known as ASPA, an acronym of Alumino-
silicate Polyacrylic Acid. The term ASPA is now applied to materials
developed by the Laboratory of the Government Chemist in the UK, and
was once also the brand name of an early commercial material. For many
years it was known as the glass-ionomer cement - indeed, that is still the
term in common use-but the International Standards Organization
officially adopted the name glass polyalkenoate cement. The term
glass-ionomer cement is now used as a generic term to cover these cements
and the new glass polyphosphonate cements invented by Ellis and Wilson
in 1987-9 (Ellis & Wilson, 1990).

5.9.2 Glasses
General
The powders used in glass polyalkenoate cement formulations are prepared
from glasses and not opaque sintered masses. In this they resemble the
traditional dental silicate cement from which they are descended. The glass
plays several roles in the chemistry and physics of the glass polyalkenoate

117
Polyalkenoate cements
cement. It acts as a source of ions for the cement-forming reaction, controls
the setting rate and strength of the cement and imparts the property,
unusual in a cement, of translucency.
Chemically these are special aluminosilicate glasses. Until quite recently,
all were calcium aluminosilicates, but now calcium is sometimes wholly or
partly replaced by strontium and lanthanum. Most glasses also contain
fluorides which, besides lowering the temperature of glass fusion, play a
role in cement formation and affect cement properties. Provided the Al/Si
ratio is high enough, these glasses are decomposed by acids to release
cement-forming ions (Wilson & Kent, 1973, 1974; Crisp & Wilson,
1978a,b, 1979; Kent, Lewis & Wilson, 1979; Wilson et al., 1980; Hill &
Wilson, 1988a). They are similar to the glasses used for dental silicate
cements, although the Al/Si ratio is higher.

Types of glass
There are a great number of potential glasses and some can be extremely
complex. All contain silica and alumina and an alkaline earth or rare earth
oxide or fluoride. The two essential glass types are SiO2-Al2O3-CaO and
SiO2-Al2O3-CaF2, from which all others are derived.
Oxide glasses have been reported by Crisp & Wilson (1978a,b, 1979),
Wilson et al. (1980), and Hill & Wilson (1988a). The fusion mixtures
contain silica, alumina and calcium carbonate to which sodium carbonate
or calcium orthophosphate may be added. They may be represented thus,
with fusion temperature given in parentheses:
SiO2-Al2O3-CaO (1350-1550 °C)
SiO2-Al2O3-CaO-P2O5 (1370-1450 °C)
SiO2-Al2O3-CaO-Na2O (1200-1350 °C)
Influorideglasses, calciumfluorideis an essential constituent, but generally
cryolite, Na3AlF6, is also added as a flux to lower the temperature of
fusion. Aluminium orthophosphate is also generally added to the fusion
mixture for various reasons. Of course, the various elements may be added
in different ways. Thus, calcium orthophosphate, aluminium fluoride and
sodium carbonate are often used in the preparation of fluoride glasses.
Apart from lowering the temperature of glass fusion, fluoride improves
the handling qualities of the cement paste, increases cement strength and
translucency, and has a therapeutic quality when used as a dental filling
material. In fluoride glasses the ratio of alumina to silica controls the
setting time of the cement; fluoride tends to slow setting while aluminium

118
Glass polyalkenoate (glass-ionomer) cement
orthophosphate improves the mixing of the paste. Sodium in the glass
improves the translucency of the cement but can affect its hydrolytic
stability. In addition, glasses have been reported where calcium is replaced
by strontium or lanthanum (Akahane, Tosaki & Hirota, 1988) which
impart radio-opacity to the cement.
Fluoride glasses are difficult to classify because the various constituents
can be added to the fusion mixture in several ways. However, glasses of the
Laboratory of the Government Chemist (Wilson & Kent, 1973; Kent,
Lewis & Wilson, 1979; Wilson et ai, 1980; Hill & Wilson, 1988a), which
form the basis of many commercial cements, can be represented as
SiO2-Al2O3-CaF2 (1150-1350 °C)
SiO2-Al2O3-CaO-CaF2 (1320-1450 °C)
SiO2-Al2O3-CaF2-AlPO4 (1150-1300 °C)
SiO2-Al2O3-CaF2-AlPO4-Na3AlF6-AlF3 (1100-1300 °C)
where again the temperature of fusion is given in parentheses.
After fusion the molten glass is shock-cooled by pouring it onto a metal
plate and then into water. The glass fragments are then finely ground to
pass either a 45-|im sieve for afillingmaterial, or a 15-jim sieve for a fine-
grained luting agent. The glass powders may be annealed after preparation
by heating at 400 to 600 °C; in general, the effect is to slow down the setting
reaction. Sometimes the powder is acid-washed to improve the mixing
qualities of the cement.

Structure of aluminosilicate glasses


The formation of a cement is dependent on the ability of the glass to release
cations to acid solutions. It is not sufficient for network-modifying cations
to be exchanged for hydrogen ions as this would restrict the attack to the
glass surface only. It is required that the glass structure itself be completely
decomposed if all the glass ions are to be available for release.
Aluminosilicate glasses have this property. To discuss why this is so it is
useful to have an appropriate conceptual framework, and one can be
developed from the Random Network model of Zachariasen (1932).
Zachariasen (1932) conceived of a glass structure as a random assembly
of oxygen polyhedra, these polyhedra consisting of a central glass-forming
cation surrounded by a small number of oxygen atoms, e.g. [SiOJ
tetrahedra. These polyhedra were considered to be linked at corners only,
via 2-coordinate oxygen atoms. This concept amounts to regarding a glass
as a type of highly crosslinked polymer based on -Si-O-Si- linkages. This

119
Polyalkenoate cements
idea, although much criticized (Rawson, 1967), has proved to be a fruitful
one. Zachariasen (1932) added another criterion, namely that the random
network was three-dimensional, and, therefore, in modern terminology,
composed only of Q4 and Q3 units. Hagg (1935) considered that this
requirement was not always necessary and a glass might contain large
irregular anionic groups. The work of Trap & Stevals (1959) supported this
view, for they prepared so-called invert glasses containing only Q2 and Q1
units, that is glasses with no crosslinking. In these glasses at least half the
oxygen atoms are non-bridging -O" groups, so the -Si-O-Si- chains are
anionic and are held together by network-modifying cations (these do not
form part of the glass structure). Today, following Ray (1975, 1983) we
would call these ionic polymers.
We are now in a position to discuss requirements for ionomer glasses
further. Consider the case of the simple silica (SiO2) glass where we can
represent the network diagrammatically thus:

O O O O
I I I I
— Si—O—Si—O—Si—O—Si — O
I I I I
This infinite three-dimensional network is electrically neutral and im-
pervious to acid attack. If so-called network-modifying cations are
introduced then this network must acquire a negative charge leading to the
breaking of an Si-O-Si bridge to form non-bridging oxygens:
\ / Ca2+ \ „ /
— Si—O—Si— • — S i — Or Ca 2+ "O — S i —
/ \ / \
This is a type of ionic polymer where the negative charge on the network
is balanced by the positively charged network modifier. Statistically all
types of [SiOJ tetrahedra, Q1, Q2, Q3 and Q4, will be present in varying
proportions, depending on the ratio of bridging to non-bridging oxygens.
Aluminosilicates are more complex as aluminium can be either a
network modifier in sixfold coordination or a network former in fourfold
coordination. In the latter case, Al3+ is able to replace Si4+ in the glass
network because it has a similar ionic radius, but the network then acquires
a negative charge. If this charge becomes sufficiently high then the network
becomes susceptible to acid attack. Again this charge on the network has
to be balanced by positively charged network-modifying cations. Thus, we

120
Glass polyalkenoate (glass-ionomer) cement

can regard an aluminosilicate glass structure as consisting of linked [SiOJ


and [A1OJ~ tetrahedra. There are restrictions on the replacement of Si4+ by
Al 3+ . The Al/Si ratio apparently cannot exceed 1:1 (Lowenstein, 1954).
Nor can all the aluminium go into the network if there are insufficient
network-modifying cations to balance the network charge. Under such
conditions aluminium adopts a sixfold coordination.
We must note that recently Ellison & Warrens (1987), using 27 A1NMR
spectroscopy, have found evidence for the existence of aluminium in
pentacoordination in asymmetric or distorted sites using previously
established assignments (Kirkpatrick et al. 1986; Risbud et al., 1987;
Cruikshank et al., 1986).
A negatively charged network of non-bridging oxygens and aluminium
sites renders these glasses susceptible to acid attack. Overall the in-
troduction of network-modifying cations and aluminium ions increases the
polarizability of oxygen ions and, therefore, vulnerability to acid attack.
The mode of acid decomposition of an aluminosilicate glass is depicted in
Figure 5.5. It can be seen that attack by hydrogen ions involves exchange
of network-modifying cations (Ca2+, Na + ) and rupture of the alumino-
silicate network at aluminium sites to yield silicic acid and aluminium ions
(Wilson, 1978b; Prosser & Wilson, 1979). Glasses used in glass poly-
alkenoate cements have been observed to release cations, fluoride if present
and silicic acid (Crisp & Wilson, 1974a; Wasson & Nicholson, 1990,1991).
Similar observations have been made for the related dental silicate cement

polymerizes
silica gel
Figure 5.5 The mode of acid decomposition of an aluminosilicate glass.

121
Polyalkenoate cements

Table 5.5. Glass compositions and acid extracts {Crisp & Wilson,
1974a; Wasson & Nicholson, 1990)

G-200 G-338

Cement Cement
Mole ratio Glass extract Glass extract

Si:Al 0-98 0-67 0-63


Ca:Al 0-89 1-17 0-26 0-36
Na:Al 014 0-25 0-44 0-65
F:A1 2-46 0-97 1-67

Table 5.6. Composition of oxide glasses used in studies on polyalkenoate


cements, parts by mass {Wilson et al., 1980; Crisp, Merson & Wilson,
1980)

G-273 G-275 G-287 G-255 G-247

SiO2 120 240 180 120 160


A12O3 102 102 102 102 100
CaO 168 112 56 56 —
Ca 3 (PO 4 ) 2 — — — — 140
Appearance clear clear clear opaque opal
Crystallites — — — An —
Properties
Powder .liquid, g cm"3 20 30 30 30 —
Setting time (37 °C), min 2-75 40 8-25 40 5-25
Strength (24 h), MPa 95 35 29 56 72

An = Anorthite

(Wilson & Kent, 1970). The release of each glass species is roughly
governed by the amount contained in the glass and so varies with glass
type (Table 5.5).
The reactivity of a glass towards acids depends on its acid-base
properties and both the Bronsted-Lowry and Lewis theories have been
applied to oxide glasses (Volf, 1984). Basic components of a glass are the
metal oxides, and acidic ones are silicon, boron or phosphorus oxides. The
important factor is the state of the oxygen atoms. In purely oxide glasses
the basicity of a glass depends on the ability of the oxygen atoms to give up

122
Glass polyalkenoate (glass-ionomer) cement
electrons. This is greatest when the oxygen atoms are associated with
cations of low electrostatic field strength, for example Na+ and Ca2+, and
least when the cations have a high electrostatic field strength, for example
the highly charged, small Si4+ ion.
Lux (1939) introduced the symbol pO (note it is not an exponent like pH)
to quantify the acid-base balance in a glass, and various attempts have
been made to obtain values for this parameter. All are based on the
electronegativity of the cation or a related characteristic, such as
electrostatic field strength (Volf, 1984).

SiO2-Al2O3-CaO glasses
In these glasses (Table 5.6) the coordination state of aluminium depends on
its chemical environment and can only be entirely fourfold when the Ca/Al
SiO
* Non-setting
D Slow setting
A Moderate setting
o Fast setting
9 Fast setting
crystalline mass
1:3 • Ultra-fast setting

1:2

Al /Si mole ratio

C9S

C,S

CaO A1 2 O 3

Figure 5.6 Triangular composition diagram for SiO2-Al2 O3-CaO glasses, showing that
glasses with cement-forming ability fall within the gehlenite and anorthite composition
region, and that only glasses with less than 61 to 62 % by mass of silica have the potential to
form a cement (Hill & Wilson, 1988a).

123
Polyalkenoate cements

Table 5.7. Properties of cements formed from glasses


corresponding to the generic formula xSi02.Al203.Ca0

Cement properties

setting time, strength,


moles(x) mass % minutes MPa

10 21-9 3-5 104


20 35-9 2-25 74
40 52-8 40 35
60 62-7 non-setting zero

SiO
* Zero strength
D Unworkable
• Weak
o Low strength
• Moderate strength
1:3

1:2

Al /Si mole ratio

C2S

C,S

CaO A12O3

Figure 5.7 Triangular composition diagram for SiO2-Al2 O3-CaO glasses. Glasses in the
gehlenite region yield stronger cements (95 to 104 MPa in compression) than those in the
anorthite region (29 to 56 MPa) (Hill & Wilson, 1988a).

124
Glass polyalkenoate (glass-ionomer) cement

Table 5.8. Composition of fluoride glasses used in studies on polyalkenoate


cements, parts by mass (Kent, Lewis & Wilson, 1979; Wilson et al., 1980)

G-200 G-307 G-309 G-235 G-237 G-338


SiO2 175 133 67 175 175 175
A12O3 100 100 100 100 100 100
CaF 2 207 100 100 166 117 90
Na 3 AlF 6 30 — — — — 135
A1F3 32 — — — — 32
A1PO4 60 — — 60 60 170
SiO 2 :Al 2 O 3 bymass 1-75 1-33 0-67 1-75 1-75 1-75
Appearance opaque clear opaque opal opaque opal
Crystallites Fl Fl,Co Fl Co Ap
Properties
Powder: liquid, g cm"3 30 30 30 3-5 30 1-8
Setting time (37 °C), min 5-2 3-5 6-5 3-0 30 3-75
Strength (24 h), MPa 185 107 166 149 199 149

Fl = Fluorite, Co = Corundum, Ap = Apatite

ratio > 1:2 and Al/Si ratio < 1:1 (Isard, 1959; Lowenstein, 1954). Thus,
aluminium is in fourfold coordination in anorthite glass (Ca: Al = 1:2,
Al:Si = 1:1); the glass is composed of Q4 and Q 3 units, i.e. it is three-
dimensional. Gehlenite glass must contain some aluminium in sixfold
coordination (Ca: Al > 1:2, Al:Si > 1:1) and is composed of paired Q 1
units, i.e. [AlSiO7].
A study by Ellison & Warrens (1987) on two of these glasses, using 27A1
and 29Si NMR, produced results not too dissimilar from theoretical
predictions. In glass G-273, Ca3Al2Si2O9 (Table 5.6), aluminium was
found to be mainly in tetrahedral coordination with a minor amount
in octahedral coordination. Similar results were found for glass G-275,
Ca2Al2Si4O13 (Table 5.6), but, in addition, some aluminium was found
to be pentacoordinate. Possible structural units were considered to be
Q 3 (1A1) and Q 2 (0A1) with some Q 4 (3A1). The number of Al replacing
Si in the second coordination sphere is given in parentheses.
Glasses that have cement-forming ability fall within the gehlenite and
anorthite composition regions of this system, and only glasses with less
than 61 to 62 % by mass of silica have potential to form a cement (Figure
5.6). Cements are not formed if the Si/Al mole ratio exceeds 3:1. When the

125
Polyalkenoate cements
ratio is less than 2:1 fast-setting cements are obtained with setting time of
2 to 10 minutes. There is only a very small region for slow-setting cements
and as Table 5.7 shows there is a critical region between setting and non-
setting. Glasses in the gehlenite region yield stronger cements (95 to
104 MPa in compression) than those in the anorthite region (29 to 56 MPa)
(Figure 5.7).

SiO2-Al2O3-CaF2 glasses
These (Table 5.8) are the basic type from which most biomedical glass
polyalkenoate cements are derived. Although thefluoridecontent is high,
many of these shock-cooled glasses are clear. Clear glasses are confined to
a narrow central compositional range at the centre of the phase diagram
where the Al2O3/CaF2 ratio is around 1:1 by mass and the SiO2/Al2O3
ratio exceeds 1-33:1 by mass (Figure 5.8). Outside this regionfluorite,and
sometimes corundum, phase-separate. Even the clear glasses can be

SiO2
• Clear, non-setting
• Opal, non-setting
A Clear, slow setting, low strength
• Opal, slow setting, low strength
o Clear, fast setting, high strength
• Opal, fast setting, high strength
1:3

CaF 2 /Si0 2 ratio by mass ratio by mass

3:1

CaF2 AI2O3
ratio by mass
Figure 5.8 Clear glasses are confined to a narrow central compositional range at the centre
of the phase diagram where the Al 2 O 3 /CaF 2 ratio lies in the region 1:1 by mass and the
SiO 2 /Al 2 O 3 ratio exceeds 1-33:1 (Hill & Wilson, 1988a).

126
Glass polyalkenoate (glass-ionomer) cement

induced to phase-separate when heated to 450 °C, and this reduces their
reactivity.
The ability of a glass to form a cement is governed by the SiO 2 /Al 2 O 3
ratio which represents the acid-base balance in these glasses. If this ratio is
3-0 or more by mass then the glass will not form a cement. If it is below 2-0
then the cements formed are rapid-setting (2-5 to 5-0 minutes). Glasses in
a very narrow band around a ratio of about 2-0 are slower-setting (6-5 to
18 minutes). The critical ratio for non-setting lies somewhere between 2-0
and 3-0. The effect of SiO 2 /Al 2 O 3 ratio on setting time and compressive
strength is shown in Figure 5.9. Note that compressive strength increases
steadily as the SiO 2 /Al 2 O 3 ratio decreases. Setting time decreases as the
SiO 2 /Al 2 O 3 ratio decreases until a point is reached when phase separation

10.0i 200 - | 1 -

8.0 0.8
150

|6.0 0.6
o>
cb
< o
^3
if)
o
100
% ••5
essi

a
0
Z 4.0 Q." 0.4
I/) £
o
u

50

2.0 0.2

Opal glasses I Clear glasses

G-309 G-308 |G-307 G-379


I I J
0.5 1.0 1.5 2.0
SiO2 • AI2O3 by mass

Figure 5.9 The effect of SiO 2 /Al 2 O 3 mass ratio on setting time, compressive strength and
opacity (Hill & Wilson, 1988a).

127
Polyalkenoate cements
occurs in the glass. Phase separation has the effect of deactivating the glass.
In these glasses the main phase is depleted in calcium and fluoride, which
reduces its reactivity. Acid attack occurs selectively at the phase-separated
droplets which are rich in calcium and fluoride. This selective attack is
shown in Figure 5.10.
Phase-separated glasses produce stronger cements than clear glasses.
The strongest cements produced from a clear glass have a compressive
strength of 130 MPa and a flexural strength of 20 MPa, whereas phase-
separated glasses produce cements with compressive strength exceeding
200 MPa and flexural strengths exceeding 35 MPa. Note that fluoride
glasses produce much stronger cements than oxide glasses. The strongest
cement produced from an oxide glass has a compressive strength of only
104 MPa.
Ellison & Warrens (1987) have reported NMR results on an atypical
phase-separated glass of extreme composition G-309 (Table 5.8) finding

Figure 5.10 In these glasses, the main phase is depleted in calcium and fluoride, which reduces
its reactivity. Acid attack occurs selectively at the phase-separated droplets which are rich in
calcium and fluoride (Hill & Wilson, 1988a).

128
Glass polyalkenoate (glass-ionomer) cement

Table 5.9. Composition (%) of oxide/fluoride glasses used in studies on


polyalkenoate cements, parts by mass. Generic composition:
2SiO2.Al2O3. (2-x)CaO.xCaF2 (Kent, Lewis & Wilson, 1979;
Wood & Hill, 1991a)

G-241 G-278 G-276 G-279 G-280 G-281 G-282

0 0-20 0-33 0-5 10 1-5 1-8


SiO2 120 120 120 120 120 120 120
A12O3 102 102 102 102 102 102 102
CaO 112 101 93 84 56 28 11-2
CaF 2 0 15-6 26 39 78 117 140
Tg nd 745 nd 717 642 nd 636
Appearance below clear clear clear clear clear clear opaque

Crystallites below — — — — — — Fl

Properties
Powder: liquid, 2-5 2-5 20 20 2-0 20 2-5
gcnr3
Setting time (37 °C), 2-25 2-25 2-5 u/w u/w u/w 3-0
min
Strength (24 h), 74 125 120 u/w u/w u/w 165
MPa

Fl = Fluorite, nd = not determined


u/w = unworkable pastes that set during mixing

that the Al is mainly in sixfold coordination, not surprising in view of the


high alumina content. The structural units are mainly Q 4 (3A1).
The role of fluoride is a matter of speculation and debate (Kumar, Ward
& Williams, 1961). According to Weyl and Marboe (1962), in addition
to [SiOJ and [A1OJ, such glasses contain tetrahedra such as [SiO 3 F],
[A1O3F] etc. The replacement of O2~ by F~ reduces the screening of the
central cation and so strengthens the remaining cation-oxygen bonds, but
fluoride is non-bridging and so structure-breaking. This role of fluoride
may be represented, thus

Q4 Q3
Another view of the role of fluoride is that metal fluorides occupy holes in
the major glass network (Rabinovich, 1983). There is experimental

129
Polyalkenoate cements
evidence to support both views. However, these views are not necessarily
mutually exclusive.

SiO2-Al2O3-CaO-CaF2 glasses
These glasses (Table 5.9) with generic composition 2SiO2.Al2O3.
(2-x)CaO.xCaF 2 , first described by Wilson et al (1980) and recently
studied by Wood & Hill (1991a), are of theoretical interest. In this series,
starting with the oxide glass G-241, O atoms are progressively replaced by
F atoms (Table 5.9). Fluoride is often considered to be a structure-breaker
and this is reflected in the reduction of the glass transition temperature, T g,
as x increases (Wood & Hill, 1991a). Another indication of structure
breaking is shown by a decrease in setting time of cements as x increases
(Wilson et al, 1980). When x reaches 0-5 (G-279) the reaction between
glass powder and polyacid liquid becomes very vigorous and cement pastes
set during mixing. Glasses with x values of 1-0 and 1-5 behave similarly.
However, when x reaches 1-8 (G-282) phase-separation of fluorite (CaF2)
occurs, fluoride is removed from the main phase, and the reactivity of the
glass is reduced. The cement of G-282 has the longest setting time of the
series.
Wood & Hill (1991b) induced phase-separation in the clear glasses by
heating them at temperatures above their transition temperatures. They
found evidence for amorphous phase-separation (APS) prior to the
formation of crystallites. Below the first exotherm, APS appeared to take
place by spinodal decomposition so that the glass had an interconnected
structure (Cahn, 1961). At higher temperatures the microstructure con-
sisted of distinct droplets in a matrix phase.
When x was 1-0, fluorite and anorthite crystallites were formed. With
glasses of lower fluoride content (x < 1-0) gehlenite crystallites were also
found. As x decreased, increasingly more gehlenite was formed at the
expense of anorthite and fluorite. In connection with this observation it
should be noted that the chemical composition of the glass corresponds
to gehlenite when x = 0 and to a mixture of anorthite and fluorite when
x=\.
Wood and Hill consider that the role of fluoride in these glasses is
uncertain. Phase-separation studies suggest that the structure of the glass
might relate to the crystalline species formed, in which case a micro-
crystallite glass model is appropriate. But other evidence cited above on the
structure-breaking role of fluoride is compatible with a random network
model.

130
Glass polyalkenoate (glass-ionomer) cement

Table 5.10. Examples of practical glasses used in glass polyalkenoate


cement {Crisp, Abel & Wilson, 1979; Wilson & McLean, 1988; Brook,
Craig & Lamb, 1991)

ASPA IV ASPAX GC ESPE Pilkington


G-200 G-338 Fuji Ketac MP4
SiO2 301 24-9 41-9 34-9 30-8
A12O3 19-9 14-2 28-6 201 38-5
A1F3 2-6 110 1-6 8-8
CaF 2 34-5 12-8 15-7 20-8
CaO 28-6
NaF 3-7 12-8 9-3 3-6
Na 2 O 21
A1PO4 100 24-2 3-8 11-8
SiO 2 : A12O3 by mass 1 51 1-75 146 1-74 0-80

G-200 and G-338 are LGC formulations used in certain commercial materials
GC Fuji and ESPE Ketac are dental cements
MP4 is used in splint bandage materials

SiO2-Al2Oz-CaF2-AlPO^ glasses
Not much needs to be said on these glasses, except that one, G-235 (Table
5.8) has been subjected to solid state NMR (Ellison & Warrens, 1987). It
would appear that aluminium is mainly in tetrahedral coordination with
some octahedral and pentahedral coordination and that the structural
units are mainly Q 3 (1A1) and Q 4 (2A1).

glasses
These are practical glasses and some examples are given in Table 5.10. One
of the most important of these is G-200, an unusual glass in that it forms
a practical cement without the need for ( + )-tartaric acid. It was used in
early commercial materials and in fundamental studies on setting and
cement structure. The glass must now be regarded as atypical as it is very
high in fluoride and low in sodium. Barry, Clinton & Wilson (1979)
examined the glass structure. They found that it was heavily opal and
phase-separated with droplets rich in calcium and fluoride of complex
morphology. These droplets were of average size 1-7 \im and volume
fraction 20%. There were also massive inclusions of fluorite. When the
glass was fused at a higher temperature, 1300 °C as against 1150 °C,

131
Polyalkenoate cements
Table 5.11. Effect of ( + )-tartaric acid on glass polyalkenoate cement
properties

G-237 G-309 G-200

Tartaric acida Absent Present Absent Present Absent Present


Working time (23 °C), 10 10 1-2 1-7 1-7 2-7
minutes
Setting time (37 °C), 4-3 2-7 3-5 2-8 60 3-8
minutes
Setting rate, arbitrary 50 165 65 100 20 55
units
Compressive strength, 121 140 125 136 100 120
MPa

Cement-forming liquid: 45 % poly(acrylic acid)


Powder: liquid ratio of 3:1
a
Added in 5 % concentration

fluoride was lost, the morphology of the glass changed to one with smaller
droplets, and it became more reactive towards poly(acrylic acid).

5.9.3 Poly(alkenoic acid)s


The poly(alkenoic acid)s used in glass polyalkenoate cement are generally
similar to those used in zinc polycarboxylate cements. They are homo-
polymers of acrylic acid and its copolymers with itaconic acid, maleic acid
and other monomers e.g. 3-butene 1,2,3-tricarboxylic acid. They have
already been described in Section 5.3. The poly (aerylie acid) is not always
contained in the liquid. Sometimes the dry acid is blended with glass
powder and the cement is activated by mixing with water or an aqueous
solution of tartaric acid (McLean, Wilson & Prosser, 1984; Prosser et al.,
1984).
Increase in concentration of the polyacid increases solution viscosity,
quite sharply above 4 5 % by mass (Crisp, Lewis & Wilson, 1977). The
strength of glass polyalkenoate cements also increases, almost linearly,
with polyacid concentration. This is achieved at the cost of producing over-
thick cement pastes and loss of working time.
The molecular weight of the polyacid affects the properties of glass
polyalkenoate cements. Strength, fracture toughness, resistance to erosion
and wear are all improved as the molecular weight of the polyacid is

132
Glass polyalkenoate (glass-ionomer) cement

Table 5.12. Effect of the various tartaric acids on glass polyalkenoate


cement properties {Crisp, Lewis & Wilson, 1979)

Setting time, Compressive strength,


minutes (37 °C) MPa (24 hours)
None 8-5 104
(+ )-tartaric acid 50 112
( —)-tartaric acid 4-75 120
Meso-tartaric acid 10-25 72
Racemic tartaric acid 5-25 109

Powder: G-200
Liquid: 47-4% poly(acrylic acid), 5-3% tartaric acid

increased (Wilson, Crisp & Abel, 1977; Hill, Wilson & Warrens, 1989;
Wilson et al.91989). Setting is accelerated and working time is lost, and this
places a restriction on improving cement properties by this route. The
maximum molecular weight of the polyacid that can be used would appear
to be 75000.

5.9.4 Reaction-controlling additives


The glass polyalkenoate cement system was not viable until Wilson and
Crisp discovered the action of ( + )-tartaric acid as a reaction-controlling
additive (Wilson & Crisp, 1975,1976, 1980; Wilson, Crisp & Ferner, 1976;
Crisp & Wilson, 1976; Crisp, Lewis & Wilson, 1979). It may be regarded
as an essential constituent and is invariably included in glass polyalkenoate
cements as a reaction-controlling additive.
It affects the nature of the setting reaction profoundly and this subject is
discussed in Section 5.9.5. It sharpens set and increases the hardening rate,
without decreasing, and even sometimes increasing, working time. Strength
is also increased (Table 5.11). Crisp, Merson & Wilson (1980) found
moreover that the addition of ( + )-tartaric acid conferred the setting
property on a glass (G-288) that otherwise did not form a cement.
No other additive has the same effect although many alternatives were
examined by Wilson, Crisp & Ferner (1976) and Prosser, Jerome & Wilson
(1982). Other multifunctional carboxylic acids, including citric acid, had
little effect, apart from a slight tendency to shorten working time and
increase the setting rate. That the effect is a subtle one is shown by the fact

133
Polyalkenoate cements
Table 5.13. Effect of fluorides on glass polyalkenoate cement
compressive strength, MPa {Crisp, Merson & Wilson, 1980)

( + )-tartaric acid

Fluoride Absent Present

None 72
A1F3 91 112
MgF 2 101 158
SnF 2 128 145
ZnF. 111 128

Powder: G-247
Liquid: 50 % acrylic-itaconic acid copolymer, 5 % tartaric acid

that raeso-tartaric acid does not have the effect of sharpening the set
(Table 5.12).
Ethanolamines and polyphosphates slow the reaction down as a whole.
Both tetrahydrofurantetracarboxylic acid and polyphosphates are some-
times to be found in commercial examples.
Crisp, Merson & Wilson (1980) found that the addition of metal
fluorides to formulations had the effect of accelerating cement formation
and increasing the strength of set cements; the effect was enhanced by the
presence of ( + )-tartaric acid (Table 5.13). Strength of cements formed
from an SiO 2 -Al 2 O 3 -Ca 3 (PO 4 ) 2 glass, G-247, can be almost doubled by
this technique.

5.9.5 Setting
General
The setting and hardening reactions are first outlined in general terms.
They can be considered to take place in a number of overlapping stages.

(1) On mixing the cement paste, the calcium aluminosilicate glass is


attacked by hydrogen ions from the poly(alkenoic acid) and
decomposes with liberation of metal ions (aluminium and cal-
cium), fluoride (if present) and silicic acid (which later condenses
to form a silica gel).
(2) As the p H of the aqueous phase rises, the poly(alkenoic acid)
ionizes and most probably creates an electrostatic field which aids
the migration of liberated cations into the aqueous phase.

134
Glass polyalkenoate (glass-ionomer) cement
(3) As the poly(alkenoic acid) ionizes, polymer chains unwind as the
negative charge on them increases, and the viscosity of the cement
paste increases. The concentration of cations increases until they
condense on the polyacid chain. Desolvation occurs and insoluble
salts precipitate, first as a sol which then converts to a gel. This
represents the initial set.
(4) After gelation or initial set, the cement continues to harden as
cations are increasingly bound to the polyanion chain and
hydration reactions continue. Recent evidence suggests that a
siliceous hydrogel may be formed in the matrix.
In considering the setting reaction in more detail, cognizance must be
taken of the nature of the glass and the presence of reaction-controlling
additives. These affect both the nature of the cement-forming reaction and
setting characteristics. These effects stem from complex formation; the two
most important complexing agents are fluoride, derived from the glass, and
(4- )-tartaric acid, by far the most important of the reaction-controlling
additives. We distinguish four possible systems: (1) oxide glasses, (2) oxide
glasses with added (+ )-tartaric acid, (3) fluoride glasses and (4) fluoride
glasses with added (H-)-tartaric acid.

Cement formation with oxide glasses


Little more can be said concerning the cement-forming reactions between
oxide glasses and poly(alkenoic acid)s because they have not been studied.
It is almost impossible to prepare cements from oxide glasses and a
polyacid, because the paste formed on mixing is intractable. The evidence
points toward the premature bonding of aluminium ions as the cause (Ellis
& Wilson, 1987). This does not occur in the related dental silicate cement,
which is formed using a concentrated solution of orthophosphoric acid;
here aluminium forms complexes with orthophosphoric acid, which delays
precipitation. Practical cements can be formed from oxide glasses if
( + )-tartaric acid is added to the system. Since aluminium forms soluble
complexes with ( + )-tartaric acid it is reasonable to suppose that the
formation of these complexes prevents the premature ion-binding of
aluminium to poly(alkenoic acid) and so allows workable cement pastes to
be formed. This view is supported by the solution studies of Ellis & Wilson
(1987).

135
Polyalkenoate cements

Table 5.14. Infrared spectroscopic bands of reference carboxylates


{Nicholson et aL, 1988b)

C-O stretch of salt, cm"1

Salt asymmetric symmetric

Ca-PAA 1550 1410


Al-PAA 1599 1460
Ca-tartrate 1595 1385
Al-tartrate 1670 1410

Cement formation with fluoride glasses - no tartaric acid


Cements can be formed using fluoride-containing glasses in the absence of
( + )-tartaric acid, but high-fluoride glasses have to be used to obtain
cements with sufficient working time (Crisp & Wilson, 1976; Kent, Lewis
& Wilson, 1979). Only very few glasses, those very high in fluoride, can be
used in practical cements. Fluoride clearly has a considerable effect on the
reaction, probably because it forms strong soluble complexes with
aluminium such as A1F2+ and A1F+; these have been reported by Connick
& Poulsen (1957), O'Reilly (1960), and Akitt, Greenwood & Lester (1971).
These complexes probably prevent the premature gelation of the poly-
anions by aluminium ions.
In this system there is a useful cooperative effect between aluminium,
fluoride and calcium, which has been demonstrated by the solution studies
of Ellis & Wilson (1987). In the absence of aluminium, calcium precipitates
as the fluoride at all pHs. Aluminium has the effect of preventing the
precipitation of calcium as fluoride, again because it forms strong soluble
complexes with fluoride.
Detailed studies of the cement-forming reaction using fluoride-con-
taining glasses and aqueous solutions of poly(acrylic acid) have been
carried out; mainly by Wilson and his coworkers (Crisp & Wilson,
1974a,b, 1976; Crisp et aL, 1974; Barry, Clinton & Wilson, 1979; Prosser,
Richards & Wilson, 1982; Hill & Wilson, 1988b; Nicholson et aL, 1988b),
but the work of Cook (1982, 1983a) should also be noted.
The following account is based mainly on the early studies of Crisp,
Wilson and coworkers (Crisp & Wilson, 1974a,b, 1976; Crisp et aL, 1974)
who used glass G-200 and 50 % solutions of poly(acrylic acid), and is
supported by the later studies of Nicholson et aL (1988b) using glass G-
309. The compositions of materials used are shown in Table 5.8.

136
Glass polyalkenoate (glass-ionomer) cement
These workers applied chemical and infrared spectroscopic methods to
study cement-formation. Infrared spectroscopy exploits the fact that
calcium polyacrylate and aluminium polyacrylate give rise to different
carboxylate bands (Table 5.14). On mixing the powder and liquid,
hydrogen ions from the poly(acrylic acid) solution rapidly attack the glass
particles, which are decomposed to silicic acid, and Al3+, Ca2+, Na+ and F~
ions are released (Crisp & Wilson, 1974a,b; Cook, 1983c; Crisp, Lewis &
Wilson, 1976d; Wasson & Nicholson, 1990). Originally, it was supposed
that this attack occurred only at the surface layer of the glass particles
(Barry, Clinton & Wilson, 1979) but later observations using 29Si NMR by
Ellison & Warrens (1987) suggest that attack occurs throughout the glass
particles. In the case of glass G-235 pentacoordinated aluminium dis-
appeared and Q4 (2A1) units were converted to Q3 (1A1) units, i.e.
aluminium was lost from the glass structure.
Since hydrogen ions are six to twelve times more mobile than other
cations, there will be a delay between loss of hydrogen ions from solution
and migration of glass cations into the aqueous phase. Presumably, this
electrical imbalance results in an electricfieldwhich acts as a driving force
for the migration of cations. Aluminium and fluoride are almost certainly
transported as cationic aluminofluoride complexes, A1F2+ and A1FJ,
mentioned above.
Figure 5.11 (Crisp & Wilson, 1974b) shows the time-dependent variation
of the concentration of soluble ions in setting and hardening cements. Note
that the concentrations of aluminium, calcium andfluoriderise to maxima
as they are released from the glass. After the maximum is reached the
concentration of soluble ions decreases as they are precipitated. Note that
this process is much more rapid for calcium than for aluminium and the
sharp decline in soluble calcium corresponds to gelation. This indication is
supported by information from infrared spectroscopy which showed that
gelation (initial set) was caused by the precipitation of calcium poly-
acrylate. Thisfindingwas later confirmed by Nicholson et al. (1988b) who,
using Fourier transform infrared spectroscopy (FTIR), found that calcium
polyacrylate could be detected in the cement paste within one minute of
mixing the cement. There was no evidence for the formation of any
aluminium polyacrylate within nine minutes and substantial amounts are
not formed for about one hour (Crisp et al., 1974).
Crisp & Wilson (1974b, 1976) attributed the slowness of binding in the
case of aluminium to several effects: preferential leaching of calcium ions,
lack of mobility of the hydrated or multinuclear aluminium species

137
Polyalkenoate cements
(Aveston, 1965; Akitt, Greenwood & Lester, 1971; Waters & Henty, 1977)
and steric problems because of the triple charge on the ion. However, when
it is remembered that the Al3+ ion is responsible for the premature gelation
of the oxide glass system, it may be that the true explanation lies in the
stability of the fluoride complexes or the formation of large multinuclear
aluminium complexes. Such complexes have been reported by a number
of workers - Aveston (1965), Akitt, Greenwood & Lester (1971) and
Waters & Henty (1977) - and may be represented by the generic formula
[Alx(OH,F)y(H2O)z]{3xy). The presence of such complexes may explain
the slow reactivity of the Al3+ ion with EDTA in slightly acid solutions.
Of course, this argument does not apply to the premature binding of Al3+
ions in the dental glass system when the pH is low.
Gelation involves an extended structure and some type of linking
between chains. The concept of salt-like crosslinks has already been
described (Section 5.5). Other possibilities may be considered. Hill, Wilson
& Warrens (1989) examined the possibility that chain entanglements might
account for the strength of polyelectrolyte cements. They used in particular

6 -

5 -
-0-v
A. 1
1
t
y • • • • • • • • • • . . .
••

3
\ ••
*•
F *'
2 -

1 -

PA
1 1 1 1 1 Mil 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 1 J

10 100 1000
time, min
Figure 5.11 The time-dependent variation, in setting and hardening cements, of the
concentration of soluble ions Al 3+ , Ca 2+ , F~ and VO\~ (expressed as P 2 O5). These ions are
released from the glass powder into the cement matrix (Crisp & Wilson, 1974b).

138
Glass polyalkenoate (glass-ionomer) cement
the reptation (chain pull-out) model described by de Gennes (1979) and
Edwards (1969). Here a polymer chain is considered to be trapped by a
tube of entanglements formed by neighbouring chains, and strength is
related to the forces required to pull out the trapped chain. They found,
however, that this model did not account for the effect of polyacid
molecular mass on cement strength (Nicholson, 1992).
After initial set the cement continues to harden and strengthen (Crisp,
Lewis & Wilson, 1976b) as cations are increasingly bound to the polyanion
chain (Cook, 1983c) and hydration processes continue (Wilson, Paddon &
Crisp, 1979; Wilson, Crisp & Paddon, 1981). There are still free COOH
groups present after 24 hours of reaction (Nicholson et al., 1988b). Cook
(1983c) observed that the transfer of aluminium and calcium from the glass
to the matrix continued for at least five weeks, during which time both
strength and modulus increased (Paddon & Wilson, 1976). The increase in
strength after set is illustrated by the results of Elliot, HoUiday & Hornsby
(1975) presented in Figure 5.12. The reaction probably never ceases
entirely, for Crisp, Lewis & Wilson (1976b) have observed that strength
continues to increase, logarithmically, for at least a year if specimens are
stored in kerosene. This slow increase in strength has been attributed to
hydration processes and the slow diffusion of cement-forming cations,
especially aluminium, seeking anionic sites.
There remains the question of the role of silica. This aspect of setting

200 r

o
\ 150 -

o
I
«» 100

a 50
E
o

1.0 10 100 1000


Time, hours
Figure 5.12 The time-dependent increase in strength after set (Elliott, HoUiday & Hornsby,
1975).

139
Polyalkenoate cements
chemistry has been neglected and there are some ambiguities in the
literature. Recently, Wasson & Nicholson (1990, 1991) have revived
interest in this topic. The release of orthosilicic acid, which accompanies
the release of glass ions, has been observed and is implicit in the theory of
acid decomposition of the glass (Section 5.9.2). During the setting reaction
orthosilicic acid is converted to silica gel (Crisp et aL, 1974). Crisp, Lewis
and Wilson (1976d) found that the reduction in the amount of silicic acid
eluted as a cement aged was matched by similar reductions in amounts of
cations and anions eluted.
Age of cement, minutes 7 60 1440
Silica (SiO2) eluted mg/g cement 3-39 0-55 010
Aluminium (A12O3) eluted mg/g cement 2-07 0-24 003
Sodium (Na2O) eluted mg/g cement 3-31 1-27 0-49
Fluoride (F) eluted mg/g cement 2-59 0-73 0-73
The insolubilization of cations and anions during the setting and hardening
process is thus paralleled by that of silica. Under acid conditions
orthosilicic acid condenses first to form polymeric silicic acid and then
silica gel (Her, 1979; Andersson, Dent Glasser & Smith, 1982). These
processes are discussed more fully in Section 6.5.4. Gelation of silica, like
the formation of salt gels, is enhanced by a reduction in the acidity of
solutions.
Much of the siliceous hydrogel formed is found enveloping the glass
particles, and the attacked glass particles maintain their original mor-
phology (Barry, Clinton & Wilson, 1979). The mechanism of this process
is obscure and it is uncertain whether it results from an ion depletion
process or deposition of a siliceous gel on the site of the decomposed glass
network. Recently, Wasson & Nicholson (1990, 1991) have suggested that
a hydrated silicate is formed in the matrix and may contribute to the
hardening process. The composition of this species is at present unknown.
It could be a silicate gel, similar to tobermorite gel found in Portland
cement (Taylor, 1966) which could account for the long term increases in
strength of glass-ionomer cement. Alternatively, it could be a type of silica
gel. Silica gel cements are known: they are weak with a compressive
strength of 14 MPa, but are acid-resistant (Trautschold, 1940). In this
connection, it may be noted that glass-ionomer cements are much more
resistant to acid attack than are the related zinc polycarboxylates. Clearly,
further studies are required into this aspect of glass-ionomer cement
chemistry.

140
Glass polyalkenoate (glass-ionomer) cement

Cement formation with fluoride glasses - ( + )-tartaric acid


The presence of ( + )-tartaric acid in a cement formulation exerts a
profound effect on the cement-forming reaction. The nature of the
underlying chemical reaction is changed and this is reflected in time-
dependent changes in viscosity.
Both Cook (1983a,b) and Hill & Wilson (1988b) have studied the effect
of (+ )-tartaric acid on the development of viscosity. Hill & Wilson used
two glasses in their study, an oxide glass, G-287 (Table 5.6), and a fluoride
glass, G-307 (Table 5.8). They noted that in the absence of ( + )-tartaric
acid there was an almost exponential increase in apparent viscosity with
time and that this effect was exaggerated when small amounts of ( + )-
tartaric acid, 0-3 %, were added to the liquid (Figure 5.13). However, when
added in larger amounts ( + )-tartaric acid reduced the apparent viscosity
of the pastes and also delayed the increase in apparent viscosity for a
period of time which depended on the amount of ( + )-tartaric acid added;
this plateau is very evident in the curves of apparent viscosity against time.
Clearly, there are competing reactions taking place.
The early reaction leading to gelation has been studied by Prosser,
Richards & Wilson (1982) who used 13C NMR to examine the interaction
between glass G-200 and poly(acrylic acid) solutions, and by Nicholson et

Viscosity (+) tartaric acid

no (+) tartaric acid

Time 1
Figure 5.13 Effect of tartaric acid on viscosity development (Hill & Wilson, 1988b).

141
Polyalkenoate cements
al (1988b) who used FTIR and glass G-309 (Table 5.8). The addition of
( + )-tartaric acid totally changes the chemistry of the cement-forming
reaction. It reacts preferentially with the glass because it is the stronger acid
and its complexes are fully formed at pH = 3-4 while those of poly (aery lie
acid) only appear at higher pH. The formation of calcium polyacrylate is
suppressed, the extent of this suppression being dependent on the amount
of ( + )-tartaric acid present. Instead, within the first minute, calcium
tartrate makes a transient appearance most probably causing gelation. It
disappears within nine minutes as aluminium polyacrylate appears.
From these experiments, it appears that ( + )-tartaric acid prolongs
working time by preventing the premature formation of calcium poly-
acrylate, and later sharpens set and accelerates hardening by enhancing the
rate at which aluminium polyacrylate is formed. By contrast, while meso-
tartaric acid delays the premature formation of calcium polyacrylate it
does not enhance the rate of formation of aluminium polyacrylate. Thus,
it prolongs working time without accelerating hardening. The effects of

Figure 5.14 The microstructure of the set cement is clearly revealed by Nomarski reflectance
optical microscopy. Glass particles are distinguished from the matrix by the presence of
etched circular areas at the site of the phase-separated droplets (Barry, Clinton & Wilson,
1979).

142
Glass polyalkenoate (glass-ionomer) cement

(H-)-tartaric acid and raeso-tartaric acid on the physical properties of


glass-ionomer cements described in Section 5.9.4 are explained in terms of
the underlying chemistry.

5.9.6 Structure
Barry, Clinton & Wilson (1979) examined the structure of cements
prepared from a glass powder from which very fine particles had been
removed to improve resolution. The microstructure of the set cement is
clearly revealed by Nomarski reflectance optical microscopy (Figure 5.14).
Glass particles are distinguished from the matrix by the presence of etched
circular areas at the site of the phase-separated droplets. The micrograph

Figure 5.15 More detail than seen in Fig. 5.14 is obtained in a scanning electron image. The
reacted glass particles are covered by a distinct reaction layer of silica gel (Barry, Clinton &
Wilson, 1979).

143
Polyalkenoate cements

is similar to that of the dental silicate cement (Section 6.5.5). More detail
can be seen in a scanning electron image (Figure 5.15). The reacted glass
particles are covered by a distinct, gel-like reaction layer which is identified
as silica gel. This layer has detached itself from the glass core, which,

Figure 5.16 (a) An electron image of a glass-ionomer cement; (b) SL diagrammatic


representation of this image: unshaded areas are glass particles, lightly shaded areas are
cement matrix, and deeply shaded areas are voids or cracks. The remaining micrographs
show the Ka characteristic radiation of (c) Ca, (d) Si, (e) Al, and (/) Si: Al (Barry, Clinton &
Wilson, 1979).

144
Glass polyalkenoate {glass-ionomer) cement
according to Brune & Smith (1982), is consistent with weak hydrogen-
bonding between layer and core. Fluorite particles appear as bright
highlights because any calcium present gives rise to enhanced electron
scattering.
Barry, Clinton & Wilson (1979), using a scanning electron microscope
fitted with an X-ray analyser, obtained an element distribution map for Ca,
Si and Al (Figure 5A6c,d,e). They found that whereas the glass particles
contained major amounts of all three elements, the matrix contained only
aluminium and calcium in major amounts, with more aluminium than
calcium; there were only minor amounts of silicon. Interestingly, element
distributions showed that the glass particles appeared slightly larger when
judged by silicon radiation than by aluminium radiation. A map of the
Si/Al ratio confirmed this observation. An electron image of a glass-
ionomer cement specimen is shown in Figure 5.16a together with an Si: Al
distribution map, Figure 5.16/. At the time, this effect was attributed to the
depletion of ions from the outer layer of the glass particles, but now we are
more inclined to believe that this layer is really due to an outgrowing of
silica gel, which, of course, has the capacity to absorb metal ions (Her,
1979; Hazel, Shock & Gordon, 1949). Certainly the siliceous gel layer in
Figure 5.15 does have the appearance of a reaction zone rather than that
of a relict. Moreover, an aluminosilicate glass matrix is decomposed to
silicic acid (Wasson & Nicholson, 1990, 1991), which is detected in
significant amounts in young cements (Crisp, Lewis & Wilson, 1976d).
The latter interpretation of data is more in accord with the recent 27A1
and 29Si NMR findings of Ellison & Warrens (1987), who found that the
structure of an appreciable fraction of the glass changed under acid attack
with some loss of aluminium including all in fivefold coordination (see
Section 5.9.2). Thus, acid attack was not entirely confined to the surface
layer of a glass particle. If this is so then silicic acid as well as ions must
migrate from the body of the particle and it is reasonable to suppose that
silicic acid deposits as siliceous gel at the particle-matrix interface.
The picture of cement microstructure that now emerges is of particles of
partially degraded glass embedded in a matrix of calcium and aluminium
polyalkenoates and sheathed in a layer of siliceous gel probably formed
just outside the particle boundary. This structure (shown in Figure 5.17)
was first proposed by Wilson & Prosser (1982, 1984) and has since been
confirmed by recent electron microscopic studies by Swift & Dogan (1990)
and Hatton & Brook (1992). The latter used transmission electron
microscopy with high resolution to confirm this model without ambiguity.

145
Polyalkenoate cements
The form of silica in the matrix is at present unknown. In the freshly
prepared cement there are appreciable amounts of silicic acid present
which decline as the cement ages (Crisp, Lewis & Wilson, 1976d). In the set
cement silica could be present as a polymeric silicic acid, a siliceous gel or
even a hydrated silicate gel, such as the tobermorite gel present in Portland
cements (Taylor, 1966).
The molecular structure is mainly a matter of conjecture but the
coordination state of aluminium in the matrix is known to be six (Ellison
& Warrens, 1987). Therefore, there must be three monovalent anions,
COO", F" or OH", and three neutral H2O. This matter has been discussed
in detail in Section 5.5. As noted there, Al3+ has the potential to link three
chains but this is sterically unlikely. Also, as Mehrotra & Bohra (1983)
have pointed out, tricarboxylates tend to hydrolyse into complex
structures that involve Al-O-Al bridges. Nevertheless we must note
that glass polyalkenoate cement loses its original plastic behaviour and
becomes increasingly rigid as it ages over the weeks following set. Such
time-dependent changes are not found with other cements and it is
tempting to speculate that they may arise from the slow transformation
in the nature of the Al bonding.

Rolysalt matrix

Figure 5.17 Diagrammatic representation of a glass-ionomer cement (Laboratory of the


Government Chemist: Crown copyright reserved).

146
Glass polyalkenoate (glass-ionomer) cement

5.9.7 General characteristics


The glass polyalkenoate cement uniquely combines translucency with the
ability to bond to untreated tooth material and bone. Indeed, the only
other cement to possess translucency is the dental silicate cement, while the
zinc polycarboxylate cement is the only other adhesive cement. It is also an
agent for the sustained release of fluoride. For these reasons the glass
polyalkenoate cement has many applications in dentistry as well as being
a candidate bone cement. Its translucency makes it a favoured material
both for the restoration of front teeth and to cement translucent porcelain
teeth and veneers. Its adhesive quality reduces and sometimes eliminates
the need for the use of the dental drill. The release of fluoride from this
cement protects neighbouring tooth material from the ravages of dental
decay. New clinical techniques have been devised to exploit the unique
characteristics of the material (McLean & Wilson, 1977a,b,c; Wilson &
McLean, 1988; Mount, 1990).
This cement also has a low setting exotherm, lower than any other
aqueous dental cement (Crisp, Jennings & Wilson, 1978), which means
that it can be mixed swiftly as there is no need to dissipate heat. This
property also gives it an advantage over bone cements based on modified
poly (methyl methacrylate) which have high exo therms.

5.9.8 Physical properties


The glass polyalkenoate cement sets rapidly within a few minutes to form
a translucent body, which when young behaves like a thermoplastic
material. Setting time (37 °C) recorded for cements mixed very thickly for
restorative work varied from 2-75 to 4-7 minutes, and for the more thinly
mixed luting agents from 4-5 to 6-25 minutes. Properties are summarized in
Table 5.15.
Strength develops rapidly and after 24 hours in water (37 °C) can reach
225 MPa (compressive) and 39 MPa (flexural) (Williams & Billington,
1989; Pearson & Atkinson, 1991; Pearson, 1991). Compressive modulus
reaches 9 to 18 GPa after 24 hours (Paddon & Wilson, 1976; Wilson,
Paddon & Crisp, 1979).
Crisp, Lewis & Wilson (1976a) found that for two early types of glass-
ionomer cement (ASPAII and ASPAIV) compressive strength continued
to increase for at least a year. Recently, Williams & Billington (1989) have
found that this behaviour does not hold for all modern commercial

147
Polyalkenoate cements

Table 5.15. Mechanical properties of glass polyalkenoate cement (Prosser


et al. 1984; 0ilo, 1988; Paddon & Wilson, 1976; Wilson, Paddon &
Crisp, 1979; Pearson & Atkinson, 1991; Williams & Billington, 1989)

Filling Luting
materials agents

Powder .liquid, g cm 3
Conventional 20-3-4 1-67-1-83
Water hardening 6-7-7-2 3-3-3-6
Consistency disc diameter, mm 18-33 21-31
Maximum particle size, urn — 20-40
Film thickness (2 min), urn — 24-40
Working time (23 °C), minutes 1-3-3-8 2-3-5-75
Setting time (37 °C), minutes 2-75-4-7 4-5-6-25
Wet compressive strength (24 h), MPa 140-225 82-162
Wet compressive modulus (24 h), GPa 9-18 —
Wet flexural strength (24 h), MPa 8-9-391 4-1-15-5
Wet tensile strength, MPa 9-0-19-3 5-3-10-9
Creep (24-48 h), % 019-0-33 0-32-1-37
Opacity, C 07 (1 mm) 0-44-0-84 0-67-0-88
Water leachables (7 min), % 0-29-2-12 0-9-3-2
Water leachables (1 h), % 0-13-0-70 0-3-1-0

Load: 2-5 Kgf for filling materials, 220 gf for luting agents, applied after
2 minutes

materials. Thus, whereas the compressive strength of Opusfil W increased


steadily from 225 MPa (after 24 hours) to 258 MPa (after 115 days), that
of Ketac Fil increased from 171 MPa (after 24 hours) to a peak of 262 MPa
(after 50 days) and, thereafter, declined to 167 MPa (after 115 days).
The young glass polyalkenoate cement is similar to zinc polycarboxylate
cement in that it is not as rigid as dental silicate cement and shows marked
stress relaxation characteristics (Paddon & Wilson, 1976). However, as the
glass polyalkenoate cement ages its properties progressively depart from
those of the zinc polycarboxylate and approach those of the dental silicate
cement. It loses its viscoelastic properties and becomes increasingly rigid.
This is shown by a significant increase in modulus and decline in stress
relaxation (Figure 5.18).

148
Glass polyalkenoate {glass-ionomer) cement

Fracture toughness
Flexural strength and fracture toughness are clinically more significant
than compressive strength. The flexural strength of a glass-ionomer
cement can reach 39 MPa after 24 hours (Pearson & Atkinson, 1991) which
is a much higher value than that attained by any dental silicate cement.

N/mm2

ExiO3 a

Cement Age
Figure 5.18 Thisfigureshows how the properties of a glass polyalkenoate cement change as
it ages. S is the compressive strength, E the modulus, a* a stress-relaxation function, and e*
a strain-conversion function from elastic to plastic strain (Paddon & Wilson, 1976).

149
Polyalkenoate cements

Table 5.16. Strength and fracture toughness of glass polyalkenoate filling


materials (Seed & Wilson, 1980; Prosser et ai, 1984; Lloyd & Mitchell,
1984; Goldman, 1985; Prosser, Powis & Wilson, 1986; Lloyd &
Adamson, 1987)

Flexural Fracture
strength toughness
a, MPa # l c ,MNm 3 / 2

Glass polyalkenoate cement 9-30 0-45-0-55


Phase-dispersed glass polyalkenoate 38 —
cement
Carbon fibre reinforced glass 53 —
polyalkenoate cement
Silver-glass polyalkenoate cement 36 0-35
Dental silicate cement 25 0-12-0-30
Anterior composite resins 29-49 0-63-1-84
Posterior composite resins 74^129 0-95-2-0
Dental amalgam 76 0-74

However, these values are less than those recorded for composite resins
used in dentistry. Goldman (1985) reports values of 29 to 49 MPa for
anterior composite resins and Lloyd & Adamson (1987) values of 76 to
125 MPa for posterior composite resins. A typical amalgam has a flexural
strength of 6 MPa (Lloyd & Adamson, 1987) (Table 5.16). However, the
flexural strengths of some glass-ionomer cements increase with time and
values as high as 59 MPa (after 3 months) and 70 MPa (after 7 days) have
been reported (Pearson & Atkinson, 1991).
Fracture toughness values for glass polyalkenoate cement vary from
0-25 to 0-55 MN m3/2 (Lloyd & Mitchell, 1984; Goldman, 1985; Lloyd &
Adamson, 1987). The values are generally higher than those found for the
traditional dental silicate cement but lower than those found for anterior
composite resins (Lloyd & Mitchell, 1984; Goldman, 1985) and much
lower than those for posterior composite resins and dental amalgams
(Lloyd & Adamson, 1987).
These low values for flexural strength and fracture toughness compared
with the values for composite resins and dental amalgams make the
glass-ionomer cement less suitable than these materials in high-stress
situations.

150
Glass polyalkenoate (glass-ionomer) cement

Translucency
A restorative material can be used for the aesthetic restoration of the front
(anterior) teeth only if it is as translucent as tooth enamel. This is because
colour matching depends on translucency as well as hue and chroma.
The glass polyalkenoate cement is translucent and so can be colour-
matched to enamel (Kent, Lewis & Wilson, 1973; McLean & Wilson,
1977a,b,c; Crisp, Abel & Wilson, 1979; Wilson & McLean, 1988); see
Figure 5.19. Early glass polyalkenoate cements were significantly deficient
in this quality because high-fluoride glasses had to be used before other
means of controlling the cement-forming reaction were discovered. These
glasses were heavily phase-separated and almost opaque. This has been
remedied in recent formulations by employing clear or slightly opalescent
glasses in combination with reaction-controlling additives.
Another barrier to achieving translucency is mismatch between the
refractive indices of the glass and the matrix; the refractive index of the
glass is greater than that of the matrix, which causes light-scattering. The
dental silicate cement tends to be naturally more translucent than the glass

Figure 5.19 The translucent appearance of glass polyalkenoate cements when placed on a
black and white striped background.

151
Polyalkenoate cements
polyalkenoate cement because the refractive index of a phosphate matrix
is greater than that of a polyacrylate one. However, by incorporating large
amounts of leachable phosphate into a polyalkenoate cement glass it has
been possible to increase matrix refractive index and to produce fully
aesthetic glass polyalkenoate cements (Crisp, Abel & Wilson, 1979).
The translucency of dental materials is normally represented as the
inverse of opacity, although the scattering coefficient is a more fundamental
property (Section 10.11). Opacity is equivalent to contrast ratio, which is
the ratio of the light reflected from a disc of cement (1 mm thick) placed on
a black background to that when it is placed on a white background. The
reflectivity of this background used in the dental context is 70 % and
opacity is reported as C0.7 values (Paffenbarger, 1937). The C0.7 values for
the enamel of incisor teeth vary from 0-31 to 0*67 (Paffenbarger,
Schoonover & Souder, 1938) and it is generally accepted that an aesthetic
filling material should have a C0.7 between 0-35 and 0-50.
The first glass polyalkenoate cement had a C0.7 of 0-76, which was far too
high, but improved modern materials are more acceptable and a value as
low as 0-52 has been reported for one of these (Crisp, Abel & Wilson, 1979).
Knibbs, Plant & Pearson (1986b) have found that most glass polyalkenoate
cements have a good optical match with tooth enamel.

5.9.9 Adhesion
Bonding to tooth material and metals
The glass polyalkenoate cement has the important property of adhering to
untreated enamel and dentine as many workers have shown (Wilson &
McLean, 1988; Lacefield, Reindl & Retief, 1985). It also appears to adhere
to bone and base metals (Hotz et al., 1977).
The bond strength to enamel (2-6 to 9-9 MPa) is greater than that to
dentine (1-5 to 4-5 MPa) (Wilson & McLean, 1988). Bond strength develops
rapidly and is complete within 15 minutes according to van Zeghbroeck
(1989). The cement must penetrate the acquired pellicle (a thin mucous
deposit adherent to all surfaces of the tooth) and also bond to debris of
calciferous tooth and the smear layer present after drilling. Whatever the
exact mode of bonding to tooth structure, the adhesion is permanent. The
principles and mechanism of adhesion have already been discussed in
Section 5.2.

Various attempts have been made to improve bonding by pre-

152
Glass polyalkenoate (glass-ionomer) cement

conditioning the tooth surfaces with surface conditioners. The earliest


conditioner, citric acid solution, was proposed by Hotz et al. (1977). It
etches the surface of enamel, revealing its prismatic structure, and removes
calciferous debris (Figure 5.20a,&). It removes the smear layer from dentine
and opens up dentinal tubules (Figure 5.20c,d). Doubt was cast on its
efficacy and since then more effective surface conditioners have been
found. These include poly(acrylic acid), which has a less drastic effect on
tooth material than citric acid, and tannic acid, which forms reaction layers
on both dentine and enamel (Powis et al., 1982). Surface cleansing agents
and the fluoride ion also improve adhesion. Prati, Nucci & Montanari

Figure 5.20 The effect of a citric acid solution on tooth structure: (a) enamel surface before
application, (b) enamel surface after application showing etching, (c) dentine surface before
application, (d) dentine surface after application showing the opening-up of the dental
tubules (Powis et al, 1982).

153
Polyalkenoate cements
(1989) have reviewed recent studies on surface conditioners using modern
glass polyalkenoate cements. They found that bond strength depended on
both the cement and the surface conditioner used. The present consensus
favours treatment with a solution of poly(acrylic acid).
Permanent adhesion is an important attribute in a restorative material
for it demands only minimal cavity preparation (i.e. drilling etc.) as there
is no need to provide a retentive undercut. In cervical erosion lesions (small
cavities at the gum line) it is especially important not to enlarge the lesion
by drilling and in this situation the glass polyalkenoate cement is the
material of choice. In addition, the glass polyalkenoate cement provides an
excellent seal because it is adhesive and shows little or no microleakage
compared with composite resins (Hembree & Andrews, 1978; Welsh &
Hembree, 1985; Powis, Prosser & Wilson, 1988). This quality accounts for
the biocompatibility of glass polyalkenoate cement because a good seal
eliminates bacterial invasion at the interface between cavity wall and
restoration. The biological consequences of this are described in Section
5.9.11.

Molecular attachment

Acid etch attachment

Enamel

Figure 5.21 The laminate restoration, showing the glass polyalkenoate cement as a dentine
substitute and a composite resin as an enamel substitute.

154
Glass polyalkenoate (glass-ionomer) cement

Bonding to composite resins


An idea with important clinical implications to dentistry was advanced by
McLean & Wilson (1977b). This was to use a laminate restoration
composed of a glass polyalkenoate cement and a composite resin. In this
technique the glass polyalkenoate cement and the composite resin are used
together in a laminate restoration - the glass polyalkenoate cement as a
dentine substitute and the composite resin as an enamel substitute (Figure
5.21). The glass polyalkenoate cement provides adhesion to dentine and
fluoride release. The overlying composite resin provides excellent aesthetics
and wear resistance and is bonded both to the enamel and the glass
polyalkenoate cement by acid etching. The advantages of both materials
are obtained with the laminate restoration.
In recent years this concept has been revived and combined with the idea
of bonding the composite resin to the glass polyalkenoate cement by using
an acid-etch technique to provide micromechanical attachment (McLean
et ah, 1985; Wilson & McLean, 1988). In this technique the surface of a set
glass polyalkenoate cement is etched using a solution of orthophosphoric
acid, and a thin coat of mobile resin is allowed to flow into the etched
surface and is then polymerized by light activation. The composite resin is
then bonded to this surface. The appearance of the etched cement surface
is shown in Figure 5.22. The bond strength appears to be about the same
as the cohesive strength of the cement; McLean et al. (1985) reported a
value of 10 MPa.
There is, however, a considerable variation in the strength of the union,
and the combination of glass polyalkenoate cement and composite resin
has to be selected with considerable care in order to achieve good results
(Hinoura, Moore & Phillips, 1987; Mount, 1989; Prati, Nucci &
Montanari, 1989). The strength of the union depends on several factors,
including the strength of the cement, its rate of development and the ability
of the bonding resin to wet the cement (Mount, 1989). There are problems
of adaptation with heavily filled composite resins, and excessive poly-
merization contraction can destroy the bond. So far in these laminates
leakage between the glass polyalkenoate cement and the dentine has not
been completely eliminated, but the use of surface pre-treatments has
significantly reduced its extent (Prati, Nucci & Montanari, 1989).

155
Polyalkenoate cements

5.9.10 Erosion, ion release and water absorption


Although the glass polyalkenoate cement is the most durable of all dental
cements it is susceptible to attack by aqueous fluids under certain
conditions. There are three related phenomena to consider: erosion, ion
release and water absorption.

Ion release and water absorption


When fully hardened, the cement is resistant to erosion provided the
solution has a pH above 4. However, the glass polyalkenoate cement is
susceptible to erosion immediately after set because some of the matrix-
forming cations and anions are still in soluble form. In fact, the hardening
process is one where these cations and anions continue to precipitate. For
this reason these cements have to be protected, temporarily, by a varnish.
When immature glass polyalkenoate cements are exposed to neutral
solutions, such as normal saliva, they release ions and absorb water. The

Figure 5.22 Effect of acid-etching on the surface of a glass polyalkenoate cement (McLean et
aL, 1985).

156
Glass polyalkenoate (glass-ionomer) cement
matrix-forming cation Al3+, but not Ca2+, can be lost and then the cement
is permanently damaged. Other ions lost are sodium andfluoride,together
with silicic acid, but loss of these species does not seem to be significant as
far as erosion is concerned. Water is also rapidly absorbed: 3-2% in 24
hours and 3-8 % in 7 days in one example cited by Crisp, Lewis & Wilson
(1980).
As the cement ages, absorption of water and loss of aluminium ions
ceases (after 7 days). Other species - sodium and fluoride ions and silicic
acid - continue to be eluted. The release of fluoride is important, for the
glass polyalkenoate cement can be seen as a device for its sustained release.

Fluoride release
A number of workers have observed the sustained release of fluoride from
the glass polyalkenoate cement (Crisp, Lewis & Wilson, 1976d; Forsten,
1977; Maldonado, Swartz & Phillips, 1978; Causton, 1981; Cranfield,
Kuhn & Winter, 1982; Swartz, Phillips & Clark, 1984; Tay & Braden,
1988; Wilson, Groffman & Kuhn, 1985). The last-named found that
release continued for at least 18 months. Release can be fitted to a log-log
plot (Cranfield, Kuhn & Winter, 1982) although it is difficult to give a
physical meaning to this expression. Wilson, Groffman & Kuhn (1985)
fitted accumulated release of fluoride, sodium and silica to an equation of
the following form:
Total amount released = C+At*+Bt (5.1)
The three terms correspond respectively to initial washout, diffusion
and erosion. Unfortunately, although the mathematical fit was good
(c. 99-9 %), C and B proved to be negative, making it difficult to assign
a physical meaning to the equation.
Wilson, Groffman & Kuhn (1985) calculated that only about 4-5% of
the totalfluorideis available for release, and Meryon & Smith (1984) found
that the amount released was not related to the fluoride content of the
cement. Wilson, Groffman & Kuhn (1985) observed that release of fluoride
was accompanied by the release of sodium, necessary to maintain
electroneutrality, and silica (as silicic acid). The release of these species
could also be fitted to equation (5.1).
Crisp, Lewis & Wilson (1980) found that these same three species were
released, in greater amounts but in roughly the same proportions, under
acid attack. The association of silica with fluoride suggests, perhaps, that
it is principally the glass particles that are attacked rather than the matrix

157
Polyalkenoate cements
phase. Cranfield, Kuhn & Winter (1982) also found that the rate of release
offluorideis greater in acid than in neutral solution. These results suggest
that release of these species may be an erosive rather than a diffusive
process, which may explain the perplexity of Cranfield, Kuhn & Winter
(1982) in attempting to elucidate the mechanism offluoriderelease in terms
of a diffusive process. Kuhn & Jones (1982) suggested that a porous
granular monolith as described by Kydonieus (1980) was appropriate to
describe fluoride release. More recently, Tay & Braden (1988), in a
prolonged study over 2\ years, suggested that there were two processes
involved, a rapid surface elution and a slower bulk diffusion. But we must
conclude, at present, that the mechanism offluoriderelease is still far from
understood.
The release offluorideis biologically important because it is taken up by
adjacent tooth material (Retief et al, 1984; Shimoke, Komatsu & Matsui,
1987), presumably by the ion exchange of F~ for OH" in hydroxyapatite.
Thisfluorideuptake has the effect of improving the resistance of the tooth
material to acid attack (Maldonado, Swartz & Phillips, 1978; Wesenberg
& Hals, 1980; Kidd, 1978). Maldonado, Swartz & Phillips (1978) have
found that the solubility of enamel in acid was reduced by 53 % when in
contact with a glass polyalkenoate cement. Also, fluoride adsorption
reduces surface energy (Glanz, 1969) making adhesion of caries-promoting
plaque more difficult (Rolla, 1977). It also decreases demineralization and
increases remineralization of teeth (Wei, 1985) and reduces the fer-
mentation of carbohydrates and the growth of plaque bacteria (Hamilton,
1977; Tanzer, 1989). Of course, it has been known since the early 1940s
that fluoride inhibits dental decay (Horowitz, 1973).

Acid erosion and clinical durability


The glass polyalkenoate cement is the most durable of all dental cements.
In a very early study, Kent, Lewis & Wilson (1973) found that the surface
of the glass polyalkenoate cement was much less affected by acids and
stained much less than traditional dental silicate cement. This early
observation has since been substantiated by both in vivo studies (Mitchem
& Gronas, 1978, 1981; Ibbetson, Setchell & Amy, 1985; Mesu & Reedijk,
1983; Pluim et al, 1984; Pluim, Arends & Havinga, 1985) and laboratory
tests (Mesu, 1982; Beech & Bandyopadhyay, 1983; Sidler & Strub, 1983;
Kuhn, Setchell & Teo, 1984; Theuniers, 1984; Pluim et al.9 1984; Pluim,
Arends & Havinga, 1985; Walls, McCabe & Murray, 1985; Wilson et ai,
1986a,b; Gulabivala, Setchell & Davies, 1987).

158
Glass polyalkenoate (glass-ionomer) cement
Clinical durability largely depends on resistance to acid erosion, for acid
conditions occur in stagnation regions of the mouth where dental plaque
accumulates. Plaque contains streptococci and lactobacilli which degrade
plaque polysaccharides and sucrose to lactic acid (Tanzer, 1989; Jenkins,
1965), and lactic acid is the most potent driving force in causing
demineralization of teeth. The lower extreme of acidity found in the mouth
is pH = 40 (Stephan, 1940; Kleinburg, 1961). Glass polyalkenoate
cements begin to erode only at this pH (Crisp, Lewis & Wilson, 1980);
Walls, McCabe & Murray (1988) and Wilson et al. (1986b) found that one
brand did not erode at all at this pH. Susceptibility to acid erosion is low
even when the pH is 2-7.
Crisp, Lewis & Wilson (1980) made a chemical study of the erosion of
a glass polyalkenoate cement under acid attack. They found that the chief
species eluted were sodium and fluoride ions and silicic acid suggesting that
attack occurred mainly on the glass particles rather than on the matrix.
Resistance to acid erosion depends on brand and varies from 0-04 to
0-54% per hour (Setchell, Teo & Kuhn, 1985; Wilson et al., 1986a; Walls,
McCabe & Murray, 1988). It would appear that cements based on
copolymers of acrylic and maleic acids are less durable than those based on
poly (aery lie acid). The extent of erosion varies inversely with the time
allowed for the cement to cure prior to exposure (Walls, McCabe &
Murray, 1988).
McKinney, Antonucci & Rupp (1987) found that the clinical wear of the
glass polyalkenoate cement compared favourably with that of the
composite resin, but they noted that it was prone to brittle fracture and
chemical erosion.
Clinical experience shows that these cements are durable. For example,
a failure rate as low as 2 % has been reported by Mount (1984) in a clinical
trial lasting seven years, and Wilson & McLean (1988) have cited a number
of clinical trials attesting to the durability of this cement.

5.9.11 Biocompatibility
Dental material
The biocompatibility of the glass polyalkenoate cement is good (Wilson &
McLean, 1988; Nicholson, Braybrook & Wasson, 1991) and its capacity to
release fluoride in a sustained fashion makes it cariostatic (Hicks, Flaitz &
Silverstone, 1986; Kidd, 1978). Its ability to provide an excellent seal
(Section 5.9.9) is an important attribute because in recent years it has

159
Polyalkenoate cements
become generally accepted that pulpal inflammation is caused not so much
by chemical toxicity as by the percolation of harmful bacteria between
cavity wall and the restorative (Brannstrom & Nyborg, 1969; Paterson,
1976; Browne et al, 1983). The seepage of harmful bacteria beneath a
restoration can be the cause of secondary caries (Bergenholtz et al, 1982)
and is the cause of much of the failure of dental amalgams (Mjor, 1985).
Early studies showed that glass polyalkenoate cement has less of an
adverse effect on the dental pulp than has silicate cement (Kawahara,
Imanishi & Oshima, 1979; Pameijer, Segal & Richardson, 1981). Ap-
parently, the glass polyalkenoate cement causes only mild pulpal inflam-
mation which reaches a maximum 14 days after placement and then
progressively diminishes (Kawahara, Imanishi & Oshima, 1979). The effect
is greatly diminished by a layer of 0-5 mm of dentine. More recent
materials appear to cause less inflammation than earlier ones (Plant et ai,
1984; Yoshii et al.9 1987). A recent careful study of biocompatibility by
Pameijer & Stanley (1988) on primates, using a standard methodology
recommended by the American National Standards Institute and the
American Dental Association, showed a modern cement to be bland.
The glass-ionomer cement is well-tolerated by living cells (Kawahara,
Imanishi & Oshima, 1979; Kasten et al, 1989). An important distinction
is to be made between the freshly mixed and fully set glass-ionomer
cement. The glass-ionomer cement exhibits an antibacterial effect when
freshly mixed which diminishes with time (Tobias, Browne & Wilson,
1985). It exhibits some cytotoxicity when freshly mixed, but none when
fully set (Tyas, 1977; Kawahara, Imanishi & Oshima, 1979; Meryon &
Browne, 1984; Hume & Mount, 1988; Brook, Craig & Lamb, 1991a;
Hetem, Jowett & Ferguson, 1989). Both the antibacterial and cytotoxic
effects appear to be associated with the leachate from the cements. The
precise cause of these effects remains unclear. Some workers have suggested
that it is the combination of fluoride and low pH in the young cement
(Leirskar & Helgeland, 1987; McComb & Ericson, 1987), while others
implicate the release of metal ions and free poly(acrylic acid)s (Kawahara,
Imanishi & Oshima, 1979; Nakamura et al., 1983). Recent studies by
Brook, Craig & Lamb (1991a,b) on bone substitutes have shed some light
on this problem and are described later.
No problems arise when the glass-ionomer cement is used to restore
abrasion/erosion lesions in primary teeth and as a lining material in
shallow cavity preparations (Tobias et ah, 1978, 1987). In deeper

160
Glass polyalkenoate {glass-ionomer) cement
preparations, when the dentine layer is very thin it is advisable to use an
antibacterial calcium hydroxide liner.
Post-operative sensitivity has occasionally been reported when the
glass-ionomer cement has been used as a luting agent. This observation is
more than anecdotal, but the reason for it is unknown. It is not connected
with pulpal irritation but may be related to hydraulic pressures (Pameijer
& Stanley, 1988). The indication is that sensitivity is related to clinical
technique and is exacerbated if certain slow-setting glass-ionomer cements
are used, especially if they are mixed too thinly.

Bone cement and bone substitute


Not surprisingly, this cement has been studied as a possible replacement
for the poly(methyl methacrylate), PMMA, bone cement, a material which
is not entirely satisfactory (Jonck, Grobbelaar & Strating, 1989a,b; Jonck
& Grobbelaar, 1990). Its use has been linked with the pathogenesis of
failure; there is tissue sensitivity and it is not suitable in reconstructive
surgery following cancer or replacement procedures following radiation
treatment of diseased bone.
Jonck, Grobbelaar & Strating (1989a,b) and Jonck & Grobbelaar (1990)
have made biological evaluations, using baboons, of glass polyalkenoate
cement for use in joint replacement surgery. It was found to be stable
within the bone environment and there were no signs of surface dissolution
over a period of three years. It was also observed to bond to bone and act
as a sealant. The promotion of new bone formation and calcification on the
surface of the cement was observed. This reparative process was tentatively
attributed to the ion-exchange properties of the cement and, in particular,
to the release of fluoride. The glass polyalkenoate cement was found to
have no inhibitory effect on bone tissue development. Rather, it appeared
to promote the formation of osteoblasts (bone-forming cells) and the
normal differentiation of haemopoietic (blood-forming) tissue.
The glass-ionomer cement was found to be non-toxic. There were no
signs of inflammation or irritation with any of the glass polyalkenoate
cement implants even after several months. By contrast a proportion of
PMMA cement implants caused swelling and bone reactions. There were
also signs of possible hyperaemia (blood congestion) and infarcts (areas
deprived of blood supply) and dead tissue.
Thus, the glass polyalkenoate cement has distinct biological advantages
stemming from its dynamic surface chemistry, which is favourable to bone

161
Polyalkenoate cements
mineral precipitation. This process, in turn, promotes normal tissue
responses essential for the processes of bone formation.
Brook, Craig & Lamb (1991a,b), who studied the use of the cement as an
alveolar bone substitute, have confirmed thesefindings.They find that the
cement forms an intimate bioactive bond with bone cells and their work
has highlighted the complex role offluoride.Their in vitro studies showed
that of the cement pastes only the one derived from a non-fluoride-
containing glass (MP4) did not inhibit cell growth. But this cement did not
integrate with bone as effectively as did cements based on fluoride-
containing glasses.
Jonck, Grobbelaar & Strating (1989a,b) suggested that the slow release
of fluoride had a beneficial in vivo effect on osteogenesis (the formation of
bone) similar to that of osteoid formation stimulated byfluoridetherapy in
the treatment of osteoporosis (abnormal porosity of the bone) (Frost,
1981). There appeared to be an optimum level of fluoride for the
stimulation of bone-forming cells (Turner et al, 1989) since mild toxic
effects are encountered in the closed in vitro situation (Kawahara, Imanishi
& Oshima, 1979; Nakamura et ai, 1983; Hetem, Jowett & Ferguson, 1989;
Kasten et al., 1989; Jonck, Grobbelaar & Strating, 1989a,b; Brook, Craig
& Lamb, 1991a,b). Brook, Craig & Lamb (1991b) have suggested that in
the dynamic in vivo situation the leaching of fluoride may stimulate
integration with the bone (osseointegration) thus accounting for the
superiority of cements based onfluoride-containingglasses in this respect.
Brook, Craig & Lamb (1991a) found that set glass-ionomer cement
implants compared favourably with those of other materials. More bone
was formed around glass-ionomer cement implants than those of
hydroxyapatite. Direct bonding of glass polyalkenoate cement to mineral-
ized collagenous extracellular bone matrix was found without an inter-
vening layer. Behaviour was similar to that of bioglass and glass ceramics.
To sum up, glass-ionomer cement forms an intimate bond to living bone,
a process which is enhanced by the release of fluoride.

5.9.12 Modified and improved materials


Various attempts have been made to improve the glass polyalkenoate
cement. We have already described (Section 5.9.5) the most important
innovation, the use of (+ )-tartaric acid to improve setting characteristics
(Wilson & Crisp, 1976). Before its use was discovered, only one glass,

162
Glass polyalkenoate (glass-ionomer) cement
G-200, could be used to form practical cements; afterwards a whole range
of glasses became available, including clear ones. The original glass poly-
alkenoate cement based on G-200 had poor translucency because this high-
fluoride glass was heavily opal. The use of the new clear glasses yielded a
new generation of cements with good translucency. Radio-opacity has also
been introduced into glasses by replacing calcium with strontium or
lanthanum (Section 5.9.2).
All additives, ( + )-tartaric acid and metal fluorides, which improve
setting also increase cement strength (Section 5.9.4). The same point can be
made about the technique of acid-washing glasses (Schmidt et al., 1981a)
which, by removing calcium ions from the surface of glass particles,
enhances the mixing and setting qualities of cements. All of which
illustrates the general point that strength is related to the working and
setting qualities of cements.
The realization that cement strength could be increased by increasing the
molecular weight of the poly(alkenoic acid) was important (Section 5.9.3).
Unfortunately, this also increased the stiffness of the cement mix,
necessitating a reduction in powder/liquid ratio. Thus, the benefits gained
on the one hand were lost on the other. Fortunately, there was a way round
this problem. By drying the poly(alkenoic acid) and blending it with the
glass powder cement, water could be used to initiate cement formation,
with the benefit of lowering the stiffness of the mix (Prosser et al., 1984;
McLean, Wilson & Prosser, 1984). This technique permitted the use of
poly(alkenoic acid) of higher molecular weight.
A combination of all these technical improvements has led to a
significant improvement in the quality of the glass polyalkenoate cement.

Reinforced glass polyalkenoate cements


More direct attempts at improving cement strength have been made.
Wilson et al (1980) and later Prosser, Powis & Wilson (1986) observed that
disperse-phase glasses yielded stronger cements. The strongest cement
yielded by a clear glass had aflexuralstrength of 21 MPa, whereas a glass
containing corundum and fluorite as disperse phase gave a cement with a
flexural strength of 33 MPa.
The use of reinforcingfillerswas examined by Seed & Wilson (1980). An
alumina-fibre cement had a flexural strength of 44 MPa, while one
reinforced by carbon fibre had a flexural strength of 53 MPa. Metal
reinforcement has also been examined. Seed & Wilson (1980) found that a
cement reinforced with silver-tin alloy had aflexuralstrength of 40 MPa.

163
Polyalkenoate cements
However, restorations made from this material could not be polished and
were aesthetically very poor. Simmonds (1983) has pursued this idea, and
a material has been placed on the market. But according to Moore, Swartz
& Phillips (1985) such cements have less resistance to abrasion than a
simple glass polyalkenoate cement.
Recently, Oldfield & Ellis (1991) have examined the reinforcement of
glass-ionomer cement with alumina (Safil) and carbon fibres. The
introduction of only small amounts of carbon fibres (5% to 7*5% by
volume) into cements based on MP4 and G-338 glasses was found to
increase considerably both the elastic modulus andflexuralstrength. There
was an increase in work of fracture attributable to fibre pull-out. A
modulus as high as 12-5 GPa has been attained with the addition of 12%
by volume of fibre into MP4 glass (Bailey et ai, 1991). Results using
alumina fibre were less promising as there was nofibrepull-out because of
the brittle nature of alumina fibres which fractured under load.
McLean & Gasser (1985) had the idea of fusing silver with ionomer glass
to produce a fused metal-glass (cermet) powder which replaced the glass
powder in the conventional cement. The idea was to improve strength,
toughness and abrasion resistance. Resistance to abrasion is improved
(Moore, Swartz & Phillips, 1985; Swift, 1988b), probably by a lubricating
effect since the surface of the restoration can take a polish (McLean &
Gasser, 1985; McKinney, Antonucci & Rupp, 1985, 1987). However,
Lloyd & Adamson (1987) found that a cermet polyalkenoate cement was
no stronger and no tougher than a conventional material. In an in vitro
study, Thornton, Retief & Bradley (1986) showed that the bond strength to
enamel and dentine was lower, and consequently microleakage has been
found to be higher (Robbins & Cooley, 1988). Fluoride release was found
to be lower than that of conventional materials (Thornton, Retief &
Bradley, 1986). The material is radio-opaque, which is an advantage, for its
presence in living tissues can then be detected by X-rays. Obviously, it
cannot be used for aesthetic restorations andfindschief use in preparing a
core to receive porcelain and gold crowns. One example of this material is
available on the market.
Very recently, Williams, Billington & Pearson (1992) have examined the
effect of reinforcement by silver or silver-tin alloy on the mechanical
properties of three glass-ionomer cements. Measurements of compressive,
flexural, tensile (measured by the diametral compressive procedure) and
shell strength are given in Table 5.17. These results show that the effect of
reinforcement varies from cement to cement but, in general, increases it.

164
Glass polyalkenoate (glass-ionomer) cement

Table 5.17. Effect of reinforcement on various strength properties of three


glass polyalkenoate cements (Williams, Billing ton & Pearson, 1992)

Cement Standard Reinforced


Powder: liquid ratio, g cm" A 3-8 3-2
B 2-4 51
C 7-0 110
Wet compressive strength (24 h), MPa A 146 137
B 136 138
C 162 221
Wet flexural strength (24 h), MPa A 21-3 290
B 24-7 36-3
C 251 670
Wet tensile strength, MPa A 18-6 12-8
B 13-3 14-3
C 16-2 251
Wet shell strength (24 h), MPa A 35-7 44-6
B 32-3 39-6
C 46-9 75-8

The effect on cement C is particularly dramatic and the flexural strength of


cement C is exceptionally high. In part, this is to be attributed to the high
powder/liquid ratio. These results are to be compared with the flexural
strengths of early polyalkenoate cements which were c. 10 MPa.

Aluminoborate glasses
Brief mention may be made of the aluminoborate glasses developed by
Bertenshaw et al. (1979). These glasses are prepared by fusing a mixture of
boric oxide (replacing silica), alumina and a metal oxide, usually zinc
oxide. The fusion temperature is much lower than for the aluminosilicate
glasses. The compositional region for glass formation is restricted and
glasses are only obtained when the alumina content lies between 1 and
13 mole. Boric oxide content ranges from 35 to 73 mole boric acid and
metal oxide content from 27 to 59 mole.
The cement-forming reaction will be similar to glass polyalkenoate
cement. The cement matrix will consist of metal polyacrylates, but boric
acid will be produced instead of silica gel. Since boric acid has a water
solubility of 2-7 % compared with the near insolubility of silica gel, it would

165
Polyalkenoate cements
be expected that these cements would be less durable than the conventional
glass polyalkenoate cements.
Compressive strengths of these cements were found by Bertenshaw et al.
(1979) to range from 20 to 50 MPa and tensile strengths from 5 to 9 MPa.
These values are inferior to those of the conventional glass polyalkenoate
cements but similar to those of the zinc polycarboxylate cements. They are
reported to have a good translucency and have a low solubility in water.
These materials do not appear to be manufactured commercially.
Recently, Wilson & Combe (1991) have studied the reactivity of
magnesium, zinc, calcium and strontium boroaluminate glasses towards
poly (aery lie acid) solutions. The controlling factor would seem to be the
alumina content of these glasses which serves to moderate the setting rate
of the cements.

5.9.13 Applications
The glass polyalkenoate cement is a versatile material and finds use in
dentistry and more generally as a biomaterial. There have also been
applications outside these fields.

Dental
The glass-ionomer cement is the most versatile of all the dental cements
and has been developed for a variety of applications (McLean & Wilson,
1974, 1977a,b,c; Swift, 1988b; van de Voorde, 1988; Wilson & McLean,
1988; Mount, 1990). Many of its applications depend on its adhesive
quality which means that, unlike the non-adhesive traditional filling
materials, it does not require the preparation of mechanical undercuts for
retention and the consequent loss of sound tooth material.
The glass polyalkenoate cement was originally intended as a substitute
for dental silicate cements for the aesthetic restoration of front (anterior)
teeth (Wilson & Kent, 1972; Knibbs, Plant & Pearson, 1986a; Osborne &
Berry, 1986; Wilson & McLean, 1988). It is suitable for restoring anterior
cavities in low-stress situations, that is when the restoration is completely
supported by surrounding tooth material. These cavities occur on the
adjacent surfaces of neighbouring teeth (class III cavities) and at the gum
line (class V cavities).
Quite early, McLean & Wilson (1977b) found that the glass poly-
alkenoate cement was particularly effective for restoring cervical lesions -

166
Glass polyalkenoate {glass-ionomer) cement

small cavities that occur at the gum line. These are often found in middle-
aged people and are caused by abrasion or erosion rather than by dental
decay (caries). These lesions are so small that it is unwise to enlarge them
to provide mechanical retention (Wilson & McLean, 1988). For this reason
the adhesive glass polyalkenoate cement is the material of choice for the
restoration of cervical lesions (Tyas & Beech, 1985).
The cement is commonly used to restore primary (children's) teeth since
the trauma of drilling may be minimized or avoided altogether (Wilson &
McLean, 1988; Walls, Murray & McCabe, 1988).
Some recent ingenious applications are of particular interest as they
exploit the adhesive properties of the glass-ionomer cement to the full.
Clinicians have now developed the concept of minimal cavity preparation
(McLean, 1980, 1986; Hunt, 1984; Knight, 1984; Wilson & McLean,
1988). The idea is that since caries is mainly a disease of the dentine, then
only a minimal amount of enamel need be removed, just sufficient to allow
for the excavation of the carious dentine. A small channel is drilled through
the enamel and the carious dentine is removed through it. Sound enamel is
thus preserved. The excavated region is thenfilledwith glass polyalkenoate
cement, which by virtue of its adhesive nature holds the enamel shell
together.
Another use of the glass polyalkenoate cement is as a base material
which can be placed under dental amalgam or composite resin in the
restoration of posterior (molar and semi-molar) teeth (Smith, Ruse &
Zuccolin, 1988; Wilson & McLean, 1988). Its role is to adhere to dentine,
provide a protective seal against bacteria and release fluoride: functions
which prevent caries occurring under the restoration. In the laminate
restoration, fully described in Section 5.9.9, it is used, in effect, as a dentine
substitute. Base cements used under other restoratives are frequently made
radio-opaque so that they can be distinguished from carious dentine. This
is achieved by adding zinc oxide, using silver-glass cermets in place of the
glass or using glasses in which the calcium is replaced by lanthanum or
strontium. Base cements are generally quick-setting.
The glass polyalkenoate cement is also used in the fitting of crowns. It is
used to build up a substructure, known as a core, if there is insufficient
tooth material to take a crown (Wilson & McLean, 1988). Core build-up
materials are generally made radio-opaque and the silver cermet is often
used. A fine-grained version was developed to be used as a luting agent for
the cementation of cement crowns and veneers; this is based on a fine-
grained glass (Wilson et aL, 1977; Wilson & McLean, 1988).

167
Polyalkenoate cements
Another important use for the glass polyalkenoate cement is in
preventive dentistry where it can be used to fill and seal naturally occurring
pits and fissures in molar teeth which are sites for the initiation of caries
(McLean & Wilson, 1974, 1977b; Komatsu, 1981; Wilson & McLean,
1988). Its adhesive quality and ability to act as a long-term fluoride-
releasing gel make it particularly suitable for this purpose. Special
formulations for this application have been placed on the market.

Splint bandage
The glass polyalkenoate cement forms the basis of a novel splint bandage
that was developed in the early 1970s (Parker, 1974; Potter et ai, 1977,
1979; Hall, 1977) and marketed by Smith & Nephew. In this application,
a powder blend of glass and poly(acrylic acid) is applied to a bandage.
When required for use the bandage is immersed in water, a technique
identical to that used for the conventional plaster bandage. The addition of
water activates the cement-forming reaction and the bandage sets hard,
but remains more flexible than the plaster bandage. It has other
advantages. It is stronger and attains strength more rapidly than the
conventional bandage. It is also impervious to water once set. These
represent significant advantages for the patient. It can be applied over a
normal plaster bandage to protect the latter from the softening effect of
water.

Bone cement
Existing bone cements for orthopaedic use are based on a modified
poly(methyl methacrylate) resin. It has disadvantages. The formation of
this polymer in situ is accompanied by the marked evolution of heat which
can damage tissues (Feith, 1975). The presence of unreacted monomer,
which can leach out, also damages tissues (Petty 1980; Pople & Phillips,
1988). The glass polyalkenoate cement has an exceptionally low exotherm
and good biocompatibility, which together with its ability to bond to bone
give it potential as an improved bone cement.
A biological evaluation of the glass polyalkenoate cement as a bone
cement has been carried out by Jonck, Grobbelaar & Strating (1989a,b;
Jonck & Grobbelaar, 1990) on baboons and is reported in Section 5.9.11;
initial findings are promising.

168
Resin glass polyalkenoate cements

Alveolar bone substitute


In the UK about 20 % of the population over 16 are without natural teeth
(Brook, Craig & Lamb, 1991b). Dentures are supported by the alveolar
ridge which over the years is subjected to progressive resorption. The
reduction of the alveolar ridge gives rise to functional problems. Brook,
Craig & Lamb (1991b) have used the glass-ionomer cement successfully to
restore this ridge showing that it successfully integrates with bone (see
Section 5.9.11).

Slip casting mould


The glass polyalkenoate cement can also be used to replace plaster as a
mould in the slip process for pottery. It possesses the same property as
plaster of Paris, of causing material to deposit on its surface from slip
suspensions. So far this property has not been exploited in the manufacture
of pottery.

5.10 Resin glass polyalkenoate cements


5.10.1 General
One of the most interesting recent developments has been the advent of the
resin glass polyalkenoate cements (Antonucci, McKinney & Stansbury,
1988; Mitra, 1989; Wilson, 1989, 1990; Mathis & Ferracane, 1989;
Minnesota Mining & Manufacturing Company, 1989; Albers, 1990). They
are dual-cure hybrids that set by a combination of acid-base and
polymerization reactions, and there are several types. Polymerization is
effected by either chemical or light initiation.
At its most basic, the resin glass polyalkenoate cement can be seen as a
glass polyalkenoate cement in which the water component is replaced by a
water-HEMA mixture. HEMA is hydroxymethyl methacrylate, its
hydroxy group making it water-soluble:

CH2 = C — C O — O — CH2-CH2-OH

CH3
Also, included in these formulations are water-soluble initiators/activators
for the polymerization of HEMA.
The resin glass polyalkenoate cements are mixed in the same way as
conventional materials. In the case of the light-activated systems they

169
Polyalkenoate cements

remain workable for 10 or more minutes unless exposed to light. When


light is shone on them they are activated and set hard in 30 seconds. There
is a dual setting reaction: the normal glass polyalkenoate cement acid-base
reaction and, additionally, a free-radical or photochemical polymerization
process, similar to that occurring in composite resins. These may be
represented as:

(1) Acid—base reaction'.

Calcium aluminosilicate Polyacrylic Calcium and aluminium


glass (base) acid polysalt hydrogel

(2) Polymerization reaction:

HEM A + photochemical initiator/activator ^PolyHEMA matrix

Two matrices are formed: a metal polyacrylate salt and a polymer. There
is a lack of water in the system because some of it has been replaced by
HEM A, and lack of water in glass polyalkenoate cements is known to slow
down the ionomer acid-base reaction (Hornsby, 1977). Thus, the initial set
of these materials results from the polymerization of HEM A and not the
characteristic acid-base reaction of glass-ionomer cements. The later
reaction serves only to harden and strengthen the already formed matrix.
The two matrix-forming reactions are shown in more detail in Figure
5.23.

5.10.2 Class I hybrids


In more complex forms of this resin hybrid, other dimethacrylates may be
present, such as the ethylene glycol dimethacrylates, and bis-GMA, when
HEMA acts as a co-solvent for water and bis-GMA (Antonucci,
McKinney & Stansbury, 1988). The general composition of these materials,
which we term class I hybrids, is summarized below:

Powder component'.
Glass (Chemfil II) + poly(acrylic acid) + tartaric acid
Liquid component {replaces water):

Water/HEMA
Other difunctional hydroxy dimethacrylates, e.g. the ethylene
glycol dimethacrylates

170
Resin glass polyalkenoate cements
Bis-GMA
Initiator/activator
In chemically-cured materials, one example of an initiator/activator
system is: hydrogen peroxide as initiator, ascorbic acid as activator and
cupric sulphate as co-activator. In light-cured materials, camphorquinone
is used as a visible-light photochemical initiator, sodium /?-toluene-
sulphinate as activator and ethyl 4-dimethylaminobenzoate as photo-
accelerator.
If there is too little water in a composition then the acid-base reaction
DUAL CURE 1
ACID — BASE REACTION

I I I
CH2 CH 2 F CH 2

CH-COOH CH-COO — A l 3 + - = 0 0 C — CH
' Calcium
I 2 aluminosilicate CH-, CH,
I •
CH-COOH CH-COO"— C a 2 + — 00C—CH
I I
Poly (acrylic acid) Ca, Al polysalt hydrogel

DUAL CURE 2

HEMA
C=CH2 POLYMERIZATION -C-CH2-

C=0 c=o
Photo or chemical
0 initiator / activator 0

I CH 2
HEMA poly HEMA
CH 2 CH 2

CH 2

I
OH

Figure 5.23 The two matrix-forming reactions in class I resin-based glass polyalkenoate
cements.

171
Polyalkenoate cements

will be completely inhibited and only the polymerization reaction will take
place, in which case the material is not strictly speaking a glass
polyalkenoate cement.

5.10.3 Class II hybrids


The two matrices in these cements are of a different nature: an ionomer salt
hydrogel and polyHEMA. For thermodynamic reasons, they do not
interpenetrate but phase-separate as they are formed. In order to prevent
phase separation, another version of resin glass polyalkenoate cement has
been formulated by Mitra (1989). This is marketed as VitraBond, which we
term a class II material. In these materials poly(acrylic acid), PAA, is
replaced by modified PAAs. In these modified PAAs a small fraction of the
pendant - C O O H groups are converted to unsaturated groups by con-
densation reaction with a methacrylate containing a reactive terminal
group. These methacrylates can be represented by the formula:

T—R—C=CH2

CH3
where T is a terminal group, for example:
HO— H 2 N— OCN— CH2 — CH —

The condensation reaction can be represented thus:

— COOH + H O — R — C = C H 2 • —CO—O—R—C=CH2

CH3 CH3

Modified PAAs can be represented by the generic formula:

—CH 2 —CH—CH 2 —CH—CH 2 —CH — CH 2 —CH—CH

CO CO CO CO

OH OH O OH
R—C=CH2

CH3
The liquids used for class II hybrids contain 2 5 ^ 5 % modified PAA and
21-41 % HEM A. The initiator system for light activation contains

172
Resin glass polyalkenoate cements
camphorquinone and diphenyliodonium chloride (Mitra, 1989). The glass
powder has the following percentage composition:
SiO2 27-24; A12O3 0-81; P2O50-95; NH4F3-37; A1F3 20-97;
Na3AlF6 10-81; ZnO 20-97; MgO215; SrO 12-74.
Formation of matrices
When a resin glass polyalkenoate cement, containing a modified PAA and
HEMA, is mixed, a paste is formed which sets only slowly in the absence
of light. When activated by light the paste sets in 30 s. Several types of
polymerization can then take place. Both HEMA and the modified PAA,
because it contains unsaturated groups, will polymerize. PolyHEMA and
a crosslinked PAA of high molecular weight will be formed. In addition,
the modified PAA may copolymerize with HEMA; thus, polyHEMA is
chemically linked to the polyacrylate matrix and so cannot phase-separate.
The matrix of such a cement contains both ionic and covalent crosslinks
(Figure 5.24). Thus, the cement matrix is reminiscent of an ion-exchange
resin.

5.10.4 Properties
These resin-modified glass polyalkenoate cements have both advantages
and disadvantages over conventional glass polyalkenoate cements. How-
ever, because of their poor translucency they are recommended only as
liners or bases.
They have improved setting characteristics. They have a long working
time because HEMA slows the acid-base reaction, yet set sharply once the
polymerization reaction is initiated by light. They are also resistant to early
contamination by water because of the formation of an organic matrix,
and so do not require protection by varnish. This combination of properties
is bound to appeal to the clinician.
The freshly set class II (Vitrabond) resin glass polyalkenoate cement
appears to have rubbery characteristics and there is some debate as to

CH2 CH 2 CH 2 CH2
2+
CH-COO" M "OOC-CH CH-COO-R-CH-CH2-CH2-CH-R-OOC-CH
CH 2 CH 2 CH 2 CH 3 CH 3 CH 2

I II I
Figure 5.24 The matrix of a class II resin-based glass polyalkenoate cement, showing ionic
and covalent crosslinks.
173
Polyalkenoate cements
Table 5.18. Strength of resin glass polyalkenoate cements (Antonucci,
McKinney & Stansbury, 1988; Wilson & McLean, 1988; Mathis &
Ferracane, 1989; Mitra, 1989; Minnesota Mining & Manufacturing
Company, 1989; Alters, 1990)

Resin Conventional
Wet strength,
MPa (24 hours) Class I Class II Liner/base Filling

Compressive 94-139 53-96 56-79 140-195


Flexural — 25-5 5-2-10-3 8-9-30-3
Tensile 16-4^33-9 11-2-17-4 3-4-9-1 9-0-19-3
Adhesion 1-6 9-8-11-3 3-4^3-9 1-7-6-8
(dentine)

whether this is advantageous or not. The reason for this rubberiness is that
the polymer is only lightly crosslinked, and at set the acid-base reaction
has not proceeded very far. Most probably, these rubbery characteristics
will disappear as the cement ages and the acid-base reaction is completed.
But this may take a very long time.
Some data have been published on the mechanical properties of these
cements (Antonucci, McKinney & Stansbury, 1988; Mathis & Ferracane,
1989; Mitra, 1989; Minnesota Mining & Manufacturing Company, 1989;
Albers, 1990), but much of it comes from patents and company reports so
it would be unwise to draw firm conclusions from thesefiguresalone.
Although both class I and II resin hybrids are stronger than conventional
liners and bases, class II materials are not as strong as conventional filling
materials (Table 5.18). According to Mathis & Ferracane (1989) a class I
material developed 82 % of its 24-hour ultimate compressive strength in
one hour, which compares favourably with a figure of about 52 % for a
conventional glass polyalkenoate cement. Rapid development of strength
is to be expected because of the polymerization process.
Both class I and class II resin glass polyalkenoate cements are claimed to
bond to dentine. This can be accepted. But that the bond is stronger and
develops more rapidly than that of the conventional glass polyalkenoate
cement, as is claimed for class II materials (Minnesota Mining &
Manufacturing Company, 1989) requires confirmation.
It may be, because of the slowness of the acid-base reaction in resin glass
polyalkenoate cements, that free poly(acrylic acid) is available for a
longer period than in conventional glass polyalkenoate cements, for the

174
References
formation of a stronger adhesive bond. However, it seems doubtful that
the adhesive bond will be developed more rapidly than in the conventional
glass polyalkenoate for, according to van Zeghbroeck (1989), the adhesive
bond of a conventional glass polyalkenoate cement develops its maximum
strength rapidly (within five minutes). The resin glass polyalkenoate
cement has the undoubted advantage of bonding directly to composite
resins and this makes it ideal for use in the glass polyalkenoate
cement/composite resin laminates.
There is bound to be one problem with resin glass polyalkenoate cement.
Because the matrix is a mixture of hydrogel salt and polymer, light-
scattering is bound to be greater than in the conventional material.
Moreover, the zinc oxide-containing glass of class II materials is bound to
be opaque. This makes it difficult to formulate a translucent material and
is the reason why their use is restricted to that of a liner or base. However,
the class II material cited will be radio-opaque because it uses strontium
and zinc, rather than calcium, in the glass.
A fundamental criticism of the resin-modified glass polyalkenoate
cements is that, to some extent, they go against the philosophy of the glass
polyalkenoate cement: namely, that the freshly mixed material should
contain no monomer. Monomers are toxic, and HEMA is no exception.
This disadvantage of composite resins is avoided in the glass polyalkenoate
cement as the polyacid is pre-polymerized during manufacture, but the
same cannot be said of these new materials. For this reason they may lack
the biocompatibility of conventional glass polyalkenoate cements. These
materials also absorb excessive amounts of water because of the hydro-
philic nature of polyHEMA (Nicholson, Anstice & McLean, 1992).
It is far too soon to make a judgement on these materials, which are of
considerable interest and only in the very early days of their development.
Most probably, much development will take place in this area.

References
Abramovich, A., Kaluza, J. J., Macchi, R. L. & Ribas, L. (1977). Enamel
surface treated with zinc polyacrylate dentine cements. Journal of Dental
Research, 56, 471-3.
Akahane, S., Tosaki, S. & Hirota, K. (1988). Fluoroaluminosilicate glass
powder for dental ionomeric cements. German Patent DE 3,804,469.
Akitt, J. W., Greenwood, N. N. & Lester, G. D. (1971). Nuclear magnetic
resonance and Raman studies of aluminium complexes formed in aqueous
solutions of aluminium salts containing phosphoric acids andfluorideions.
Journal of the Chemical Society, A, 2450-7.

175
Polyalkenoate cements

Albers, H. F. (ed.) (1990). Light-cured fluoride releasing liners. The Adept


Report. 1990; 1, No. 1, l^k
Andersson, K. R., Dent Glasser, L. S. & Smith, D. (1982). Polymerization and
colloid formation in silicate solutions. In Falcone, J. G. (ed.) Soluble Silicates.
ACS Symposium Series No. 194, Chapter 8. Washington, DC: American
Chemical Society.
Antonucci, J. M., McKinney, J. E. & Stansbury, J. W. (1988). Resin-modified
glass-ionomer cement. US Patent Application 160,856.
Anzai, M., Hirose, H., Kikuchi, H., Goto, J., Azuma, F. & Higasaki, S. (1977).
Studies on soluble elements and solubility of dental cements. I. Solubilities of
zinc phosphate cement, carboxylate cement and silicate cement in distilled
water. Journal of the Nihon University School of Dentistry, 19, 26-39.
Aveston, J. (1965). The hydrolysis of the aluminium ion: ultracentrifugal and
acidity measurements. Journal of the Chemical Society, 4438-43.
Bailey, J. E., Ellis, B., Howarth, L. G. & Oldfield, C. W. B. (1991). The fracture
of glass-ionomer cements. Conference on the Fractography of Glasses and
Ceramics II, New York State College of Ceramics, Alfred University, New
York, July 1990. American Ceramic Society.
Barnes, D. S. & Turner, E. P. (1971). Initial response of the human pulp to zinc
polycarboxylate cement. Journal of the Canadian Dental Association, 37,265-6.
Barry, T. I., Clinton, D. J. & Wilson, A. D. (1979). The structure of a
glass-ionomer cement and its relationship to the setting process. Journal of
Dental Research, 58, 1072-9.
Barton, J. R., Brauer, G. M., Antonucci, J. M. & Raney, M. J. (1975).
Reinforced polycarboxylate cements. Journal of Dental Research, 54, 310-23.
Baumann, E. & Gerhard, V. (1970). Vormischung, Verfahren zu ihrer
Herstellung und Verwendung. German Patent 1,903,807.
Beagrie, G. S., Main, J. H. P. & Smith, D. C. (1972). Inflammatory reaction
evoked by zinc polyacrylate and zinc eugenate cements: a comparison. British
Dental Journal, 132, 351-7.
Beagrie, G. S., Main, J. H. P., Smith, D. C. & Walshaw, P. R. (1974).
Polycarboxylate cements as a pulp capping agent. Journal of the Canadian
Dental Association, 40, 378-83.
Beech, D. R. (1972). A spectroscopic study of the interaction between human
tooth enamel and polyacrylic acid (polycarboxylate cement). Archives of Oral
Biology, 17, 907-11.
Beech, D. R. (1973). Improvement in the adhesion of polycarboxylate cements
to human dentine. British Dental Journal, 135, 442-5.
Beech, D. R. & Bandyopadhyay, S. (1983). A new laboratory method for
evaluating the relative solubility and erosion of dental cements. Journal of
Oral Rehabilitation, 10, 57-63.
Beech, D. R., Solomon, A. & Bernier, R. (1985). Bond strength of
polycarboxylic acid cements to treated dentine. Dental Materials, 1, 154-7.
Bergenholtz, G., Cox, C. F., Loesche, W. J. & Syed, S. A. (1982). Bacterial
leakage around dental restorations: its effect on dental pulp. Journal of Oral
Pathology, 11, 439-50.

176
References

Bertenshaw, B. W. & Combe, E. C. (1972a). Studies on polycarboxylates and


related cements. I. Analysis of cement liquids. Journal of Dental Research, 1,
13-16.
Bertenshaw, B. W. & Combe, E. C. (1972b). Studies on polycarboxylates and
related cements. II. Analysis of cement powders. Journal of Dental Research,
1, 65-8.
Bertenshaw, B. W. & Combe, E. C. (1976). Studies on polycarboxylates and
related cements. III. Molecular weight determinations. Journal of Dental
Research, 4, 87-90.
Bertenshaw, B. W., Combe, E. C. & Grant, A. A. (1979). Studies on
polycarboxylates and related cements. 4. Properties of cements. Journal of
Dentistry,!, 117-25.
Bertenshaw, B. W., Combe, E. C, Tidy, P. C. & Laycock, J. N. C. (1979).
Surgical cement composition containing aluminoborate glass and a polymer
glass and a polymer containing recurring carboxylic or carboxylate groups.
US Patent 4,174,334.
Bitner, T. J. & Weir, S. H. (1973). Fluoride uptake and acid solubility of enamel
exposed to carboxylate cement containing MFP. Journal of Dental Research,
52, 157-62.
Bitter, N. C. (1986). Glass ionomer-microfil technique for restoring cervical
lesions. Journal of Prosthetic Dentistry, 56, 661-2.
Bovis, S. C, Harrington, E. & Wilson, H. J. (1971). Setting characteristics of
composite filling materials. British Dental Journal, 131, 352-6.
Brannstrdm, M. & Nyborg, H. (1969). Comparison of pulpal effect of two
liners. Ada Odontologica Scandinavica, 27, 433-51.
Brook, I. M., Craig, G. T. & Lamb, D. J. (1991a). In vitro interaction between
primary bone organ cultures, glass-ionomer cements and
hydroxyapatite/tricalcium phosphate ceramics. Biomaterials, 12, 179-86.
Brook, I. M., Craig, G. T. & Lamb, D. J. (1991b). Initial in-vitro evaluation of
glass-ionomer cements for use as alveolar bone substitutes. Clinical Materials,
7, 295-300.
Brown, D. & Combe, E. C. (1973). Effects of stainless steelfilleron the
properties of polycarboxylate cement. Journal of Dental Research, 52, 388.
Browne, R. M., Tobias, R. S., Crombie, I. K. & Plant, G. C. (1983). Bacterial
microleakage and pulpal inflammation in experimental cavities. International
Endodontics Journal, 16, 147-55.
Brune, D. & Smith, D. (1982). Microstructure and strength properties of silicate
and glass-ionomer cements. Acta Odontologica Scandinavica, 40, 389-96.
Buonocore, M. G. (1955). A simple method of increasing the adhesion of acrylic
materials to enamel surfaces. Journal of Dental Research, 34, 849-53.
Buonocore, M. G. (1961). Test of an adhesive containing glycerophosphoric
acid dimethacrylate. In Phillips, R. W. & Ryge, G. (eds.) Adhesive Restorative
Dental Materials, pp. 172-6. Spencer, Indiana: Owen Litho Service.
Cahn, J. W. (1961). On spinodal decomposition. Acta Metallurgica, 9, 795-801.
Causton, B. E. (1981). The physio-mechanical consequences of exposing glass
ionomer cements to water during setting. Biomaterials, 2, 112-15.

177
Polyalkenoate cements

Causton, B. E. (1982). Primers and mineralizing solutions. In Smith, D. C. &


Williams, D. F. (eds.) Biocompatibility of Dental Materials. Volume II.
Biocompatibility of Preventive Dental Materials and Bonding Agents, Chapter
7. Boca Raton, Florida: CRC Press Inc.
Causton, B. E. & Johnson, N. W. (1982). Improvement of polycarboxylate
adhesion to dentine by the use of a new calcifying solution. British Dental
Journal, 152, 9-11.
Chamberlain, B. B. & Powers, J. N. (1976). Physical and mechanical properties
of three zinc polycarboxylate cements. Journal of the Michigan Dental
Association, 58, 494-500.
Connick, R. E. & Poulsen, R. E. (1957). Nuclear magnetic resonance studies of
aluminium fluoride complexes. Journal of the American Chemical Society, 79,
5153-7.
Cook, W. D. (1982). Dental polyelectrolyte cements. I. Chemistry of the early
stages of the setting reaction. Biomaterials, 3, 232-6.
Cook, W. D. (1983a). Dental polyelectrolyte cements. II. Effect of
powder/liquid ratio on their rheology. Biomaterials, 4, 21-4.
Cook, W. D. (1983b). Dental polyelectrolyte cements. III. Effect of additives on
their rheology. Biomaterials, 4, 85-8.
Cook, W. D. (1983c). Degradative analysis of glass-ionomer polyelectrolyte
cements. Journal of Biomedical Materials Research, 17, 1015-27.
Cornell, J. (1961). Adhesion to natural tooth. In Phillips, R. W. & Ryge, G.
(eds.) Adhesive Restorative Dental Materials, p. 159. Spencer, Indiana: Owen
Litho Service.
Cotton, F. A. & Wilkinson, G. (1966). Advanced Inorganic Chemistry. New
York: Wiley Interscience.
Cranfield, M., Kuhn, A. T. & Winter, G. (1982). Factors relating to the rate of
fluoride-ion release from glass-ionomer cement. Journal of Dentistry, 10,
333^1.
Crisp, S., Abel, G. & Wilson, A. D. (1979). The quantitative measurement of
the opacity of aesthetic dental materials. Journal of Dental Research, 58,
1585-96.
Crisp, S., Ferner, A. J., Lewis, B. G. & Wilson, A. D. (1975). Properties of
improved glass-ionomer cement formulations. Journal of Dentistry, 3, 125-30.
Crisp, S., Jennings, M. A. & Wilson, A. D. (1978). A study of temperature
changes occurring in setting dental cements. Journal of Oral Rehabilitation, 5,
139^4.
Crisp, S., Kent, B. E., Lewis, B. G., Ferner, A. J. & Wilson, A. D. (1980).
Glass-ionomer cement formulations. II. The synthesis of novel polycarboxylie
acids. Journal of Dental Research, 59, 1055-63.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1975). Gelation of polyacrylic acid
aqueous solutions and the measurement of viscosity. Journal of Dental
Research, 54, 1173-5.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1976a). Zinc polycarboxylate cements.
A chemical study of erosion and its relationship to molecular structure.
Journal of Dental Research, 55, 299-308.

178
References
Crisp, S., Lewis, B. G. & Wilson, A. D. (1976b). Characterization of
glass-ionomer cements. 1. Long term hardness and compressive strength.
Journal of Dentistry, 4, 162-6.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1976c). Characterization of
glass-ionomer cements. 2. Effect of powder: liquid ratio on the physical
properties. Journal of Dentistry, 4, 287-90.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1976d). Glass-ionomer cements:
chemistry of erosion. Journal of Dental Research, 55, 1032-41.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1977). Characterization of
glass-ionomer cements. 3. Effect of poly acid concentration on the physical
properties. Journal of Dentistry, 5, 51-6.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1979). Characterization of
glass-ionomer cements. 5. Effect of tartaric acid concentration in the liquid
component. Journal of Dentistry, 7, 304-5.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1980). Characterization of
glass-ionomer cements. 6. A study of erosion and water absorption in both
neutral and acidic media. Journal of Dentistry, 8, 68-74.
Crisp, S., Merson, S. A. & Wilson, A. D. (1980). Modification of ionomer
cements by the addition of simple metal salts. Industrial & Engineering
Chemistry, Product Research & Development, 19, 403-8.
Crisp, S., Merson, S., Wilson, A. D., Elliott, J. H. & Hornsby, P. R. (1979). The
formation and properties of mineral—poly acid cements. Part 1. Ortho- and
pyro-silicates. Journal of Materials Science, 14, 2941-58.
Crisp, S., Pringuer, M. A., Wardleworth, D. & Wilson, A. D. (1974). Reactions
in glass—ionomer cements. II. An infrared spectroscopic study. Journal of
Dental Research, 53, 1414-19.
Crisp, S., Prosser, H. J. & Wilson, A. D. (1976). An infra-red spectroscopic
study of cement formation between metal oxides and aqueous solutions of
poly(acrylic acid). Journal of Materials Science, 11, 36-48.
Crisp, S. & Wilson, A. D. (1974a). Reactions in glass ionomer cements. I.
Decomposition of the powder. Journal of Dental Research, 53, 1408-14.
Crisp, S. & Wilson, A. D. (1974b). Reactions in glass ionomer cements. III. The
precipitation reaction. Journal of Dental Research, 53, 1420-4.
Crisp, S. & Wilson, A. D. (1974c). Poly(carboxylate) cements. British Patent
1,484,454.
Crisp, S. & Wilson, A. D. (1976). Reactions in glass ionomer cements. V. Effect
of incorporating tartaric acid in the cement liquid. Journal of Dental Research,
55, 1023-31.
Crisp, S. & Wilson, A. D. (1977). Cements comprising acrylic and itaconic acid
copolymers and fluoroaluminosilicate glass powder. US Patent 4,016,124.
Crisp, S. & Wilson, A. D. (1978a). Poly(carboxylate) cements. British Patent
1,532,954.
Crisp, S. & Wilson, A. D. (1978b). Fluoroaluminosilicate glasses. British Patent
1,532,955.
Crisp, S. & Wilson, A. D. (1979). Cements. US Patent 4,143,018.
Crisp, S., Wilson, A. D., Elliott, J. H. & Hornsby, P. R. (1977). Cement-forming

179
Polyalkenoate cements

ability of silicate minerals and polyacid solution. Journal of Applied


Chemistry, 27, 369-74.
Cruikshank, M. C , Dent Glasser, L. S., Barri, S. A. I. & Pollett, J. F. (1986).
Penta-co-ordinated aluminium: a solid-state 27A1 N.M.R. study. Journal of
the Chemical Society, Chemical Communications, 23—40.
Eames, W. B., Hendrix, K. & Mohler, H. C. (1979). Pulpal responses in rhesus
monkeys to cementation agents and cleaners. Journal of the American Dental
Association, 98, 40-5.
Edwards, S. F. (1969). Theory of crosslinked polymerized material. Proceedings
of the Physical Society of London. Solid State Physics, 2 (2), 1-13.
Eick, J. D., Johnson, L. N., Fromer, J. R., Good, R. J. & Neuman, A. W.
(1972). Surface topography: its influence on wetting and adhesion in a dental
adhesive system. Journal of Dental Research, 51, 780-8.
El-Kafrawy, A. H., Dickey, D. M., Mitchell, D. F. & Phillips, R. W. (1974).
Pulp reaction to a polycarboxylate cement in monkeys. Journal of Dental
Research, S3, 15-19.
Elliott, J., Holliday, L. & Hornsby, P. R. (1975). Physical and mechanical
properties of glass-ionomer cement. British Polymer Journal, 7, 297-306.
Ellis, J., Jackson, A. M., Scott, R. P. & Wilson, A. D. (1990). Adhesion of
carboxylate cements to hydroxyapatite. III. Adsorption of poly(alkenoic
acid)s. Biomaterials, 11, 379-84.
Ellis, J. & Wilson, A. D. (1987). Report. Laboratory of the Government Chemist.
Ellis, J. & Wilson, A. D. (1990). Polyphosphonate cements: a new class of
dental materials. Journal of Materials Science Letters, 9, 1058-60.
Ellison, S. & Warrens, C. (1987). Solid-state nmr study of aluminosilicate
glasses and derived dental cements. Report of the Laboratory of the
Government Chemist.
ESPE. (1975). Mixing liquid for dental filling materials. British Patent 1,382,882.
Feith, R. (1975). Side effect of acrylic cement implanted into bone. Acta
Orthopaedica Scandinavica Supplementum, 161, 1-36.
Fields, J. E. & Nielsen, L. E. (1968). Dynamic mechanical properties of some
polymeric zinc salts. Journal of Applied Building Science, Yl, 1041-51.
Flood, H. & Forland, T. (1947a). The acidic and basic properties of oxides.
Acta Chemica Scandinavica, 1, 592-604.
Flood, H. & Forland, T. (1947b). The acidic and basic properties of oxides. II.
The thermal decomposition of pyrosulphates. Acta Chemica Scandinavica, 1,
781-9.
Flory, P. J. (1953). Principles of Polymer Chemistry, Chapter 11. Ithaca, New
York: Cornell University Press.
Flory, P. J. (1974). Introductory lecture. In Gels and Gelling Processes. Faraday
Discussions of the Chemical Society, No. 57, pp. 7-18.
Forsten, L. (1977). Fluoride release from a glass ionomer cement. Scandinavian
Journal of Dental Research, 85, 503-4.
Foster, J. F. & Dovey, E. H. (1974). Surgical cements of improved compressive
strength containing stannous fluoride and poly aery lie acid. US Patent
3,856,737.

180
References
Foster, J. F. & Dovey, E. H. (1976). Improvements in and relating to surgical
cements. British Patent 1,423,133.
Frost, H. M. (1981). Coherence treatment of osteoporosis. Orthopedic Clinics of
North America, 12, 649-69.
de Gennes, P. (1979). Scaling Concepts in Polymer Physics. Ithaca: Cornell
University Press.
Glanz, P-O. (1969). On wettability and adhesiveness. Odontologisk Revy, 20,
Supplement 20.
Goldman, M. (1985). Fracture properties of composite and glass ionomer dental
restorative materials. Journal of Biomedical Materials Research, 19, 771-83.
Gourley, J. M. & Rose, D. E. (1972). Cavity bases under liners. Journal of the
Canadian Dental Association, 38, 246.
Granath, L. E. (1982). Pulp capping materials. In Smith, D. C. & Williams,
D. F. (eds.) Biocompatibility of Dental Materials. Volume II. Biocompatibility
of Preventive Dental Materials and Bonding Agents, Chapter 11. Boca Raton:
CRC Press Inc.
Gregor, H. P., Luttinger, L. B. & Loebl, E. M. (1955a). Metal polyelectrolyte
complexes. I. The polyacrylic acid-copper complex. Journal of Physical
Chemistry, 59, 34^9.
Gregor, H. P., Luttinger, L. B. & Loebl, E. M. (1955b). Metal polyelectrolyte
complexes. IV. Complexes of polyacrylic acid with magnesium, calcium,
manganese, cobalt and zinc. Journal of Physical Chemistry, 59, 990-1.
Gulabivala, K., Setchell, D. J. & Davies, E. H. (1987). An application of the jet
test method for the evaluation of disintegration of dental luting cements in
marginal gaps analogous to those of crowns and bridges. Clinical Materials,
2, 221-30.
Hagg, G. (1935). The vitreous state. Journal of Chemical Physics, 3, 42-9.
Hall, J. I. (1977). Orthopedic bandage. US Patent 4,044,761.
Hamilton, I. R. (1977). The effects of fluoride on enzymatic regulation of
bacterial carbohydrate metabolism. Caries Research, 11 (Supplement 1),
321-7.
Hatton, P. V. & Brook, I. M. (1992). Characterization of the ultrastructure of
glass-ionomer (glass polyalkenoate) cements. British Dental Journal, 173,
275-7.
Hazel, F. M., Shock, R. U. & Gordon, M. (1949). Interactions of ferric ions
with silicic acid. Journal of the American Chemical Society, 71, 2256-7.
Helgeland, K. (1982). In vitro testing of dental cements. In Smith, D. C. &
Williams, D. F. (eds.) Biocompatibility of Dental Materials. Volume II.
Biocompatibility of Preventive Dental Materials and Bonding Agents, Chapter
9. Boca Raton, Florida: CRC Press Inc.
Hembree, J. H. & Andrews, J. T. (1978). Microleakage of several class V
anterior restorative materials: a laboratory study. Journal of the American
Dental Association, 91, 179-83.
Hertet, R. S., Huget, E. F., de Simon, L. & Cosgrave, J. H. (1975). Short-term
stress relaxation behaviour of non-metallic restoratives. Journal of Dental
Research, 54, 1149-53.

181
Polyalkenoate cements
Hess, W. C. (1961). Organic components of enamel and dentine. In Phillips,
R. W. & Ryge, G. (eds.) Adhesive Restorative Dental Materials, pp. 10-14.
Spencer, Indiana: Owen Litho Service.
Hetem, S., Jowett, A. K. & Ferguson, M. W. (1989). Biocompatibility of a
posterior composite and dental cements using a new organ culture model.
Journal of Dentistry, 17, 155-61.
Hicks, M. J., Flaitz, C. M. & Silverstone, L. M. (1986). Secondary caries
formation in vitro around glass ionomer restorations. Quintessence
International, 17, 527-32.
Hill, R. G. & Wilson, A. D. (1988a). Some structural aspects of glasses used in
ionomer cements. Glass Technology, 29, 150-88.
Hill, R. G. & Wilson, A. D. (1988b). A rheological study of the role of additives
on the setting of glass ionomer cements. Journal of Dental Research, 68, 89-94.
Hill, R. G., Wilson, A. D. & Warrens, C. P. (1989). The influence of poly(acrylic
acid) molecular weight on the fracture toughness of glass-ionomer cements.
Journal of Materials Science, 24, 363-71.
Hinoura, K., Moore, B. K. & Phillips, R. W. (1986). Influence of dentin surface
treatments on the bond strength of dentin-lining cements. Operative Dentistry,
11, 147-54.
Hinoura, K., Moore, B. K. & Phillips, R. W. (1987). Bonding agent influence on
glass ionomer-composite resin. Journal of the American Dental Association,
114, 167-72.
Hodd, K. A. & Reader, A. L. (1976). The formation and hydrolytic stability of
metal ion-polyacid gels. British Polymer Journal, 8, 131-9.
Hopkins, R. P. (1955). Acrylate salts of divalent metals. Industrial & Engineering
Chemistry, 47, 2258-65.
Hornsby, P. R. (1977). A study of the formation and properties of ionic polymer
cements. Thesis for PhD, Brunei University, Middlesex, England.
Hornsby, P. R., Merson, S., Prosser, H. J. & Wilson, A. D. (1982). The
formation and properties of mineral-poly acid cements. Part 2. Chain, sheet
and three-dimensional silicates. Journal of Materials Science, 17, 3575-92.
Horowitz, H. S. (1973). A review of systematic and topicalfluoridesfor the
prevention of dental caries. Community Dentistry Oral Epidemiology, 1, 104-14.
Hotz, P., McLean, J. W., Seed, I. & Wilson, A. D. (1977). The bonding of
glass-ionomer cements to metal and tooth substrates. British Dental Journal,
142, 41-7.
Hume, W. R. & Mount, G. J. (1988). In vitro studies on the potential for pulpal
cytotoxicity of glass-ionomer cements. Journal of Dental Research, 67,
915-18.
Hunt, P. R. (1984). A modified class II cavity preparation for glass ionomer
restorative materials. Quintessence International, 15, 1011-18.
Ibbetson, R. J., Setchell, D. J. & Amy, D. J. (1985). An alternative method for
the clinical evaluation of the disintegration of dental cements. Journal of
Dental Research, 64, 672 Abstract No. 89.
Her, R. K. (1979). The polymerization of silica. The Chemistry of Silica,
Chapters 3 & 6. New York: Wiley-Inter science.

182
References
Isard, J. O. (1959). Electrical conduction in aluminosilicate glasses. Journal of
the Society of Glass Technology, 43, 113-23.
Jemt, T., Stalblad, P. A. & 0ilo, G. (1986). Adhesion of polycarboxylate-based
dental cements to enamel: an in vivo study. Journal of Dental Research, 65,
885-7.
Jendresen, M. D. & Trowbridge, H. D. (1972). Biological and physical
properties of a zinc polycarboxylate cement. Journal of Prosthetic Dentistry,
28, 264^71.
Jenkins, G. N. (1965). The equilibrium between plaque and enamel in relation to
dental caries. In Wolstenholme, G. E. W. & O'Connor, M. (eds.) Caries
Resistant Teeth, pp. 192-210. London: Churchill.
Jonck, L. M. & Grobbelaar, C. J. (1990). Ionos bone cement (glass-ionomer):
an experimental and clinical evaluation in joint replacement. Clinical
Materials, 6, 323-59.
Jonck, L. M., Grobbelaar, C. J. & Strating, H. (1989a). The biocompatibility of
glass-ionomer cement in joint replacement: bulk testing. Clinical Materials, 4,
85-107.
Jonck, L. M., Grobbelaar, C. J. & Strating, H. (1989b). Biological evaluation of
glass-ionomer cement (Ketac-0) as an interface material in total joint
replacement. A screening test. Clinical Materials, 4, 201-24.
Jurecic, A. (1973). Water-soluble acrylic copolymer dental cements. British
Patent 1,304,987; US Patent 3,741,926.
Kasten, F. H., Pineda, L. F., Schneider, P. E., Rawls, H. R. & Foster, T. A.
(1989). Biocompatibility testing of an experimental fluoride releasing resin
using human gingival epithelial cells in vitro. In Vitro Cellular Development
Biology, 25, 57-62.
Kawahara, H., Imanishi, Y. & Oshima, H. (1979). Biological evaluation on
glass-ionomer cement. Journal of Dental Research, 58, 1080-6.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1973). The properties of a
glass-ionomer cement. British Dental Journal, 135, 322-6.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1979). Glass ionomer cement
formulations. I. The preparation of novelfluoridealuminosilicateglasses high
in fluorine. Journal of Dental Research, 58, 1607-19.
Kidd, E. A. M. (1978). Cavity sealing ability of composite and glass ionomer
restoration; an assessment in vitro. British Dental Journal, 144, 139-42.
Kirkpatrick, R. J., Oestrike, R., Weis, jr, C. A., Smith, K. & Oldfield, E.
(1986). Higher resolution aluminium-27 and silicon-29 NMR spectroscopy
of glasses and crystals along the join calcium magnesium silicate-calcium
aluminium silicate (CaMgSi2 O6-CaAl2 SiO6). American Mineralogist, 71,
705-11.
Kleinburg, I. (1961). Studies on dental plaque. I. The effect of different
concentrations of glucose on the pH of dental plaque in vivo. Journal of
Dental Research, 40, 1087-110.
von Klotzer, W. T., Tronstad, L., Dowden, W. E. & Langeland, K. (1970).
Polycarboxylatzemente im physikalischen und biologischen Test. Deutsche
Zahndrztliche Zeitschrift, 25, 877-86.

183
Polyalkenoate cements
Knibbs, P. J., Plant, C. G. & Pearson, G. J. (1986a). The use of a glass-ionomer
cement to restore class III cavities. Restorative Dentistry, 2, 42-8.
Knibbs, P. J., Plant, C. G. & Pearson, G. J. (1986b). A clinical assessment of an
anhydrous glass-ionomer cement. British Dental Journal, 161, 99-103.
Knight, G. M. (1984). The use of adhesive materials in the conservative
restoration of selected posterior teeth. Australian Dental Journal, 29, 324-31.
Kohmura, T. T. & Ida, K. (1979). A new type of hydraulic cement. Journal of
Dental Research, 58, 1461.
Komatsu, H. (1981). Glass ionomer cement for caries prevention. Journal of
Conservative Dentistry, 24, 32-45.
Kramer, I. R. H. & McLean, J. W. (1952). Alterations in the staining reactions
of dentine resulting from a constituent of a new self-polymerizing resin.
British Dental Journal, 150, 150-3.
Kuhn, A. T. & Jones, M. P. (1982). A model for the dissolution and fluoride
release from dental cements. Biomaterials, Medical Devices and Artificial
Organs, 10, 281-93.
Kuhn, A. T., Setchell, D. J. & Teo, C. K. (1984). An assessment of the jet
method for solubility measurements of dental cements. Biomaterials, 5, 161-8.
Kuhn, A. T. & Wilson, A. D. (1985). The dissolution mechanisms of silicate and
glass-ionomer dental cements. Biomaterials, 6, 378-82.
Kumar, D., Ward, R. G. & Williams, D. J. (1961). Effects of fluorides on
silicates and phosphates. Discussions of the Faraday Society, 32, 147-54.
Kydonieus, A, F. (1980). Controlled Release Technologies. West Palm Beach,
Florida: CRC Press.
Lacefield, W. R., Reindl, M. C. & Retief, D. H. (1985). Tensile bond strength of
a glass-ionomer cement. Journal of Prosthetic Dentistry, 53, 194-8.
Lambe, T. W. (1951). Stabilization of soils with calcium acrylate. Journal of the
Boston Society of Civil Engineers, 38, 127-54.
Lambe, T. W. & Michaels, A. S. (1954). Altering soil properties with chemicals.
Chemical Engineering News, 32, 488-92.
Larsen, E. S. & Berman, H. (1934). Microscopic determination of the
nonopaque minerals. US Geological Survey Bulletin, No. 834. Washington.
Lawrence, L. G. (1979). Cervical glass-ionomer cement restorations: a clinical
study. Journal of the Canadian Dental Association, 45, 58-60.
Lawrence, L. G., Beagrie, G. S. & Smith, D. C. (1975). Zinc polyacrylate
cements as alloplastic bone implant. Journal of the Canadian Dental
Association, 41, 456-61.
Lawrence, L. G. & Smith, D. C. (1973). Strength modification of
polycarboxylate cements with fillers. Journal of the Canadian Dental
Association, 39, 405-9.
Lee, H. & Orlowski, J. (1973). Handbook of Dental Composite Restoratives, p.
3107. South El Monte, California: Lee Pharmaceuticals.
LeGeros, R. Z. & LeGeros, P. (1984). Phosphate minerals in human tissues. In
Nriagu, J. O. & Moore, P. B. (eds.) Phosphate Minerals, Chapter 12. Berlin:
Springer-Verlag.

184
References
Leirskar, J. & Helgeland, K. (1977). Toxicity of some dental cements in cell
culture systems. Scandinavian Journal of Dental Research, 85, 471-9.
Leirskar, J. & Helgeland, K. (1987). Mechanism of an in vitro toxicity of
restorative materials: pH, fluoride and zinc. International Endodontic Journal,
20, 246-7.
Levine, R. S., Beech, D. R. & Garton, B. (1977). Improving the bond strength
of poly aery late cements to dentine. A rapid technique. British Dental Journal,
143, 275-7.
Lloyd, C. H. & Adamson, M. (1987). The development of fracture toughness
and fracture strength in posterior restorative materials. Dental Materials, 3,
225-31.
Lloyd, C. H. & Mitchell, L. (1984). The fracture toughness of tooth coloured
restorative materials. Journal of Oral Rehabilitation, 11, 257-72.
Low, T. (1981). The treatment of a hypersensitive cervical abrasion cavity using
ASPA cement. Journal of Oral Rehabilitation, 8, 81-9.
Lowenstein, W. (1954). The distribution of aluminum in the tetrahedra of
silicates and aluminates. American Mineralogist, 39, 92-6.
Lux, H. (1939). 'Sauren' und 'Basen' in Schmelzfluss: Die Bestimmung der
Sauerstoffionen-Konzentration. Zeitschrift fur Elektrochemie, 45, 303-9.
McComb, D. & Ericson, D. (1987). Antimicrobial action of new, proprietary
lining cements. Journal of Dental Research, 66, 1025-8.
McKinney, J. E., Antonucci, J. M. & Rupp, N. W. (1985). Wear and micro-
hardness of a metal-filled ionomer cement. Journal of Dental Research, 64,
Special Issue 344 (IADR Abs 1577).
McKinney, J. E., Antonucci, J. M. & Rupp, N. W. (1987). Wear and
microhardness of glass-ionomer cement. Journal of Dental Research, 66,
134^9.
McLean, J. W. (1972). Polycarboxylate cements. Five years experience in
general practice. British Dental Journal, 132, 9-15.
McLean, J. W. (1980). Aesthetics in restorative dentistry: the challenge for the
future. British Dental Journal, 149, 368-72.
McLean, J. W. (1986). New concepts in cosmetic dentistry using glass-ionomer
cements and composites. Journal of the Californian Dental Association, April
1986, 20-7.
McLean, J. W. & Gasser, O. (1985). Glass-cermet cements. Quintessence, 16,
333-43.
McLean, J. W., Powis, D. R., Prosser, H. J. & Wilson, A. D. (1985). The use of
the glass-ionomer cement in bonding composite resins to dentine. British
Dental Journal, 158, 410-14.
McLean, J. W. & Wilson, A. D. (1974). Fissure sealing and filling with a
glass-ionomer cement. British Dental Journal, 136, 269-76.
McLean, J. W. & Wilson, A. D. (1977a). The clinical development of the
glass-ionomer cements. I. Formulations and properties. Australian Dental
Journal, 22, 31-6.
McLean, J. W. & Wilson, A. D. (1977b). The clinical development of the

185
Polyalkenoate cements
glass-ionomer cements. II. Some clinical applications. Australian Dental
Journal, 22, 120-7.
McLean, J. W. & Wilson, A. D. (1977c). The clinical development of the
glass—ionomer cements. III. The erosion lesion. Australian Dental Journal, 22,
190-5.
McLean, J. W., Wilson, A. D. & Prosser, H. J. (1984). Development and use of
water-hardening glass-ionomer luting cements. Journal of Prosthetic
Dentistry, 52, 175-81.
Main, J. H. P., Mock, D., Beagrie, G. S. & Smith, D. C. (1975). Investigations
of possible oncogenic action of zinc polycarboxylate cement. Journal of
Biomedical Materials Research, 9, 69-78.
Maldonado, A., Swartz, M. L. & Phillips, R. W. (1978). An in vitro study of
certain properties of a glass-ionomer cement. Journal of the American Dental
Association, 96, 785-91.
Mase, H. (1961). The reactivity of various silicate minerals toward acids.
Bulletin of the Chemical Society of Japan, 34, 214-25.
Mathis, R. S. & Ferracane, J. L. (1989). Properties of a glass-ionomer cement
resin-composite hybrid material. Dental Materials, 5, 355-8.
Mehrotra, R. C. & Bohra, R. (1983). Metal Carboxylates. London & New
York: Academic Press.
Mehrotra, R. C , Bohra, R. & Gauer, D. P. (1978). Metal Diketonates and Allied
Derivatives. London & New York: Academic Press.
de Mello, V. H. F., Hauser, E. A. & Lambe, T. W. (1953). Stabilization of soils.
US Patent 2,65\,6\9.
Meryon, S. D. & Smith, A. J. (1984). A comparison of fluoride release from
three glass ionomer cements and a polycarboxylate cement. International
Endodontic Journal, 17, 16-24.
Meryon, S. D. & Browne, R. M. (1984). In vitro cytotoxicity of a new
generation. Cell Biochemistry and Function, 2, 43-8.
Mesu, F. P. (1982). Degradation of luting cements measured in vitro. Journal of
Dental Research, 61, 665-72.
Mesu, F. P. & Reedijk, T. (1983). Degradation of luting cements measured in
vitro and in vivo. Journal of Dental Research, 62, 1236-40.
Michaeli, I. (1960). Ion-binding and the formation of insoluble polymethacrylic
salts. Journal of Polymer Science, 48, 291-9.
Minnesota Mining & Manufacturing Company. (1989). Vitrabond Light Cure
Glass Ionomer Cement Liner/Base. Report of Dental Products, 3M Health
Care.
Mitchem, J. C. & Gronas, D. G. (1978). Clinical evaluation of cement solubility.
Journal of Prosthetic Dentistry, 40, 453-6.
Mitchem, J. C. & Gronas, D. G. (1981). Continued evaluation of the
clinical solubility of luting cements. Journal of Prosthetic Dentistry, 45,
289-91.
Mitra, S. B. (1989). Photocurable ionomer cement systems. European Patent
Application 0 323 120 A2.
Mizrahi, E. & Smith, D. C. (1969a). Direct cementation of orthodontic brackets

186
References
to dental enamel. An investigation using zinc polycarboxylate cement. British
Dental Journal, 127, 371-410.
Mizrahi, E. & Smith, D. C. (1969b). The bond strengths of a zinc
polycarboxylate cement. British Dental Journal, 127, 410-14.
Mizrahi, E. & Smith, D. C. (1971). Direct attachment of orthodontic brackets to
dental enamel. British Dental Journal, 130, 392-6.
Mjor, I. A. (1985). Frequency of secondary caries at various anatomical
locations. Operative Dentistry, 10, 88-92.
Moore, B. K., Swartz, M. L. & Phillips, R. W. (1985). Abrasion resistance of
metal reinforced glass ionomer cement. Journal of Dental Research, 64,
Special Issue 371 (Abstract 1766).
Morawetz, H. (1965). Macromolecules in Solution. New York: Interscience
Publishers.
Mortimer, K. V. & Tranter, T. C. (1969). A preliminary laboratory evaluation
of polycarboxylate cements. British Dental Journal, 111, 365-9.
Moser, J. B., Brown, D. B. & Greener, E. H. (1974). Short-term bond strengths
between adhesive cements and dental alloys. Journal of Dental Research, 53,
1377-80.
Mount, G. J. (1984). Glass ionomer cements: clinical considerations. In Clarke,
J. W. (ed.) Clinical Dentistry, Chapter 20A. Philadelphia: Harper & Row.
Mount, G. J. (1986). Longevity of glass-ionomer cements. Journal of Prosthetic
Dentistry, 55, 682-5.
Mount, G. J. (1988). The tensile strength of the union between various glass
ionomer cements and various composite resins. Australian Dental Journal, 34,
136-46.
Mount, G. J. (1989). The wettability of bonding resins used in the composite
resin/glass—ionomer 'sandwich' technique. Australian Dental Journal, 34,
32-5.
Mount, G. J. (1990). An Atlas of Glass-ionomer Cements. London: Martin
Dunitz.
Murata, K. J. (1943). Internal structure of silicate minerals that gelatinize with
acid. American Mineralogist, 28, 545-62.
Murata, K. J. (1946). The significance of internal structure of gelatinizing
silicate minerals. Contributions to Geochemistry, 1942-5. In Geological Survey
Bulletin, No. 950, 25-33.
Nakamoto, K. (1963). Infrared Spectra of Inorganic and Coordination
Compounds. New York: John Wiley.
Nakamura, M., Kawahara, H., Imia, K., Tomoda, S., Kawata, Y. & Hikari, S.
(1983). Long-term biocompatibility test of composite resins and
glass-ionomer cement (in vitro). Dental Materials Journal, 1, 100-12.
Nicholson, J. W. (1992). The application of the reptation hypothesis to
polyelectrolyte biomaterials. Journal of Materials Science, Materials in
Medicine, 3, 157-9.
Nicholson, J. W., Anstice, H. M. & McLean, J. W. (1992). A preliminary report
on the effect of storage in water on the properties of commercial light-cured
glass-ionomer cements. British Dental Journal, 173, 98-101.

187
Polyalkenoate cements

Nicholson, J. W., Braybrook, J. H. & Wasson, E. A. (1991). The


biocompatibility of glass-poly(alkenoate) (glass-ionomer): a review. Journal
of Biomedical Science, Polymer Edition, 2, 277-85.
Nicholson, J. W., Brookman, P. J., Lacy, O. M., Sayers, G. S. & Wilson, A. D.
(1988a). A study of the nature and formation of zinc polyacrylate cement
using Fourier transform infrared spectroscopy. Journal of Biomedical
Materials Research, 22, 623-31.
Nicholson, J. W., Brookman, P. J., Lacy, O. M. & Wilson, A. D. (1988b).
Fourier transform infrared spectroscopic study of the role of tartaric acid in
glass-ionomer cements. Journal of Dental Research, 67, 1450—4.
0ilo, G. (1981). Bond strength of new ionomer cements to dentin. Scandinavian
Journal of Dental Research, 89, 344-7.
0ilo, G. (1988). Characterization of glass-ionomer filling materials. Dental
Materials, 4, 129-33.
0ilo, G. & Espevik, S. (1978). Stress/strain behaviour of some dental luting
cements. Acta Odontologica Scandinavica, 36, 45-9.
0ilo, G. & Evje, D. M. (1986). Film thickness of dental luting cements. Dental
Materials, 2, 85-9.
Oldfield, C. W. B. & Ellis, B. (1991). Fibrous reinforcement of glass-ionomer
cements. Clinical Materials, 7, 313-24.
Oosawa, F. (1971). Poly electrolytes. New York: Marcel Dekker.
O'Reilly, D. E. (1960). NMR chemical shifts of aluminum: experimental data
and variational calculation. Journal of Chemical Physics, yi, 1007-12.
Osborne, J. W. & Berry, T. G. (1986). Clinical assessment of glass ionomer
cements as Class III restorations: a one-year report. Dental Materials, 2,
147-50.
Osborne, J. W., Swartz, M. L., Goodacre, C. J., Phillips, R. W. & Gale, E. M.
(1978). A method for assessing the clinical solubility and disintegration of
luting cements. Journal of Prosthetic Dentistry, 40, 413-17.
Paddon, J. M. & Wilson, A. D. (1976). Stress relaxation studies on dental
materials. 1. Dental cements. Journal of Dentistry, 4, 183-9.
Paffenbarger, G. C. (1937). Dental silicate cements. In Judd, D. B. Optical
specifications of light scattering materials. Report No. 1026. Journal of the
National Bureau of Standards, 19, 314-16.
Paffenbarger, G. C , Schoonover, I. C. & Souder, W. (1938). Dental silicate
cements: Physical and chemical properties and a specification. Journal of the
American Dental Association, 25, 52-87.
Pameijer, C. H., Segal, E. & Richardson, J. (1981). Pulpal response to a
glass ionomer cement in primates. Journal of Prosthetic Dentistry, 46, 36-40.
Pameijer, C. H. & Stanley, H. R. (1988). Biocompatibility of a glass ionomer
luting agent in primates. Part I. American Journal of Dentistry, 1, 71-6.
Parker, R. S. R. (1974). Hardenable sheet materials. British Patent 1,504,972.
Paterson, R. C. (1976). Bacterial contamination and the exposed pulp. British
Dental Journal, 140, 231-6.
Pearson, G. J. (1991). Physical properties of glass-ionomer cements influencing
clinical behaviour. Clinical Materials, 1, 325-32.

188
References

Pearson, G. J. & Atkinson, A. S. (1991). Long-term flexural strength of


glass-ionomer cements. Biomaterials, 12, 658-60.
Peddy, M. (1981). The bond strength of polycarboxylic acid cements to dentine:
effect of surface modification and time after extraction. Australian Dental
Journal 26, 178-80.
Peters, W. J., Jackson, R. W., Iwako, K. & Smith, D. C. (1972). The biological
response to zinc polyacrylate carboxylate. Clinical Orthopaedics and Related
Research, 88, 228-33.
Peters, W. J., Jackson, R. W. & Smith, D. C. (1974). Studies of the stability and
toxicity of zinc polyacrylate cements (PA2). Journal of Biomedical Materials
Research, 8, 53-60.
Petty, W. (1980). Methylmethacrylate concentrations in tissues adjacent to bone
cement. Journal of Biomedical Materials Research, 14, 88-95.
Petzold, A. (1966). Die Reaktionfahigkeit einiger complexer Hydrosilikater
gegeniiber dem Angriff von Saure und ihre strukturelle und atomistische
Deutung. Wissenschaftliche Zeitschrift der Hochschule fur Architektur und
Bauwesen Weimar, 113, 343-50.
Phillips, R. W., Hamilton, A. J., Jendresen, M. D., McHarris, W. H. &
Schallhorn, R. G. (1986). Report of the Committee on Scientific Investigation
of the American Academy of Restorative Dentistry. Journal of Prosthetic
Dentistry, 55, 736-72.
Phillips, R. W. & Ryge, G. (1961). Adhesive Restorative Dental Materials.
Spencer, Indiana: Owen Litho Service.
Plant, C. G. (1970). The effect of polycarboxylate containing stannous fluoride
on the pulp. British Dental Journal, 135, 317-21.
Plant, C. G., Browne, R. M., Knibbs, P. J., Britton, A. S. & Sorhahan, T.
(1984). Pulpal effects of glass ionomer cements. International Endodontic
Journal, 17, 51-9.
Plant, C. G., Jones, I. H. & Wilson, H. J. (1972). Setting characteristics of lining
and cementing materials. British Dental Journal, 133, 21-4.
Plant, C. G. & Wilson, H. J. (1970). Early strengths of lining materials. British
Dental Journal, 129, 269-74.
Pluim, L. J. & Arends, J. (1981). In vivo solubility of dental cements. Journal of
Dental Research, 60, 1213, Abstract No. 61.
Pluim, L. J. & Arends, J. (1987). The relationship between salivary
properties and in vivo solubility of dental cements. Dental Materials, 3,
13-18.
Pluim, L. J., Arends, J. & Havinga, P. (1985). Comparison of in vivo and in vitro
solubility of 6 luting cements. Journal of Dental Research, 64, 714, Abstract
No. 98.
Pluim, L. J., Arends, J., Havinga, P., Jongebloed, W. J. & Stokroos, I. (1984).
Quantitative cement solubility experiments in vivo. Journal of Oral
Rehabilitation, 11, 171-9.
Pople, I. K. & Phillips, H. (1988). Bone cement and the liver. A dose-related
effect? Journal of Bone and Joint Surgery, 70B, 364-6.
Posner, A. S. (1961). The structure of the mineral portion of teeth. In Phillips,

189
Polyalkenoate cements
R. W. & Ryge, G. (eds.) Adhesive Restorative Dental Materials, pp. 15-27.
Spencer, Indiana: Owen Litho Service.
Potter, W. D., Barclay, A. C, Dunning, R. J. & Parry, R. R. (1977). Curable
compositions. US Patent 4,034,327.
Potter, W. D., Barclay, A. C, Dunning, R. J. & Parry, R. R. (1979). Calcium
fluoroaluminosilicate glass. British Patent 1,554,555.
Powers, J. M., Farah, J. W. & Craig, R. G. (1976). Modulus of elasticity and
strength properties of dental cements. Journal of the American Dental
Association, 92, 588-91.
Powers, J. M., Johnson, Z. G. & Craig, R. G. (1974). Physical and mechanical
properties of zinc polyacrylate dental cements. Journal of the American Dental
Association, 88, 380-3.
Powis, D. R., Folleras, T., Merson, S. A. & Wilson, A. D. (1982). Improved
adhesion of a glass ionomer cement to dentin and enamel. Journal of Dental
Research, 61, 1416-22.
Powis, D. R., Prosser, H. J. & Wilson, A. D. (1988). Long-term monitoring of
microleakage of dental cements by radiochemical diffusion. Journal of
Prosthetic Dentistry, 59, 651-7.
Prati, C, Nucci, C. & Montanari, G. (1989). Effects of acid and cleansing
agents on shear bond strength and marginal microleakage of glass-ionomer
cements. Dental Materials, 5, 260-5.
Prosser, H. J., Jerome, S. M. & Wilson, A. D. (1982). The effect of additives on
the setting properties of a glass-ionomer cement. Journal of Dental Research,
61, 1195-8.
Prosser, H. J., Powis, D. R., Brant, P. & Wilson, A. D. (1984). Characterization
of glass-ionomer cements. 7. The physical properties of current materials.
Journal of Dentistry, 12, 231^40.
Prosser, H. J., Powis, D. R. & Wilson, A. D. (1986). Glass-ionomer
cements of improved flexural strength. Journal of Dental Research, 65,
146-8.
Prosser, H. J., Richards, C. P. & Wilson, A. D. (1982). NMR spectroscopy of
dental cements. II. The role of tartaric acid in glass-ionomer cements. Journal
of Biomedical Materials Research, 16, 431-45.
Prosser, H. J. & Wilson, A. D. (1979). Litho-ionomer cements and their
technological applications. Journal of Chemical Technology and Biotechnology,
29, 69-87.
Rabinovich, E. M. (1983). On the structural role of fluorine in silicate glasses.
Journal of Physical Chemistry, 24, 54-6.
Rawson, H. (1967). Inorganic Glass-Forming Systems, Chapter 2. London &
New York: Academic Press.
Ray, N. H. (1975). Inorganic glasses as ionic polymers. In Holliday, L. (ed.)
Ionic Polymers, Chapter 8. London: Applied Science Publishers.
Ray, N. H. (1983). Oxide glasses as ionic polymers. In Wilson, A. D. & Prosser,
H. J. (eds.) Developments in Ionic Polymers- 1, Chapter 2. London and New
York: Applied Science Publishers.
Retief, D. H., Bradley, E. L., Denton, J. C. & Switzer, P. (1984). Enamel and

190
References
cementum fluoride uptake from a glass ionomer cement. Caries Research, 18,
250-7.
Richter, W. A. & Ueno, H. (1975). Clinical evaluation and dental clinical
durability. Journal of Prosthetic Dentistry, 33, 294-9.
Risbud, S. H., Kirkpatrick, R. J., Taglialavore, A. P. & Montez, B. (1987).
Solid-state NMR evidence of 4-, 5- and 6-fold aluminium sites in roller-
quenched SiO2-Al2 O3 glasses. Journal of the American Ceramic Society, 70,
C10-12.
Robbins, J. W. & Cooley, R. L. (1988). Microleakage of Ketac-Silver in the
tunnel preparation. Operative Dentistry, 13, 8-11.
Rolla, G. (1977). Effects of fluoride on initiation of plaque formation. Caries
Research, 11, 243-61.
Saito, C, Sakai, Y., Node, H. & Fusayama, T. (1976). Adhesion of
polycarboxylate cements to dental casting alloys. Journal of Prosthetic
Dentistry, 35, 543-8.
Seed, I. & Wilson, A. D. (1980). Poly(carboxylic acid) hardenable compositions.
British Patent Application GB 2,028,855A.
Schmidt, W., Purrmann, R., Jochum, P. & Glasser, O. (1981a). Calcium
aluminosilicate glass powder and its use. European Patent Application 23,013.
Schmidt, W., Purrmann, R., Jochum, P. & Glasser, O. (1981b). Mixing
compounds for glass-ionomer cements and use of a copolymer for preparing
the mixing components. European Patent Application 24,056.
Scott, R. P., Jackson, A. M. & Wilson, A. D. (1990). Adhesion of carboxylate
cements to hydroxyapatite. II. Adsorption of aromatic carboxylates.
Biomaterials, 11, 341-4.
Setchell, D. J., Teo, C. K. & Kuhn, A. T. (1985). The relative solubilities of four
modern glass-ionomer cements. British Dental Journal, 158, 220-2.
Shimoke, H., Komatsu, H. & Matsui, I. (1987). Fluoride content in human
enamel after removal of the applied glass ionomer cement. Journal of Dental
Research, 66, Special Issue 131, Abstract No. 196.
Sidler, P. & Strub, J. R. (1983). In-vivo Untersuchung der Loslichkeit und des
Abdichtungsvermogens von drei Befestigungszementen. Deutsche
Zahndrztliche Zeitschrift, 38, 564-71.
Silverstone, L. M. (1982). The structure and characteristics of human dental
enamel. In Smith, D. C. & Williams, D. F. (eds.) Biocompatibility of Dental
Materials. Volume I. Characteristics of Dental Tissues and their Response to
Dental Materials, Chapter 2. Boca Raton: CRC Press Inc.
Simmonds, J. J. (1983). The miracle mixture glass-ionomer and alloy powder.
Texas Dentist, October 6-12.
Skinner, J. C, Prosser, H. J., Scott, R. P. & Wilson, A. D. (1986). Adhesion of
carboxylate cements to hydroxy apatite. I. The effect of the structure of
aliphatic carboxylates on their uptake by hydroxy apatite. Biomaterials, 7,
438^0.
Smink, L. A. & Arends, J. (1980). Oplosbaarheid en disintegratie van
Tandhelkundige cementen in vitro. Nederlands Tijdschrift voor Tandheelkunde,
87, 389-93.

191
Polyalkenoate cements
Smith, D. C. (1968). A new dental cement. British Dental Journal, 125, 381^4.
Smith, D. C. (1969). Improvements relating to surgical cements. British Patent
1,139,430.
Smith, D. C. (1971). A review of the zinc polycarboxylate cements. Journal of
the Canadian Dental Association, 37, 22-9.
Smith, D. C. (1975). Approaches to adhesion to tooth structure. In Silverstone,
L. M. & Dogson, I. L. (eds.) The Acid Etch Technique, pp. 119-38. St Paul,
Minnesota: North Central Publishing Company.
Smith, D. C. (1977). Past, present and future of dental cements. In Craig, R. G.
(ed.) Dental Materials Review. Ann Arbor: University of Michigan Press.
Smith, D. C. (1982a). Composition and characteristics of dental cements. In
Smith, D. C. & Williams, D. F. (eds.) Biocompatibility of Dental Materials.
Volume II. Biocompatibility of Preventive Dental Materials and Bonding
Agents, Chapter 8. Boca Raton: CRC Press Inc.
Smith, D. C. (1982b). Tissue reactions to cements. In Smith, D. C. & Williams,
D. F. (eds.) Biocompatibility of Dental Materials. Volume II. Biocompatibility
of Preventive Dental Materials and Bonding Agents, Chapter 10. Boca Raton:
CRC Press Inc.
Smith, D. C. & Cartz, L. (1973). Crystalline interface formed by polyacrylic acid
and tooth enamel. Journal of Dental Research, 52, 1155.
Smith, D. C , Ruse, D. N. & Zuccolin, D. (1988). Some characteristics of glass
ionomer cement lining materials. Canadian Dental Journal, 54, 903-8.
Sneed, W. D. & Looper, S. W. (1985). Shear bond strength of a composite resin
to an etched glass-ionomer. Dental Materials, 1, 127-8.
Spangberg, L., Rodrigues, H. & Langeland, K. (1974). Biologic effects of dental
materials. 4. Effect of polycarboxylate cements on HeLa cells in vitro. Oral
Surgery, Oral Medicine, Oral Pathology, 37, 113-17.
Steinke, R., Newcomer, P., Komarneni, S. & Roy, R. (1988). Dental cements:
investigation of chemical bonding. Materials Research Bulletin, 23, 13-22.
Stephan, R. M. (1940). Changes in the hydrogen-ion concentration on tooth
surfaces and in carious lesions. Journal of the American Dental Association,
27, 718-23.
Suzaki, M. (1976). Dental cement composition. US Patent 3,962,267.
Swartz, M. L., Phillips, R. W. & Clark, H. E. (1984). Long-term F release from
glass ionomer cements. Journal of Dental Research, 63, 158-60.
Swift, E. J. (1988a). Silver-glass ionomers. A status report for the American
Journal of Dentistry. American Journal of Dentistry, 1, 81-3.
Swift, E. J. (1988b). An update on glass ionomer cements. Operative Dentistry,
19, 125-30.
Swift, E. J. & Dogan, A. U. (1990). Analysis of glass-ionomer cement with
use of scanning electron microscopy. Journal of Prosthetic Dentistry, 64,
167-74.
Tanzer, J. M. (1989). On changing the cariogenic chemistry of coronal plaque.
Journal of Dental Research, 68, 1576-87.
Tay, W. M. & Braden, M. (1988). Fluoride ion diffusion from polyalkenoate
(glass-ionomer) cements. Biomaterials, 9, 454-6.

192
References

Taylor, H. F. W. (1966). The Chemistry of Cements. Lecture Series 1966, No. 2.


London: Royal Institute of Chemistry.
Ten Cate, A. R. & Torneck, C. (1982). The dentin-pulp complex: development,
structure and repair. In Smith, D. C. & Williams, D. F. (eds.) Biocompatibility
of Dental Materials. Volume I. Characteristics of Dental Tissues and their
Response to Dental Materials, Chapter 3. Boca Raton: CRC Press Inc.
Theuniers, G. (1984). Een onderzock naar een duurzame afdichting door kroon-
en brugcementen. Thesis. Leuvan.
Thornton, J. B., Retief, D. H. & Bradley, E. L. (1986). Fluoride release from
and tensile bond strength of Ketac-Fil and Ketac-Silver to enamel and
dentine. Dental Materials, 2, 241-5.
Tobias, R. S., Browne, R. M., Plant, C. G. & Ingram, D. V. (1978). Pulpal
response to a glass ionomer cement. British Dental Journal, 144, 345-50.
Tobias, R. S., Browne, R. M. & Wilson, C. A. (1985). Antibacterial activity of
dental restorative materials. International Dental Research, 18, 161-71.
Tobias, R. S., Plant, C. G., Browne, R. M., Knibbs, P. J. & Britton, A. (1987).
Pulpal response to an anhydrous glass-ionomer luting cement. Journal of
Dental Research, 66, 836. Abstract 12.
Trap, H. J. L. & Stevals, J. M. (1959). Physical properties of invert glasses.
Glastechnische Berichte, 32K, VI/31-52.
Trautschold, R. (1940). Non-corrosive, non-metallic, materials. Rayon Textile
Monthly, 21, 373-4.
Turner, R. T., Francis, R., Brown, D., Garand, J., Hannon, K. S. & Bell, N. H.
(1989). The effects of fluoride on bone implant histomorphology in growing
rats. Journal of Bone Mineralogy Research, 4, 477-84.
Tyas, M. J. (1977). A method for the in vitro toxicity testing of dentine
restorative materials. Journal of Dental Research, 56, 1285.
Tyas, M. J. & Beech, D. R. (1985). Clinical performance of three restorative
materials for cervical abrasion lesions. Australian Dental Journal, 30, 260-4.
Tyas, M. J., Alexander, S. B., Beech, D. R., Brockenhurst, P. J. & Cook, W. D.
(1988). Bonding - retrospect and prospect. Australian Dental Journal, 33,364-74.
Volf, M. B. (1984). Chemical Approach to Glass, Chapter 2. Amsterdam, etc.:
Elsevier.
van de Voorde, A. (1988). Clinical uses of glass ionomer cement: a literature
review. Quintessence International, 19, 53-60.
Wall, F. T. & Drenan, J. W. (1951). Gelation of polyacrylic acids by divalent
cations. Journal of Polymer Science, 7, 83-8.
Walls, A. W. G. (1986). Glass polyalkenoate (glass-ionomer) cements: a review.
Journal of Dentistry, 14, 231-46.
Walls, A. W. G., McCabe, J. R. & Murray, J. J. (1985). An erosion test for
dental cements. Journal of Dental Research, 64, 1100-4.
Walls, A. W. G., McCabe, J. F. & Murray, J. J. (1988). The effect of variation
of glass polyalkenoate (ionomer) cements. British Dental Journal, 164, 141-4.
Walls, A. W. G., Murray, J. J. & McCabe, J. F. (1988). The use of glass
polyalkenoate (ionomer) cements in deciduous dentition. British Dental
Journal, 165, 13-17.

193
Polyalkenoate cements
Wasson, E. A. & Nicholson, J. W. (1990). A study of the relationship between
setting chemistry and properties of modified glass-poly(alkenoate) cements.
British Polymer Journal 23, 179-83.
Wasson, E. A. & Nicholson, J. W. (1991). Studies on the setting chemistry of
glass-ionomer cements. Clinical Materials, 7, 289-93.
Waters, D. N. & Henty, M. S. (1977). Raman spectra of aqueous solutions of
hydrolysed aluminium(III) salts. Journal of the Chemical Society: Dalton
Transactions, 243-5.
Watts, D. C , Combe, E. C. & Greener, E. H. (1979). Effect of storage
conditions on the mechanical properties of polyelectrolyte cements. Journal of
Dental Research, 58, Special Issue C, Abstract No. 18.
Wei, S. H. Y. (1985). Clinical Uses of Fluoride. Philadelphia: Lea & Febiger.
Welker, D. & Neupert, G. (1974). Vergleichender biologischer Test von
Polyakrylate- und Phosphatzement an Monolayer-Kulturen. Stomatologie
(DDR), 24, 602-10.
Welsh, E. L. & Hembree, J. H. (1985). Microleakage of the gingival wall with
four class V anterior restorative materials. Journal of Prosthetic Dentistry, 54,
370-2.
Wesenberg, G. & Hals, E. (1980). The in vitro effect of a glass ionomer cement
on dentine and enamel wall. Journal of Oral Rehabilitation, 7, 35-42.
Weyl, W. A. & Marboe, E. C. (1962). The Constitution of Glasses: a Dynamic
Interpretation. Volume 1. Fundamentals of the Structure of Inorganic Liquids
and Solids. New York: Interscience Publishers.
Williams, J. & Billington, R. W. (1989). Increase in compressive strength of
glass-ionomer cements with respect to time: a guide to their use in the
posterior deciduous dentition. Journal of Oral Rehabilitation, 16,
475-9.
Williams, J. & Billington, R. W. (1991). Changes in compressive strength of
glass ionomer restorative materials with respect to time periods of 24 h to 4
months. Journal of Oral Rehabilitation, 18, 163-8.
Williams, J. A., Billington, R. W. & Pearson, G. J. (1992). The comparative
strengths of commercial glass-ionomer cements with and without metal
additions. British Dental Journal, 111, 279-82.
Wilson, A. D. (1968). Dental silicate cements. VII. Alternative liquid acid
formers. Journal of Dental Research, 47, 1133-6.
Wilson, A. D. (1974). Alumino-silicate poly aerylie acid and related cements.
British Polymer Journal, 6, 165-79.
Wilson, A. D. (1975a). Dental cements - general. In von Fraunhofer, J. A. (ed.)
Scientific Aspects of Dental Materials, Chapter 4. London and Boston:
Butterworths.
Wilson, A. D. (1975b). Zinc oxide dental cements. In von Fraunhofer, J. A. (ed.)
Scientific Aspects of Dental Materials, Chapter 5. London and Boston:
Butterworths.
Wilson, A. D. (1975c). Dental cements based on ion-leachable glasses. In von
Fraunhofer, J. A. (ed.) Scientific Aspects of Dental Materials, Chapter 6.
London and Boston: Butterworths.

194
References
Wilson, A. D. (1978a). The chemistry of dental cements. Chemical Society
Reviews, 7, 265-96.
Wilson, A. D. (1978b). Glass-ionomer cements - ceramic polymers. In Young,
J. F. (ed.) Chemical Research Progress. Columbus, Ohio: American Ceramic
Society.
Wilson, A. D. (1982). The nature of the zinc polycarboxylate cement matrix.
Journal of Biomedical Materials Research, 16, 549-57.
Wilson, A. D. (1989). Developments in glass-ionomer cements. International
Journal of Prosthodontics, 2, 438-46.
Wilson, A. D. (1990). Resin modified glass-ionomer cements. International
Journal of Prosthodontics, 3, 425—46.
Wilson, A. D. (1991). Glass-ionomer cement - origins, development and future.
Clinical Materials, 7, 275-82.
Wilson, A. D. & Crisp, S. (1975). Ionomer cements. British Polymer Journal, 1,
279-96.
Wilson, A. D. & Crisp, S. (1976). Poly(carboxylate) cements. British Patent
1,422,337.
Wilson, A. D. & Crisp, S. (1977). Polymer-clay compounds and soil treatment.
In Organolithic Macromolecular Materials, Chapter 5. London: Applied
Science Publishers.
Wilson, A. D. & Crisp, S. (1980). Dental cement containing poly(carboxylic
acid), chelating agent and glass ionomer powder. US Patent 4,209,434.
Wilson, A. D., Crisp, S. & Abel, G. (1977). Characterization of glass-ionomer
cements. 4. Effect of molecular weight on physical properties. Journal of
Dentistry, 5, 117-20.
Wilson, A. D., Crisp, S. & Ferner, A. J. (1976). Reactions in glass ionomer
cements: IV. Effect of chelating comonomers on setting behavior. Journal of
Dental Research, 55, 489-95.
Wilson, A. D., Crisp, S., Lewis, B. G. & McLean, J. W. (1977). Experimental
luting agents based on the glass-ionomer cements. British Dental Journal, 142,
117-22.
Wilson, A. D., Crisp, S. & Paddon, J. M. (1981). The hydration of a
glass-ionomer (ASPA) cement. British Polymer Journal, 13, 66-70.
Wilson, A. D., Crisp, S., Prosser, H. J., Lewis, B. G. & Merson, S. A. (1980).
Aluminosilicate glasses for polyelectrolyte cements. Industrial & Engineering
Chemistry Product Research & Development, 19, 263-70.
Wilson, A. D. & Ellis, J. (1989). Poly-vinylphosphonic acid glass ionomer
cement. British Patent Application 8,924,129.3.
Wilson, A. D., Groffman, D. M. & Kuhn, A. T. (1985). The release of fluoride
and other chemical species from a glass-ionomer cement. Biomaterials, 6, 431-3.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986a). An
evaluation of the significance of the impinging jet method for measuring the
acid erosion of dental cements. Biomaterials, 7, 55-60.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986b). A study
of variables affecting the impinging jet method for measuring the erosion of
dental cements. Biomaterials, 7, 217-20.

195
Polyalkenoate cements
Wilson, A. D., Hill, R. G., Warrens, C. P. & Lewis, B. G. (1989). The influence
of poly(acrylic acid) molecular weight on some properties of glass-ionomer
cement. Journal of Dental Research, 68, 89-94.
Wilson, A. D. & Kent, B. E. (1970). Dental silicate cements. IX. Decomposition
of the powder. Journal of Dental Research, 49, 7-13.
Wilson, A. D. & Kent, B. E. (1971). The glass-ionomer cement, a new
translucent cement for dentistry. Journal of Applied Chemistry and
Biotechnology, 21, 313.
Wilson, A. D. & Kent, B. E. (1972). A new translucent cement for dentistry.
British Dental Journal, 132, 133-5.
Wilson, A. D. & Kent, B. E. (1973). Surgical cements. British Patent 1,316,129.
Wilson, A. D. & Kent, B. E. (1974). Poly(carboxylic acid)-fluoroaluminosilicate
glasses surgical cement. US Patent 3,814,717.
Wilson, A. D. & Lewis, B. G. (1980). The flow properties of dental cements.
Journal of Biomedical Materials Research, 14, 383-91.
Wilson, A. D. & McLean, J. W. (1988). Glass-ionomer Cement. Chicago,
London, etc.: Quintessence Publishing Company Inc.
Wilson, A. D., Paddon, J. M. & Crisp, S. (1979). The hydration of dental
cements. Journal of Dental Research, 58, 1065-71.
Wilson, A. D. & Prosser, H. J. (1982). Biocompatibility of the glass ionomer
cement. Journal of the Dental Association of South Africa, 37, 872-9.
Wilson, A. D. & Prosser, H. J. (1984). A survey of inorganic and polyelectrolyte
cements. British Dental Journal, 157, 449-54.
Wilson, A. D., Prosser, H. J. & Powis, D. M. (1983). Mechanism of adhesion of
polyelectrolyte cement to hydroxyapatite. Journal of Dental Research, 62,
590-2.
Wilson, M. A. & Combe, E. C. (1991). Effects of glass composition and
pretreatment on the reactivity of a novel glass polyalkenoate (glass ionomer)
dental cement. Clinical Materials, 7, 15-21.
Wood, D. & Hill, R. (1991a). Structure-property relationships in ionomer
glasses. Clinical Materials, 1, 301-12.
Wood, D. & Hill, R. (1991b). Glass ceramic approach to controlling the
properties of a glass-ionomer cement. Biomaterials, 12, 164-70.
Woodberry, N. T. (1961). A new, anionic, polyacrylamide flocculant. Tappi, 44,
156A-60A.
Wygant, J. F. (1958). Cementitious bonding in ceramic fabrication. In Kingery,
W. D. (ed.) Ceramic Fabrication Processes, pp. 171-88. New York: John
Wiley & Sons.
Yoshii, E., Homma, T., Hirota, K. & Tomioka, K. (1987). Cytotoxic evaluation
of the improved glass-ionomer cement. Journal of Dental Research, 66,
Special Issue 133, Abstract No. 215.
Zachariasen, W. H. (1932). The atomic arrangement in glass. Journal of the
American Chemical Society, 54, 3841-51.
van Zeghbroeck, L. (1989). Bond capacity of adhesive luting cements. Thesis for
Doctor in de Tandheelkunde, Leuvan University, Belgium.

196
6 Phosphate bonded cements

6.1 General
The phosphate bonded cements described in this chapter are the products
of the simple acid-base reaction between an aqueous solution of
orthophosphoric acid and a basic oxide or silicate. Such reactions take
place at room temperature. Excluded from this chapter are the cemen-
titious substances that are formed by the heat treatment of aqueous
solutions of acid metal phosphates.
The most important of these are the refractory cements formed by the
heat treatment of aluminium acid phosphate solutions. This subject has
been well reviewed by Kingery (1950a), Morris et al. (1977), Cassidy (1977)
and O'Hara, Duga & Sheets (1972). The chemistry of these binders is
extremely complex as the action of heat on acid phosphates gives rise to
polymeric phosphates, with P-O-P linkages, and these are very complex
systems (Ray, 1979).
Here we are concerned with the cement-forming reaction between
orthophosphoric acid solutions and basic oxides and silicates where the
reaction is much simpler. Polymeric phosphates are not involved, there are
no P-O-P bonds, and the structural unit is the simple [POJ tetrahedron.

6.1.1 Orthophosphoric acid solutions


Concentrated solutions of orthophosphoric acid, often containing metal
salts, are used to form cements with metal oxides and aluminosilicate
glasses. Orthophosphoric acid, often referred to simply as phosphoric acid,
is a white crystalline solid (m.p. 42-35 °C) and there is a crystalline
hemihydrate, 2H 3 PO 4 .H 2 O, which melts at 29-35 °C. The acid is tribasic
and in aqueous solution has three ionization constants (pKa): 2-15, 7-1 and
12-4.

197
Phosphate bonded cements
X-ray diffraction (XRD) studies have shown that the crystalline acid and
its hydrate contain tetrahedral [POJ groups (Van Wazer, 1958). In the
anhydrous acid, three of the oxygen atoms are bonded to hydrogen atoms
and the P-O bonds are 0-157 nm or 0-158 nm in length. The P-O bond of
the fourth O has more n character than the others and is shorter (0-152 nm).
The [POJ tetrahedra are interconnected by hydrogen bonds and there are
also internal hydrogen bonds between pairs of O atoms within the [POJ
tetrahedra. Radial distribution curves from XRD studies indicate that the
intramolecular hydrogen bonds persist in 86 % phosphoric acid solution.
Mostly, a single hydrogen bond connects any two [POJ groups, but double
and triple hydrogen bonding occurs to a lesser extent. In more dilute
solution (54 %) the [POJ groups are linked to the water lattice rather than
to other [POJ groups. Raman spectroscopy supports these structural
views. In 75 % solution not all [POJ groups are hydrogen-bonded to water
molecules (Wilson & Mesley, 1968).
Results from infrared spectroscopy indicate that the only species present
in 50 % phosphoric acid are H3PO4, H2PO4 and their oligomers (Wilson &
Mesley, 1968). There is evidence that H6P2O8, the phosphoric acid dimer,
and H5P2O~, the triple ion HgPO^. H + . H2PO~, are also present (Elmore,
Mason & Christensen, 1946; Selvaratnam & Spiro, 1965). Akitt, Green-
wood & Lester (1971), on the basis of 31P NMR studies, suggest further
that there are oligomers of the type (H3P4O)W.

6.1.2 Cations in phosphoric acid solutions


Cement-forming phosphoric acid liquids nearly always contain cations.
The most important of these are aluminium and zinc, but other metals may
be used. Manly et al. (1951) have laid down criteria for the choice of
modifying metals. They must be moderately soluble in oxide solution, not
form coloured sulphides and be non-toxic. To these criteria, following
Kingery (1950a,b), may be added another, that the modifying cations must
be capable of remaining in vitreous form in the cement gel. This criterion
is satisfied only by cations having a low coordination number, multiple
charge and small ionic radius, that is having a high ionic potential. These
are ions of amphoteric or weakly basic metals such as aluminium, zinc,
beryllium and magnesium. More basic metals, for example calcium,
barium and thorium, weaken the cement.
As noted above, however, aluminium and zinc are the most important
and are often found in combination in the liquids used for the zinc

198
General

phosphate and dental silicate cement. The presence of aluminium in


phosphoric acid solution serves to retard the setting reaction in zinc
phosphate cements and accelerate it in dental silicate cements.
Both these metals are soluble to a limited extent in phosphoric acid
solutions. Phase diagrams have been constructed for the systems
ZnO-P2O5-H2O (Figure 6.1) and A12O3-P2O5-H2O (Figure 6.2). Stable
zinc phosphate phases that have been found to exist are Zn3(PO4)2. 4H2O,
ZnHPO4. 3H2O, ZnHPO4. H2O, Zn(H2PO4)2. 2H2O and
Zn(H2PO4)2.2H3PO4 (Eberly, Gross & Crowell, 1920; Salmon & Terrey,
1950). All are apparently crystalline. Stable aluminium phosphate phases
that have been found to exist are 2A1PO4.7H2O, 2A1PO4.4H2O,
2A1PO4.2H3PO46H2O, 2A1PO4.2H3PO4.3H2O; with the possible excep-
tion of the first all are crystalline (Jameson & Salmon, 1954). Although
these phase diagrams have been used to deduce the course of cement
formation, they are of limited use because it is doubtful whether conditions
of thermodynamic equilibrium are reached; moreover, many cements are
mainly amorphous.
H2O

range
for
cements

ZnO P2O5

Figure 6.1 The system ZnO-P 2 O 5 -H 2 O (Salmon & Terrey, 1950).

199
Phosphate bonded cements

The actions of zinc and aluminium differ. In general, metal ions such as
zinc merely serve to neutralize the acid and are present in solution as simple
ions (Holroyd & Salmon, 1956; O'Neill et ai, 1982). But aluminium has a
special effect: in contrast to zinc, it prevents the formation of crystallites
during the cement-forming reaction in zinc phosphate cements.
Aluminium has long been known to form complexes with phosphoric
acid (Bjerrum & Dahm, 1931; Jameson & Salmon, 1954; Genge et aL,
1955; Holroyd & Salmon, 1956; Salmon & Wall, 1958; Van Wazer, 1958;
Van Wazer & Callis, 1958; Genge & Salmon, 1959). When it is dissolved in
phosphoric acid solution the viscosity of the solution increases sharply as,
according to Sveshnikova & Zaitseva (1964), aluminophosphoric acids are
formed which behave as polyelectrolytes. A number of workers using 31P
NMR spectroscopy have found evidence for the formation of alumino-
phosphoric acid complexes (Akitt, Greenwood, & Lester, 1971; O'Neill et
al., 1982). Combining these NMR observations with those from infrared
spectroscopy (Wilson & Mesley, 1968) the species present in 50%

range
for
cements

A12O3 P2O5

Figure 6.2 The system A1 2 O 3 -P 2 O 5 -H 2 O (Jameson & Salmon, 1954).

200
General

Table 6.1. Cement-forming oxides

Condition Modification
of oxide of liquid

BeO
Be(OH)2
ZnO Calcined Al salt
CuO Calcined
Cu 2 O
MgO Calcined Al or NH 4 salt
CaO Calcined Al salt
Bi2O3
CdO
SnO
Pb 3 O 4
Co(OH) 3
Aluminosilicate glass

phosphoric acid solution containing aluminium appear to be H 3 PO 4 ,


H 6 P 2 O 8 , H 2 P O " , H 5 P 2 O - , A l H 3 P O r , A1H 2 POJ + , A1(H 2 PO 4 )+ and
A1(H 3 PO 4 ) W , where n ^ 2, of u n k n o w n protonation. Binuclear aluminium
phosphate complexes are also present and Salmon & Wall (1958) consider
that bridge structures exist. All this points to the formation of an
aluminophosphate polymer in the solution based on - P - O - A l - b o n d i n g .

6.1.3 Reactions between oxides and phosphoric acid solutions

In a classic study, Kingery (1950b) examined a large number of oxides for


cement formation with orthophosphoric acid. He observed three types of
reaction: no reaction, violent reaction with crystallization, and controlled
reaction with cement formation.
There was no reaction with oxides of an acidic or inert nature, for
example SiO2, A12O3, ZrO 2 and Co 2 O 3 . There was violent reaction with
reactive oxides, yielding non-cementitious products, which were crys-
talline, porous and friable. This type of reaction tended to occur when the
oxides were alkaline, although it could be affected by calcining the oxide.
Examples cited were calcined CaO, SrO and BaO, and uncalcined MgO
and La 2 O 3 . However, calcined CaO did form cements when ortho-
phosphoric acid was partly neutralized by CaO.

201
Phosphate bonded cements
Lastly, there was the cementitous reaction which Kingery (1959b)
reported with BeO, Be(OH)2, CuO, Cu2O, CdO, SnO and Pb3O4. In
addition, calcined ZnO and MgO formed cements.
These observations require some comment and amendment. The
formation of cements with ZnO, CuO and Bi2O3 has been known for many
years and they are found in dental cements (Wilson, 1975a,b; 1978). Also,
although Co2O3 does not form a cement, Co(OH)3 does (Prosser et al.,
1986). MgO is a doubtful cement-former with orthophosphoric acid, but
forms useful cements with ammonium acid phosphate and aluminium-
containing orthophosphoric acid solutions (Finch & Sharp, 1989). There is
also cement formation with calcium aluminosilicate glasses (Wilson 1975c,
1978). This material, the dental silicate cement, is unusual in being
translucent.
Table 6.1 summarizes known basic cement-formers based on the
observations of the workers cited above.
Because of its importance and the breadth of the investigation, the work
of Kingery (1950b) requires critical examination. He considered that the
essential feature of phosphate-bonded cements was an acid phosphate
matrix. Extended hydrogen bridges between acid phosphate groups can
then be cited as the matrix-forming bond (Wygant, 1958). However, we
consider that Kingery was mistaken, for his work has its limitations.
Kingery based his conclusions on a literature survey and some
experimental studies. The literature survey indicated that in many cases the
reaction product of a metal oxide with orthophosphoric acid was an acid
phosphate. However, these studies did not relate to cements, where the
metal oxide is always present in excess. For example, he cites the phase
diagram of Eberly, Gross & Crowell (1920) as showing ZnHPO4. 3H2O as
the reaction product. But in the presence of excess ZnO the phase diagram
shows Zn3(PO4)2. 4H2O as the stable species. Kingery also neglected to
consider the effect of phosphoric acid concentration on the nature of the
reaction product.
Kingery himself used XRD, but except in one case failed to positively
identify crystalline hydrogen phosphates in any of the cements he
examined. He found either the neutral orthophosphate or unidentified
crystalline species and then the lines were weak.
Although an acid phosphate matrix cannot be excluded it is not essential
for cement formation. In fact, it must be remembered that when these
cements are prepared the oxide or silicate powder is normally in excess of
that required for the reaction. Under these conditions most oxides (MgO

202
General

is a notable exception) form neutral orthophosphates. An acid phosphate


matrix is to be expected only when there is excess acid in the cement mix;
the only practical cements where these conditions obtain are those used for
controlled release which are designed to be hydrolytically unstable (Prosser
et al., 1986). Hydrated neutral orthophosphates are the normal reaction
products in zinc phosphate and dental silicate cements, both of which have
been studied in detail. The reaction between magnesium oxide and
phosphoric acid is an exception. An acid phosphate is formed, but it is
soluble in water (Finch & Sharp, 1989) and, in fact, MgO forms useful
cements only with ammonium dihydrogen phosphate. Hydrogen bonds
can still be considered as playing an important role, even in the case of a
neutral orthophosphate, but they would act via water of hydration.
As research progresses over the years it is becoming apparent that the
majority of these cements are essentially amorphous, and that crystaUinity
is secondary and sometimes very slight. Kingery's arguments based solely
on XRD data are, perhaps, not very relevant.

6.1.4 Effect of cations in phosphoric acid solutions


As we have already shown, the presence of cations in orthophosphoric acid
solution can have a decisive effect on cement formation. As noted above,
Kingery (1950b) found it necessary to modify orthophosphoric acid, by the
addition of calcium, to obtain cement formation with calcium oxide. Also,
Finch and Sharp (1989) had to modify orthophosphoric acid, with either
ammonium or aluminium, to achieve cement formation with magnesium
oxide.
Even when modifiers are not necessary for cement formation, they can
lead to improved cement properties. Kingery (1950b) also examined this
effect. He found that optimum bonding was achieved with cations that had
small ionic radii and were amphoteric or weakly basic, such as beryllium,
aluminium, magnesium and iron. By contrast, cations that were highly
basic and had large ionic radii, for example calcium, thorium and barium,
had a detrimental effect on bonding.
We have noted earlier that aluminium is unusual in forming alumino-
phosphate complexes in phosphoric acid solution which may be of a
polymeric nature. Bearing in mind the analogies between aluminium
phosphate and silica structures, it may well be that during cement
formation an aluminium phosphate hydrogel is formed. Its character may
be analogous to that of silica gel, where a structure is built up by the

203
Phosphate bonded cements

condensation of pairs of hydroxyl groups to form oxygen bridges. Thus,


the structure may consist of P in 4- and Al in 6-coordination, linked by
oxygen bridges with H2O and OH~ as other ligands.

6.1.5 Important cement-formers


The most important of the phosphate bonded cements are the zinc
phosphate, dental silicate and magnesium ammonium phosphate cements.
The first two are used in dentistry and the last as a building material.
Copper(II) oxide forms a good cement, but it is of minor practical value.
In addition, certain phosphate cements have been suggested for use as
controlled release agents. The various phosphate cements are described in
more detail in the remainder of this chapter.

6.2 Zinc phosphate cement


6.2.1 General
Zinc phosphate cement, as its name implies, is composed principally of zinc
and phosphate. It is formed by mixing a powder, which is mainly zinc
oxide, with a solution based on phosphoric acid. However, it is not as
simple chemically as it appears because satisfactory cements cannot be
formed by simply mixing zinc oxide with phosphoric acid solution.
The zinc phosphate cement finds use only in dentistry. Here it is used
mainly as a 'luting agent' for the attachment of inlays, crowns, bridges,
posts and orthodontic bands (Wilson, 1975a,b; Smith, 1982). It is used also
as a cavity liner in crowns and bridges (dental prosthesis). Although new
types of cement have been introduced in dentistry in the 1970s and 1980s,
this traditional cement continues to hold its own, particularly on the
continent of Europe.

6.2.2 History
The early history of the material is obscure. According to Palmer (1891) it
goes back to 1832, but this statement has never been corroborated.
Rostaing (1878) patented a series of pyrophosphate cements which could
include Zn, Mg, Cd, Ba and Ca. Rollins (1879) described a cement formed
from zinc oxide and syrupy phosphoric acid. In the same paper he
mentions zinc phosphate cements recently introduced by Fletcher and
Weston. Similar information is given in a discussion of the Pennsylvania

204
Zinc phosphate cement
Association of Dental Surgeons (1879), where Peirce describes a cement
similar to that of Rollins. Many brands were on the market by 1881
(Miller, 1881).
The earliest formulations, as reported by Rollins (1879), Gaylord (1889),
Ames (1893), Hinkins & Acree (1901) and Fleck (1902), were variously
based on syrupy orthophosphoric acid or unstable mixtures of meta-
phosphoric acid and sodium metaphosphate in solution. Some used solid
pyrophosphoric acid. Many were grossly inferior cements which were
hydrolytically unstable.
Later, better cements appeared based on c. 50 % solutions of ortho-
phosphoric acid. But even these were far from satisfactory. As always with
dental cements, the problems revolved around the control of the setting
reaction: the reaction between zinc oxide and orthophosphoric acid was
found to be far too fierce. By the time of Fleck's 1902 paper these problems
had been solved. The importance of densifying and deactivating the zinc
oxide powder to moderate the cement reaction had been recognized. Of
equal importance was the realization that satisfactory cements could be
produced only if aluminium was incorporated into the orthophosphoric
acid solution. The basic science underlying this empirical finding was
elucidated only in the 1970s.
Comparison of the chemical composition of brands available in the
1960s and 1970s (Axelsson, 1965; Wilson, Abel & Lewis, 1974) shows little
variation from those of the 1930s (Paffenbarger, Sweeney & Isaacs, 1933)
and it is doubtful whether the composition has changed in essence since the
beginning of the century (Table 6.2).

6.2.3 Composition
Powder
The powder is principally composed of zinc oxide (Table 6.2). Magnesium
oxide is found in all current commercial brands in amounts that range from
3 to 10%. Alumina and silica are sometimes to be found. Present day
compositions show less variation than formerly when bismuth, calcium
and barium oxides, or sometimes no additives, were to be found in
commercial examples (Paffenbarger, Sweeney & Isaacs, 1933).
The chief problem with these cements, as with many AB cements, is to
moderate the cement-forming reaction. If the reaction is over-vigorous
then a crystalline mass rather than a cement is formed (Komrska & Satava,
1970; Crisp et al, 1978). Therefore, the zinc oxide used in these cements

205
Phosphate bonded cements

Table 6.2. Chemical composition of commercial zinc phosphate cements


(Axelsson, 1965; Wilson, Abel & Lewis, 1974)

Powders Liquids

Species % by mass Species % by mass


ZnO 89-1-92-7 H 3 PO 4 45-3-63-2
MgO 3-2-9-7 Al 1-0-3-1
A12O3 0-0-6-8 Zn 0-0-9-9
SiO2 0-0-2-1

Based on the results of nine examples.

has to be deactivated and densified by sintering at temperatures which


range from 1000 to 1350 °C to moderate its reaction with aqueous
phosphoric acid solutions. The sintering of a non-stoichiometric solid,
such as zinc oxide, is affected by its initial physical condition and the
surrounding atmosphere. Several processes are involved. Sintering reduces
specific surface area and densities zinc oxide. It also reduces surface energy.
According to Dollimore & Spooner (1971) freshly prepared zinc oxide has
a high surface energy because of its preparation in an oxygen-rich
atmosphere. They suggested that during sintering the initial excess of
oxygen at the surface is reduced by the diffusion of zinc ions from the bulk
to the surface.
Magnesium oxide is always blended with the zinc oxide prior to ignition.
Magnesium oxide promotes densification of the zinc oxide, preserves its
whiteness and renders the sintered powder easier to pulverize (Crowell,
1929). The sintered mixed oxide has been shown to contain zinc oxide and
a solid solution of zinc oxide in magnesium oxide (Zhuravlev, Volfson &
Sheveleva, 1950). Specific surface area is reduced compared with that of
pure zinc oxide and cements prepared from the mixed oxides are stronger
(Crowell, 1929; Zhuravlev, Volfson & Sheveleva, 1950).
In the presence of silica, zinc silicate is formed, the sintering process is
improved and the increase in grain size is enhanced (Zhuravlev, Volfson, &
Sheveleva, 1950). Mineralizers, such as fluorite, cryolite and borax, have a
similar effect (Zhuravlev, Volfson & Sheveleva, 1950). These mineralizers
enhance sintering and promote growth in grain size. As a result the
sintering temperature can be reduced from 1350 °C to 1150-1200 °C.

206
Zinc phosphate cement

Liquid
The liquid is an aqueous solution of phosphoric acid, always containing 1
to 3 % of aluminium, which is essential to the cement-forming reaction
(Table 6.2). Zinc is often found in amounts that range from 0 to 10% to
moderate the reaction. Whereas zinc is present as simple ions, aluminium
forms a series of complexes with phosphoric acid (Section 6.1.1). This has
important consequences, as we shall see, in the cement-forming reaction.

6.2.4 Cement-forming reaction


The cement sets rapidly within a few minutes of preparation. The reaction
is strongly exothermic and is greater than with any other dental cement
(Crisp, Jennings & Wilson, 1978). The excessive heat generated in the
reaction has to be dissipated by progressively incorporating the powder
into the liquid. Strength develops rapidly. About half the ultimate strength
is attained within ten minutes of preparation and 80 % after one hour
(Plant & Wilson, 1970; Williams & Smith, 1971).
The setting reaction is an acid-base one and the course of the reaction is
shown by pH changes in the cement. Two minutes after mixing the pH is
as low as 1-6, after 60 minutes it increases to about 4 and reaches between
6 and 7 after 24 hours (Plant & Tyas, 1970).
The nature of the setting reaction and the set cement remained
imperfectly understood for many years. This is not surprising, for the
products of the reaction depend on a number of factors, including the
phosphoric acid concentration and the presence or absence of aluminium
in the solution. These complexities have caused considerable confusion in
the literature.
Early workers, and some later ones, ignored the fact that aluminium is
always found in the orthophosphoric acid liquid of the practical cement;
its presence profoundly affects the course of the cement-forming reaction.
It affects crystallinity and phase composition, and renders deductions
based on phase diagrams inappropriate. Nevertheless we first describe the
simple reaction between zinc oxide and pure orthophosphoric acid
solution, which was the system studied by the earliest workers.
In the earliest attempt to explain the reaction, Crowell (1929) used, in
part, arguments based on the phase diagrams of the ZnO-P2O5-H2O
system constructed by Eberly, Gross & Crowell (1920). Later, Darvell
(1984) advanced similar arguments using the phase diagrams of Salmon &

207
Phosphate bonded cements

Terrey (1950). In the case of the simple zinc oxide-orthophosphoric acid


system, but only for this simple system, these phase-diagram arguments are
valid.
In the presence of excess zinc oxide, the final product of reaction with
orthophosphoric acid solution is always hopeite, Zn 3 (PO 4 ) 2 . 4H 2 O, and as
early as 1933 Halla & Kutzeilnigg (1933) found that zinc phosphate
cements in service in the mouth contain hopeite. But this substance is not
responsible for initial set in the absence of aluminium. Crowell (1929)
attributed setting to the formation of ZnHPO 4 . 3H2O, which he found
slowly converted over the weeks to hopeite, Zn 3 (PO 4 ) 2 . 4H 2 O. Vieira & De
Arujo (1963) confirmed this result. The most definitive study in this field
was that of Komrska & Satava (1970) who used XRD analysis to identify
the crystalline products formed. They found that a cement prepared from
ZnO and 82-5% H 3 PO 4 set as the result of the formation of
Zn(H 2 PO 4 ) 2 . 2H 2 O crystallites. Under humid conditions at 37 °C the
following conversions occurred:

Zn(H 2 PO 4 ) 2 . 2H 2 O - ZnHPO 4 . H 2 O -> ZnHPO 4 . 3H2O ->


Zn 3 (PO 4 ) 2 .4H 2 O

After 7 days all four species were found in the cement. These conversions
were speeded up when the cement was placed in water, so that after 7 days
only Zn 3 (PO 4 ) 2 . 4H 2 O was found. With 65 % H 3 PO 4 , setting was found to
result from the formation of ZnHPO 4 .H 2 O crystallites. On ageing, the
same conversions occurred as with the 82*5% H 3 PO 4 cement. All these
changes are in accordance with phase-diagram predictions.
Unfortunately, Komrska & Satava (1970) did not examine the reaction
with orthophosphoric acid solutions having concentrations in the range
45 to 6 3 % H 3 PO 4 which, according to Wilson (1975b), is the range
encountered in practical materials. As Darvell (1984) has deduced, when
ZnO is added to orthophosphoric acid solutions within this compositional
range, ZnHPO 4 . 3H 2 O starts to precipitate. As precipitation continues, the
acid concentration declines until it reaches the isoelectric point (22-5 %) at
which ZnHPO 4 . 3H 2 O becomes unstable with respect to the liquid and
starts to convert to Zn 3 (PO 4 ) 2 . 4H 2 O, the final reaction product. Darvell's
deductions for higher initial concentrations of acid are in accordance with
the experimental findings of Komrska & Satava (1970).
These simple zinc oxide-orthophosphoric acid cements are very weak;
indeed, it may be that they are just weakly-bonded aggregates of

208
Zinc phosphate cement
crystallites. Nevertheless, they have been used to advance theories of
cementation. Kingery (1950b) in his XRD investigations of the reaction
reported a crystalline matrix that was not hopeite but was taken to be an
acid phosphate. He was apparently unaware of the slow conversion of the
matrix to this species. Unfortunately, both he and Wygant (1958) went on
to use this and similar observations to construct a theory that attributed
cementation to hydrogen bonding between acid phosphate units. This idea
can now be seen as dubious. The role of hydrogen bonding in the cement
matrix is to be envisaged between particles of colloidal dimensions rather
than between molecular units.
For many years the picture remained of a cement consisting of zinc oxide
particles bonded by a crystalline matrix of hopeite (Dobrowsky, 1942). But
these ideas were erroneous for several reasons. Phase-diagram arguments
applied only to systems in thermodynamic equilibrium and, clearly, these
conditions are not obtained in rapidly setting cements. XRD analysis can
be misleading as it ignores amorphous phases. Finally, these early ideas
ignored the role of aluminium, although the necessity of having it in the
liquid for the production of satisfactory cements had long been known
(Section 6.2.2).
Komrska & Satava (1970) showed that these accounts apply only to the
reaction between pure zinc oxide and phosphoric acid. They found that the
setting reaction was profoundly modified by the presence of aluminium
ions. Crystallite formation was inhibited and the cement set to an
amorphous mass. Only later (7 to 14 days) did XRD analysis reveal that
the mass had crystallized directly to hopeite. Servais & Cartz (1971) and
Cartz, Servais & Rossi (1972) confirmed the importance of aluminium. In
its absence they found that the reaction produced a mass of hopeite
crystallites with little mechanical strength. In its presence an amorphous
matrix was formed. The amorphous matrix was stable, it did not crystallize
in the bulk and hopeite crystals only grew from its surface under moist
conditions. Thus, the picture grew of a surface matrix with some tendency
for surface crystallization.
Surprisingly, the effect of aluminium on the reaction had been
anticipated by van Dalen many years before in his thesis of 1933, but had
not made its way into the scientific literature. The authors are indebted to
Dr L. J. Pluim of the Rijksuniversiteit te Groningen for this information.
Van Dalen (1933) was convinced that aluminium had an important role
in cement formation and that Crowell had been wrong to ignore it. Van
Dalen found that the reaction between zinc oxide and phosphoric acid was

209
Phosphate bonded cements

greatly moderated by the presence of aluminium in the liquid. He attributed


this to the formation of a gelatinous coating of aluminium phosphate
around each zinc oxide particle. This observation has never been repeated,
but it appears entirely reasonable. Van Dalen also found that aluminium
inhibited the formation of crystallites and that this inhibition increased
with increase in the aluminium content of the liquid.
Crisp et al. (1978) were able to follow the course of the cement-forming
reaction using infrared spectroscopy and to confirm previous observations.
They found that the technique could be used to distinguish between
crystalline and amorphous phases of the cement. Hopeite shows a number
of bands between 1105 and 1000 cm"1; this multiplicity has been explained
by postulating a distortion of the tetrahedral orthophosphate anion. (Two-
thirds of the zinc ions are tetrahedrally coordinated to four phosphate ions,
and the remainder are octahedrally coordinated to two phosphate and four
water ligands.)
Crisp et al. (1978) were able to detect the formation of crystallites both
on the surface and in the bulk of the reaction product. In the absence of
aluminium the reaction between zinc oxide and phosphoric acid was very
rapid and the cement set in less than two minutes. Hopeite was formed,
within minutes, both at the surface and in the bulk of the reaction mass. It
was doubted whether this mass constituted a true cement.
The addition of aluminium to the liquid slowed down the reaction. An
amorphous cement was formed and there was no crystallization in the bulk
of the cement. However, after some time crystallites were formed at the
surface. Thus, the presence of aluminium exerts a decisive influence on the
course of the cement-forming reaction. This effect is to be attributed to the
formation of aluminophosphate complexes (see Sections 6.1.2 and 4.1.1).
These complexes may delay the precipitation of zinc from solution and also
introduce an element of disorder into the structure, thus inhibiting
crystallization. It is significant that zinc, which does not form complexes,
has little effect on the nature or speed of the reaction.
Crisp and coworkers found that the development of surface crystallinity
was related to the speed of set. The faster the reaction, the shorter was the
inhibition period before surface crystallization took place. When the
setting time of a cement was between two and three minutes, surface
crystallinity developed in a few minutes. When it was seven minutes,
surface crystallinity was delayed by three hours. The reaction rate was
affected by the chemical composition and physical state of the cement
components. Well-ignited zinc oxide, the presence of magnesium in the

210
Zinc phosphate cement
oxide powder, high phosphoric acid concentration, a low powder/liquid
ratio and the presence of aluminium in the liquid all served to retard the
reaction. It is notable that these workers, like Servais & Cartz, found no
evidence for the formation of crystalline acid phosphates in the cements.
More recently, Steinke et al. (1988) examined four commercial cements
and found that the matrices were mainly amorphous; indeed, they found
hopeite in only one of them. This indicates that manufacturers today are
able to formulate to prevent crystallization.
Wilson, Paddon & Crisp (1979) have shown that the water present in the
cement can be divided, somewhat arbitrarily, into bound water of
hydration (non-evaporable) and loosely held (evaporable) water. The
amount of tightly bound water increases as the cement ages and in one
example reached 42 % of the total water.
It is interesting that this cement has been known for over 100 years and
yet certain features of its chemistry remain obscure. The exact nature of the
matrix is still a matter for conjecture. It is known that the principal phase
is amorphous, as a result of the presence of aluminium in the liquid. It is
also known that after a lapse of time, crystallites sometimes form on the
surface of the cement. A cement gel may be likened to a glass and this
process of crystallization could be likened to the devitrification of a glass.
Therefore, it is reasonable to suppose that the gel matrix is a zinc
aluminophosphate and that entry of aluminium into the zinc phosphate
matrix causes disorder and prevents crystallization. It is not so easy to
accept the alternative explanation that there are two amorphous phases,
one of aluminium phosphate and the other of zinc phosphate. This is
because it is difficult to see how aluminium could act in this case to prevent
zinc phosphate from crystallizing.

Summary of experimental evidence


A summary of this evidence may be attempted to give a probable reaction
mechanism. After mixing, the zinc oxide powder is attacked by the acid
solution, water acting as the reaction medium. Zinc ions are extracted and
the pH at the powder-liquid interface rises, causing aluminium phosphate
or more probably a zinc aluminophosphate to precipitate as a gel at the
particle surface. This gel coating moderates the reaction. Zinc ions diffuse
through this layer and, as the pH rises, precipitate as an amorphous gel,
probably a zinc aluminium phosphate. This reaction mechanism thus
postulates both a topochemical and a through-solution reaction. As
reaction proceeds, the cement matrix becomes increasingly hydrated.

211
Phosphate bonded cements

Increase in concentration of aluminium and phosphoric acid in the liquid


serves to slow the reaction. This observation is in line with the above
reaction scheme. Increase in the aluminium content will serve to increase
the thickness of the coating formed around zinc oxide particles. Increase
in phosphoric acid content implies a decrease in water content and an
impairment of the hydration reaction.
The phenomenon of surface crystallization could be represented by the
equation.

H2O

Zinc aluminophosphate -> Zn3PO4. 4H2O + A1PO4. «H2O


amorphous hydrogel hopeite amorphous gel

6.2.5 Structure
The set cement is completely opaque and can be regarded as a composite
of unconsumed zinc oxide particles, possibly coated with an aluminium
orthophosphate gel, bonded together by an amorphous neutral zinc
orthophosphate gel. There is some tendency for hopeite crystallites to grow
from the surface (Servais & Cartz, 1971; Cartz, Servais & Rossi, 1972;
Crisp et al, 1978). Growth is related to the speed of set and the presence of
moisture. Under dry conditions the surface is stable and undulating with
no sign of crystallites (Figure 6.3a). When the environment is maintained
at 100% relative humidity, crystal growth is observed (Figure 63b) and
may be compared with the devitrification of a glass. Servais & Cartz (1971)
observed that under certain conditions the layer could be 8 |am thick. It is
only loosely attached to the body of the cement matrix. The presence of
this layer of crystallites may explain why the cement lacks adhesion.
The bulk of the cement is extremely porous as the fractured surface of a
specimen shows (Figure 6.3c). The pores are 0-5 \im in diameter and more
abundant in the depth of the cement. The porosity arises from excess
unbound water which separates out as globules in the cement and is
trapped by the rapid setting. Subsequent diffusion of these globules leaves
the cement porous. This makes the cement permeable to dyes (Wisth,
1972).
As mentioned previously, the cement contains both tightly bound and
loosely bound water. The set cement can both lose and gain water
depending on its environment. Under drying conditions (say 50 % relative
humidity) it loses water and shrinks. When placed in water there is an

212
Zinc phosphate cement

(b)
Figure 6.3 The effect of environmental conditions on the surface of a zinc phosphate cement:
(a) stable and undulating surface with no sign of crystallites observed under dry conditions,
(b) crystal growth observed in an atmosphere of 100 % relative humidity, (c) extreme porosity
observed in the bulk of the cement; pores are 0-5 urn in diameter (Servais & Cartz, 1971).

213
Phosphate bonded cements

Table 6.3. Specification properties of commercial zinc phosphate cements


{Wilson 1975b)

Value Specification limits"

Powder: liquid6 20-30 —


Setting time (37 °C), min 3-9-7-5 5-9
Film thickness, urn 24-40 25 maximum
Compressivec 70-131 70 minimum
strength (24 h), MPa
Solubility & 004-3-3 0-2 maximum
disintegration (24 h), %

a
BS 3364: 1961 Specification for Dental Zinc Phosphate Cement.
b
For a consistency spread of 30 mm diameter for 0-5 cm3 of cement paste under
a load of 1-96 N (200 g weight) applied after 3 minutes at 23 °C.
c
After storage for 24 hours in water at 37 °C, based on the results of nine
examples.

equilibrium water uptake: 9 % for two zinc phosphate cements examined


by Eichner, Lautenschlager & von Radnoth (1968).

6.2.6 Properties
Zinc phosphate cement is used as a luting agent for the cementing of
crowns and bridges in dentistry. It is not an ideal material. It does not
possess the adhesion of modern polyelectrolyte cements, but its clinical
performance has been good enough to ensure it a place in dentistry for
about 100 years. Despite the challenge of modern polyelectrolyte cements
it still holds its own. Undoubtedly this is because it is easy to mix, the fluid
paste is easy to manipulate, its working time is long and the paste sets
sharply to a hard mass. These are important properties for a dental
material. In practice it means that these cements are easier to use than the
polyelectrolyte cements and this alone explains their continued popularity
with the dentist.

Preparation and setting


Zinc phosphate cement is prepared by introducing small incremental
amounts of powder into the liquid and mixing the paste over a large area
on a glass slab in order to dissipate heat because of the excessive exotherm

214
Zinc phosphate cement

Table 6.4. Mechanical properties of commercial zinc phosphate


cements (Housten & Miller, 1968; Wilson, 1975b; Wilson &
Lewis, 1980; Powers, Farah & Craig, 1976; 0ilo & Espevik,
1978)

Property" Value

Compressive strength 70-131 MPa


Compressive modulus 11-9—13-5 GPa
Tensile strength 4-3-8-3 MPa
Strain 0-2 %b
Creep (over 24 hours) 0-13 %b

a
After storage for 24 hours in water at 37 °C.
b
Only one example.

of this cement (Crisp, Jennings & Wilson, 1978). To achieve a consistency


suitable for cementing crowns this cement is normally mixed with a
powder/liquid ratio that ranges from 2-3 to 2-7 g cm" 3 (Table 6.3) (Wilson,
1975b). The working time (at 23 °C) then varies from 3 to 6 minutes and
setting time from 5 to 14 minutes (Plant, Jones & Wilson, 1972; Jendresen,
1973; Wilson, 1975b; Myers, Drake&Brantley, 1978; E a r n e s t a l , 1977).
The linear contraction of the cement on setting (0-5 %) can give rise to slits
at the cement-tooth and cement-restoration interfaces (0ilo, 1978).
Zinc phosphate cement mixes to a paste which is thin and mobile. Under
pressure it flows readily to give a film 24 to 40 jim thick (Table 6.3). This
film thickness is adequate to seat restorations, especially as McLean & von
Fraunhofer (1971) and Dimashkieh, Davies & von Fraunhofer (1974) have
shown that in practice the gap between tooth and restoration can be as
much as 100 |iim or more.

Mechanical properties
Fully hardened cements have brittle characteristics (Williams & Smith,
1971; Skibell & Shannon, 1973) and show little creep under load (Wilson
& Lewis, 1980). When mixed to a luting (cementation) consistency, their
compressive strength reaches 70 to 131 MPa after 24 hours (Wilson,
1975b) depending on brand (Table 6.4). There is little subsequent increase
in strength (Paffenbarger, Sweeney & Isaacs, 1933; Smith, 1977).
The tensile strength of these cements lies between 4-3 and 8-3 MPa and
is thus lower than their compressive strength (Table 6.4) (Williams &

215
Phosphate bonded cements

Smith, 1971; Hannah & Smith, 1971; Powers, Farah & Craig, 1976). The
modulus of elasticity in compression is 12 to 13 GPa (Table 6.4) (Powers,
Farah & Craig, 1976; Wilson, Paddon & Crisp, 1979).

Erosion
Good cements show little dissolution in water, less than 0-1 % using the
standard test (see Chapter 10); the amount can be much greater for poorer
examples. Dissolution represents the amount of material eluted from a
one-hour-old cement as it ages in water for a further 24 hours. As the
cement ages further the rate of dissolution falls although it always remains
significant (Wilson, Kent & Lewis, 1970; Wilson, 1976; de Freitas, 1973).
Wilson, Abel & Lewis (1974), in a detailed chemical study of erosion in
aqueous solution, found that in the first 24 hours of the cement's life the
ions eluted were Zn2+, Mg2+, HPO2" and H2PC>4. Far more Mg2+ ions were
eluted than Zn2+ ions, despite zinc being the major metal constituent of the
zinc phosphate cement. These workers deduced that magnesium is far less
firmly bound to phosphate than is zinc and that, consequently, its presence
in the oxide is a source of weakness. These results were later confirmed by
Anzai et al (1977).
Wilson, Kent & Lewis (1970) in a long-term study found that most of
the soluble phosphate was eluted from the cement in the first 24 hours. By
contrast the rate of elution of zinc remained constant for 160 days, the
length of the study. They concluded that long-term erosion took place at
the surface of the oxide particles rather than in the matrix.
Of more significance, however, were their results for the effect of pH.
These showed that the cement is at its most stable in neutral solution and
that dissolution increases sharply with increasing acidity (Figure 6.4). This
is of clinical significance, for pHs as low as 4 can occur in the stagnation
regions of the mouth (Stephan, 1940; Kleinburg, 1961), and it is now
generally believed that the life of a dental cement is determined by its
resistance to acid conditions. The dissolution in lactic acid and especially
citric acid is much higher (Norman, Swartz & Phillips, 1957). This effect
must be attributed to the complexing effect of these acids.
The laboratory impinging jet test for evaluating the acid erosion of
dental cements is described in Chapter 10. Using this method with lactic
acid-lactate solutions, Wilson et al. (1986b) found, for one cement, that the
erosion rate was virtually zero at pH = 5-0, 0-38 % at pH = 4-0 and 5-7 %
at pH = 2-7. For a range of cements Wilson et al. (1986a) found erosion
rates varying from 3-0 to 5-7 % in lactic acid solutions of pH = 2-7. The

216
Zinc phosphate cement
zinc phosphate cement is markedly less resistant to acid erosion than the
aluminosilicate glass cements, glass-ionomer cements and dental silicate
cements. They also found that, with one exception, zinc phosphate cements
were somewhat more resistant to acid erosion than zinc polycarboxylate
cement. These results have been confirmed by other workers using similar
methods (Beech & Bandyopadhyay, 1983; Kuhn, Setchell & Teo, 1984;
Gulabivala, Setchell & Davies, 1987; Mesu, 1982).
In vivo studies have indicated that zinc phosphate cements erode under
oral conditions. Also, cements based on zinc oxide, including the zinc
phosphate cement, are less durable in the mouth than those based on
aluminosilicate glasses, the dental silicate and glass-ionomer (Norman et
al., 1969; Ritcher & Ueno, 1975; Mitchem & Gronas, 1978,1981; Osborne
et al, 1978; Pluim & Arends, 1981, 1987; Sidler & Strub, 1983; Mesu &
Reedijk, 1983; Theuniers, 1984; Pluim et al, 1984, Arends & Havinga,
1985). However, there is some disagreement on whether the zinc phosphate
cement is more durable than the zinc polycarboxylate cement.
Dissolution of the cement has been associated with increased marginal

\P/L26q/ml

silicate
cement

P/L >
2", 4 2q/ml

VJ
\ \

I I > ~r
pH
Figure 6.4 Effect of pH on the elution of phosphate from a zinc phosphate cement mixed at
two different consistencies (Wilson, Kent & Lewis, 1970).

217
Phosphate bonded cements
leakage (Andrews & Hembree, 1976) and the penetration of bacteria
(Brannstrom & Nyborg, 1974).

6.2.7 Factors affecting properties


Cement properties are affected by a number of factors. Some are
determined by the manufacturer, for example the chemical composition of
the cement components. Others are under the clinician's control. These
include the powder/liquid ratio of the cement mix and the temperature of
the surgery. Increase in either of these variables accelerates the reaction
and affects properties.
Properties are also affected by the composition of the phosphoric acid
liquid. We have already pointed out that the liquid has to contain
aluminium for satisfactory cement formation. In addition the H 3PO4/H2O
ratio exerts a considerable effect on cement properties. This was investi-
gated by Worner & Docking (1958). If the cement liquid is exposed to air,
water can be readily gained or lost, depending on humidity, and this clearly
affects the H3PO4/H2O ratio and therefore the cement properties (Figure
6.5). If water is lost, then setting time is prolonged, strength increased and

3O 4O
°h water in the liquid
Figure 6.5 Effect of water content of the liquid (H 2 O:H 3 PO 4 ) on the properties of a zinc
phosphate cement (Worner & Docking, 1958).

218
Zinc phosphate cement
resistance to aqueous attack decreased. The converse also occurs, although
the reduction in setting time is slight.
Both setting time and working time are reduced if the mixing
temperature is increased. Longer mixing time gives slower set (Paffen-
barger, Sweeney & Isaacs, 1933). As long ago as 1892, Evans advocated the
use of a cool slab for cement preparation. Low-temperature preparation
extends working time and enables more powder to be incorporated into the
cement; this is advantageous for it increases strength and resistance to
dissolution, although this varies from brand to brand. This effect has since
been confirmed by many workers: Paffenbarger, Sweeney & Isaacs (1933);
Henschel (1943); Jendresen (1973); Myers, Drake & Brantley (1978);
Windeler (1978); Kendzior, Leinfelder & Hershey (1976); Tuenge, Sugel &
Izutsu (1978); Williams et al. (1979). However, there is a danger with this
technique, that if the atmosphere is excessively moist, water may condense
on the slab and the cement be weakened (Tuenge, Sugel & Izutsu, 1978;
Norman et al., 1970).
The powder/liquid ratio used in the cement mix affects a number of
properties. As it is increased, setting time and working time are reduced.
Compressive strength increases almost linearly with powder/liquid ratio
(Savignac, Fairhurst & Ryge, 1965).
Film thickness is controlled by a number of factors. The grain size of the
powder imposes a lower limit on its value and rheological characteristics of
the cement affect flow (Jorgensen & Peterson, 1963). An increase in the
powder/liquid ratio or a delay in seating a restoration leads to an increase
infilmthickness. The geometry of the surfaces to be cemented also affects
flow and hence film thickness (Windeler, 1979).

6.2.8 Biological effects


The zinc phosphate cement is not bland towards living tissues. When
freshly mixed it is highly acidic with a pH as low as 1-6 (Plant & Tyas,
1970). Even after it has aged one hour the pH may be lower than 4. This can
give rise to pulpal irritation and pain. Prolonged pulpal irritation in deep
cavities cannot be allowed and some form of pulpal protection is needed in
these cases.
There are other causes of pain and pulpal irritation. Hydraulic pressure
developed during the seating of a restoration can lead to pulpal damage
(Hoard et al., 1978). The movement of fluid under osmotic pressure has
been cited as a cause of pain (Brannstrom & Astrom, 1972).

219
Phosphate bonded cements
If the cement is mixed too thinly it may etch the tooth enamel because of
its excess acidity (Docking et al., 1953; Abramovich, Macchi & Ribas,
1976). Of course, etching can promote mechanical attachment to the tooth
(Ware, 1971).

6.2.9 Modified zinc phosphate cements


Fluoride is found in some zinc phosphate cements, generally as stannous
fluoride. The cements are weaker and have less resistance to dissolution
than normal zinc phosphate cements (Myers, Drake & Brantley, 1978;
Williams et al., 1979). They releasefluorideover a long period (de Freitas,
1973) and this is taken up by enamel (Wei & Sierk, 1971). This results in
reduced enamel solubility (Gursin 1965; Skibell & Shannon, 1973) and
increased hardness (Yamano, 1968). Fluoride-releasing cements should
reduce the incidence of enamel decalcification under orthodontic bands
but this effect has not been recorded.
Althoughfluorideis added as the tin salt,fluoriderelease is accompanied
by the release of aluminium and not tin (de Freitas, 1973). There is little
leaching out of tin apart from an initial wash-out. Of course, aluminium is
not released from the normal cement (Wilson, 1976; Wilson, Abel & Lewis,
1974).
These cements are of very minor interest in dentistry.

6.2.10 Hydrophosphate cements


Interesting attempts have been made to formulate water-setting cements
by blending solid acid phosphates with the zinc oxide powder. The cement
is then prepared by mixing this powder blend with water. These attempts
may be considered to have failed.
Nakazawa et al. (1965) used calcium or magnesium dihydrogen
phosphate as the acid phosphate. The powders were hygroscopic. The
magnesium salt yielded cements which showed excessive contraction on
setting and were weak and prone to aqueous attack. The calcium salt was
somewhat better.
A more successful approach was that of Higashi et al. (1969a,b 1972).
They blended solid acid phosphate salts with zinc oxide powder. One acid
salt used was a precipitated hydrate of ZnH2PO4. The cement was formed
by mixing this powder blend with water. Work progressed to the point
where three commercial brands of these so-called 'hydrophosphate'
cements appeared on the market. None met the specification requirements

220
Transition-metal phosphate cements
(Laswell et ai, 1971; Arato, 1974). All were prone to excessive dissolution
and only one had adequate strength and film thickness. Their working
characteristics were found to be unduly sensitive to changes in temperature
and humidity (Simmons, D'Anton & Hudson, 1968). All were inferior to
conventional zinc phosphate cements. No further development of these
cements has taken place, nor is it likely that interest in them will be revived.
The modern water-activated glass-ionomer cement has filled this niche
and has vastly superior properties including adhesion to tooth material.

6.3 Transition-metal phosphate cements


Copper oxides are good cement-formers but copper phosphate cements
find little practical use. They have a very minor use in dentistry, the last
description of their chemistry appearing in 1940 (Worner, 1940). They
have also been proposed for the controlled release of copper as a trace
element supplement. They were introduced into dentistry by Ames in 1892
and came into extensive use between 1914 and 1916; Poetschke described
them in 1916. The rationale for their use was based on the germicidal
action of copper. There appears to be some basis for this belief (Worner,
1940; Babin, Hurst & Feary, 1978). They are, however, excessively acidic
and cause pulpal irritation (Worner, 1940; Ware, 1971). At one time they
were used for thefillingof pits andfissuresof deciduous (children's) teeth,
lining, restoring badly decayed teeth, luting and the placement of
orthodontic bands. By 1971 their use was confined to the last two
applications (Ware, 1971).
There is virtually no knowledge of the setting and structure of copper
phosphate cements. Mostly, they are complex materials. The simplest was
based on a powder containing 91-5% CuO and 8-4% Co3O4. Others
contained respectively 62-2 % CuO and 29-8 % ZnO, and 23-9 % Cu2O and
66-7 % ZnO, with other metal oxides. The strength of these cements is
about the same as the zinc phosphate cement (Ware, 1971). There are also
pseudo-copper cements, which are zinc phosphate cements coloured by
minor amounts of copper(II) oxide.
Vashkevitch & Sychev (1982) have identified the main reaction product
of the cement-forming reaction between copper(II) oxide and phosphoric
acid as Cu3(PO4)2. 3H2O. The addition of polymers - poly(vinyl acetate)
and latex - was found to inhibit the reaction and to reduce the compressive
strength of these cements. However, impact strength and water resistance
were improved.

221
Phosphate bonded cements

More recently copper phosphate cements have been suggested for use as
controlled-release agents for supplying trace amounts of copper to cattle
and sheep over an extended period (Allen et al, 1984; Manston et al., 1985;
Prosser et al., 1986). The cements were prepared with a Cu/P ratio of 1:1
to ensure that the matrix was an acid phosphate and so subject to
dissolution in aqueous solutions. They released copper at a constant rate
for 90 days.
Here we might note that cobalt(II) hydroxide, but not the oxide, also
forms cements (Allen et al., 1984; Manston & Gleed, 1985; Prosser et al.,
1986). It also is used in controlled-release devices for supplying trace
elements to cattle and sheep. Nothing is known of its structure.

6.4 Magnesium phosphate cements


6.4.1 General
Cements based on the reaction between magnesium oxide and phosphates
have long been known as investment materials for the casting of alloys
(Prosen, 1939, 1941; Earnshaw, 1960a,b). More recently versions have
been used as materials for the emergency repair of roads, runways and
industrial floors (El-Jazairi, 1982). This is because of their fast-setting
properties and ability to set at low temperatures. Setting reactions have
only been elucidated in the last decade or so. The important point to note
is that although a cementitious mass is formed when magnesium oxide is
reacted with phosphoric acid, it is soluble in water (Finch & Sharp, 1989).
Practical cements are formed only when ammonium or aluminium
phosphates are incorporated into the cement liquid. Interestingly, they
appear to occur as a natural cementitious binder in kidney stones
(Abdelrazig & Sharp, 1988). Unlike other phosphate-bonded cements that
have been fully studied, these appear to have a mainly crystalline matrix.

6.4.2 Composition
Magnesium (or magnesia) phosphate cements are based on the reaction
between ignited magnesium oxide and acid phosphates, which are generally
modified by the addition of ammonium and aluminium salts. The
phosphates may be either in solution or blended in solid form with the
magnesium oxide. In the latter form the cement is formed by mixing the
powder blend with water.

222
Magnesium phosphate cements

Table 6.5. Crystalline phosphate species found in magnesium


phosphate cements

Crystalline species Formula


Dittmarite MgNH4PO4. H2O
' Hayesite' MgHPO4. 2H2O or MgHPO4. H2O ?
Newberyite MgHPO4. 3H2O
Phosphorresslerite MgHPO4. 7H2O
Schertelite Mg(NH4)2(HPO4)2. 4H2O
Stercorite NaNH4HPO4. 4H2O
Struvite MgNH4PO4. 6H2O

The most important characteristic of the magnesium oxide powder used


in these cements is its reactivity (Glasson, 1963). Magnesium oxide needs
to be calcined to reduce this, otherwise the cement pastes are too reactive
to allow for placement. Surface area and crystal size are important and
relate to the calcination temperature (Eubank, 1951; Harper, 1967; Sorrell
& Armstrong, 1976; Matkovic et ai, 1977). The lower reactivity of
calcined magnesium oxide relates to a lower surface area and a larger
crystallite size.

6.4.3 Types

There are several types of cement and mortar (cement plus filler):
(1) Those based on the reaction between magnesium oxide powder
and orthophosphoric acid solution.
(2) Those based on the reaction between magnesium oxide and
ammonium dihydrogen phosphate (ADP), often in the presence
of sodium tripolyphosphate (STPP), which is mixed with water
(El-Jazairi, 1982; Abdelrazig et aL, 1984; Abdelrazig, Sharp & E1-
Jazairi, 1988, 1989).
(3) Those based on the reaction between magnesium oxide and
diammonium hydrogen phosphate (DAP) (Sugama & Kukacka,
1983a).
(4) Those based on the reaction between magnesium oxide and
ammonium polyphosphate (APP) (Sugama & Kukacka, 1983b).

223
Phosphate bonded cements

(5) Those based on the reaction between magnesium oxide and


aluminium hydrogen phosphate solutions (Finch & Sharp, 1989).

6.4.4 Cement formation and properties


Cement formation between MgO and various acid phosphates involves
both acid-base and hydration reactions. The reaction products can be
either crystalline or amorphous; some crystalline species are shown in
Table 6.5. The presence of ammonium or aluminium ions exerts a decisive
influence on the course of the cement-forming reaction.

6.4.5 Cement formation with phosphoric acid


The reaction between magnesium oxide and 85-3 % phosphoric acid has
been studied by Finch & Sharp (1989). The reaction products were found
to be two highly crystalline phases, one unidentified, and an amorphous
phase. One crystalline phase was most probably Mg(H 2 PO 4 ) 2 . 2H 2 O.
The other was believed to be another dihydrogen phosphate,
Mg(H 2 PO 4 ) 2 .4H 2 O. These findings are in accordance with the phase
diagrams of Belopolsky, Shpunt & Shulgina (1950) who reported both
these species.
These cements are soluble in water and so of no practical significance.

6.4.6 Cement formation with ammonium dihydrogen phosphate


The reaction between MgO and ammonium dihydrogen phosphate (ADP)
in aqueous solution yields struvite, MgNH 4 PO 4 . 6H 2 O, and schertelite,
Mg(NH 4 ) 2 (HPO 4 ) 2 . 4H 2 O, as the main reaction products. Both may be
regarded as cementing species. Only minor amounts of MgO are consumed
during these reactions as it is present in excess. The reaction is exothermic.
Sodium tripolyphosphate (STPP) or borax may be added to retard the
reaction. The main course of the reactions may be represented thus:
MgO + 2NH£ + 2H 2 PO 4 + 3H 2 O = Mg(NH 4 ) 2 (HPO 4 ) 2 . 4H 2 O [6.2]
Mg(NH 4 ) 2 (HPO 4 ) 2 . 4H 2 O + MgO + 7H 2 O = 2MgNH 4 PO 4 . 6H 2 O [6.3]
The reaction will not go to completion unless there is sufficient water.
Mortars of this system are prepared by blending ignited magnesium
oxide, ADP and STPP with a filler, normally quartz sand. On mixing with
water a cementitious mass is formed. The reaction has been studied by a
number of workers: Kato et al (1976), Takeda et al. (1979), Neiman &

224
Magnesium phosphate cements
Sarma (1980), Abdelrazig et al (1984), Popovics, Rajendran & Penko
(1987), Abdelrazig, Sharp & El-Jazairi (1988, 1989). The systems used
differed in the proportions of MgO: ADP: H2O which did affect some
aspects of the reaction. Nevertheless there are common features to all
compositions and some broad conclusions can be drawn.
On mixing, an exothermic reaction takes place with some loss of
ammonia. As the reaction proceeds crystalline phases are formed. All
workers are in agreement that these are the tetrahydrate schertelite, or the
hexahydrate struvite, or a mixture of both. They also agree that schertelite
isfirstformed, which is then hydrated further, if water is available, to form
struvite. The relative amounts thus depend on the water available for
hydration. Sometimes dittmarite, MgNH4PO4. H2O, and stercorite,
NaNH4HPO4. 4H2O, are found in minor amounts.
The most extensive studies on these materials are those of Abdelrazig
and coworkers, who used XRD, differential thermal analysis, thermo-
gravimetry and scanning electron microscopy to characterize the reaction
products. The reaction has been described by Abdelrazig et al. (1984), with
subsequent revision (Abdelrazig, Sharp & El-Jazairi, 1988, 1989). On
adding water to ADP a solution is immediately formed that contains
ammonium and various complex phosphate ions. This solution reacts with
the suspended magnesium oxide particles. Hydration products are formed
probably by a through-solution mechanism, an idea originally advanced
by Neiman & Sarma (1980), rather than by a topochemical reaction.
Initially, when the concentration of phosphate is high, the tetrahydrate
schertelite (MgO:PO4 = 1:2) is formed. Schertelite can be seen as a
reaction intermediate which reacts further with MgO. As reaction
proceeds, the phosphate content of the solution drops and the formation of
struvite or dittmarite (MgO:PO4 = 1:1) is favoured. The reaction can be
frozen by lack of either water or phosphate. If there is an abundance of
water then the formation of the hexahydrate struvite is favoured. If,
however, water is lacking then some monohydrate dittmarite is formed,
and moreover the reaction does not proceed. Schertelite remains in the set
material and some ADP remains undissolved. Thus, the reaction is not one
of progressive hydration from monohydrate to tetrahydrate to hexa-
hydrate.
The presence of colloidal species has been a subject of some debate.
Neiman & Sarma (1980) found with their material that crystallinity did not
develop for at least the first two hours of setting. However, Abdelrazig
et al (1984) and Abdelrazig, Sharp & El-Jazairi (1988, 1989) reported that

225
Phosphate bonded cements
abundant amounts of crystallites are formed in only a few minutes. These
differences probably relate to the use of different molar ratios of reactants,
although this cannot be confirmed because Neiman & Sarma did not give
the molar ratios that they used.

(3) 100 ym (b) 1 0 0 ]im

(c) 1 0 pm (d) 100


For legend see opposite

226
Magnesium phosphate cements
Abdelrazig, Sharp & El-Jazairi (1988, 1989) prepared a series of
mortars based on a powder blend of MgO and ADP with a quartz
sand filler. They were hydrated by mixing with water. A mortar I
(MgO: ADP: silica: water = 17-1:12-9:70-0:12-5), with a water/solid ratio
of 1:8, formed a workable paste which set in 7 minutes with evolution of
ammonia. The main hydration product, struvite, was formed in ap-
preciable amounts within 5 minutes and continued to increase. Schertelite
also appeared, but only in minor amounts, within the first 5 minutes and
persisted only during thefirsthour of the reaction. Dittmarite appeared in
minor amounts after 15 minutes, and persisted.
Microstructural changes were observed during hydration. After the first
hour the microstructure appeared to be predominantly crystalline,
although the crystallinity was poor. The crystallites appeared to grow
between silica grains (Figure 6.6a). After seven hours the crystallinity
became well defined (Figure 6.6b); the platy and tabular morphologies
observed were tentatively assigned to dittmarite and struvite respectively
(Figure 6.6c). After 24 hours there was an abundance of struvite crystallites
with a rod-like appearance (Figure 6.6d,e).
The compressive strength of the mortar reached 14-9 MPa in 24 hours.

V
(e) 10 \im
Figure 6.6 Scanning electron micrographs of magnesium ammonium phosphate mortar I
(Abdelrazig, Sharp & El-Jazairi, 1989) after hydration at 22 °C: (a) after 1 hour, low
magnification, (b) after 7 hours, low magnification, (c) after 7 hours, high magnification
showing platy and tabular morphologies, (d) after 24 hours, low magnification, (e) after 24
hours, high magnification showing rod-like crystallites of struvite.

227
Phosphate bonded cements
Total pore (3-75 nm to 7-5 (am) volume was 73-1 mm3 g"1 and coarse pore
(1 |im to 7-5 |im) volume was 60-1 mm3 g"1.
The addition of STPP (1-7%) acted as a retarder and increased
compressive strength (mortar II). Less heat and ammonia were evolved
and the cement set more slowly in 10 minutes. The paste hardened in 30 to
60 minutes. Traces of ADP persisted for 30 minutes but no STPP was
detected in the reaction products. Struvite, the main hydration product,
schertelite and dittmarite all appeared within 5 minutes. Struvite continued
to increase in amount as the cement aged; schertelite disappeared after 3
hours and dittmarite after a week. Stercorite was found only during the
first 7 hours.
The presence of STPP affected the morphology of the cement. After 24
hours struvite with ellipsoidal morphology was observed (Figure 6.1a).
Later (7 and 28 days) there was a morphological change with the formation
of well-crystalline hexagonal plates (Figure 6.1b). This morphology is quite
different from that of mortar I.
The addition of STPP improved the compressive strength of the mortar
which reached 19-5 MPa in 24 hours. The total pore volume was reduced
to 70-4 mm3 g"1 and the coarse pore volume to 55-4 mm3 g"1.

•J
10 10 ym
Figure 6.7 Scanning electron micrographs of magnesium ammonium phosphate mortar II
(Abdelrazig, Sharp & El-Jazairi, 1989): (a) after 24 hours, high magnification showing
struvite of ellipsoidal morphology, (b) after 28 days, high magnification showing well-
crystalline hexagonal plates of struvite.

228
Magnesium phosphate cements

When a mortar was prepared with insufficient water (mortar III,


water: solid = 1:16) some ADP remained even after 3 weeks and the
mortar took longer to set (12 minutes). Schertelite was the major reaction
product for the first hour; thereafter struvite became the major product,
but schertelite remained a significant constituent. The microstructure
differed from that of the other two mortars. After one hour there was a
dense crystalline microstructure with needle-like and cuboid crystallites
growing between silica grains (Figure 6.&a,b). These crystallites were
tentatively assigned to struvite and schertelite respectively. These features
did not change with time. The compressive strength of the mortar was
27-4 MPa, the total pore volume was reduced considerably to 20-6 mm3 g"1
and the coarse pore volume to 5-3 mm3 g"1. This incompletely hydrated
mortar was stronger than the other two fully hydrated mortars. In this,
these materials obey the general rule that strength of a cement is increased
as the water content is reduced, irrespective of the exact microstructure of
the matrix.
Abdelrazig et al. (1984) studied the commercial FEB SET-45 cements
and mortars (Set Products Inc., Master Builders Division, Martin Marietta
Corporation). Their hydration behaviour is similar to those described
above. The mortars normally set in 15 minutes and hardened in 30 to 60

100 \m 10 ym
Figure 6.8 Scanning electron micrographs of magnesium ammonium phosphate mortar III
(Abdelrazig, Sharp & El-Jazairi, 1989): (a) after 1 hour, low magnification, (b) after 1 hour,
high magnification showing needle-like and cuboid crystallites.

229
Phosphate bonded cements

minutes. The compressive strength of an FEB SET-45 mortar (water: solid


= 1:16) increases with time. Compressive strength reached 24 MPa in the
first hour and 47 MPa after 24 hours. Thereafter, compressive strength
increased more slowly to 54 MPa after one week and 56 MPa after one

4U

Figure 6.9 The morphology of a commercial mortar, showing well-developed needle-like


crystallites. Micrograph span (a) 75 \im, (b) 30 (im (Abdelrazig et aL, 1984).

230
Magnesium phosphate cements
month. Strength depends on the amount of water in the system, and
compressive strength was drastically reduced when the water/ADP ratio
was increased to 1:8.
Microstructure is also affected by water content. The 1:16 mortar
contained hexagonal plates, while the 1:8 mortar had an ellipsoidal
morphology. The 1:5 mortar when prepared at — 5 °C developed striking,
well-formed, needle-like crystals (Figure 6.9).
The action of heat on these cements is complex (Abdelrazig & Sharp,
1988). The principal sequence based on XRD and thermal analysis is
shown in Figure 6.10.

6.4.7 Cement formation with diammonium hydrogen phosphate


Sugama & Kukacka (1983a) described cements based on magnesium oxide
and a 40 % solution of diammonium hydrogen phosphate (DAP) liquid.
The powder was a fine magnesium oxide that had been calcined above
1500 °C and had a surface area ofc. 1 m2 g"1. These cements set within 3
minutes and developed an early strength of 5-7 MPa after 30 minutes and
19-3 MPa in 15 hours. Sugama & Kukacka, using XRD, considered that
Mg3(PO4)2. 4H2O was the main reaction product. They also reported the
presence of other hydrates, Mg(OH)2 and struvite. However, Abdelrazig &
Sharp (1985) discounted the presence of Mg3(PO4)2. 4H2O and Mg(OH)2
in this system, and we are inclined to agree that it is unlikely that such

MgNH4PO4.6H2O Mg(NH 4 ) 2 (HP0 4 ) 2 .4H 2 0

MgNH4PO4.H2O [6.4]

l
MgNH 4 PO 4 .H 2 O (amorphous phase)
[6.5]
Temperajture
increasing \
[6.6]

MgHPO 4 (amorphous phase)


[6.7]

Mg 21P 2+OMgO
7

Mg3(PO4)2 [6.8]
Figure 6.10 Action of heat on ammonium magnesium phosphate cements (Abdelrazig &
Sharp, 1988).

231
Phosphate bonded cements

species are formed in the presence of ammonium ions. Most likely, the
reaction products are the hydrates of magnesium ammonium phosphate:
schertelite, struvite and dittmarite.

6.4.8 Cement formation with ammonium polyphosphate


Sugama & Kukacka (1983b) described cements based on magnesium oxide
and a 56 % aqueous solution of ammonium polyphosphate (APP). The
powder was a fine magnesium oxide that had been calcined above 1300 °C
and had a surface area of 1 to 5 m2 g"1. The reaction was strongly
exothermic; the cements set within 3 minutes and developed an early
strength of 13*8 MPa after 1 hour and over 20 MPa after 5 hours.
Micrographs taken after 30 minutes reaction showed the matrix to be
amorphous (Figure 6.11). After 1-5 hours, crystallites were observed. XRD
analysis showed that the most abundant phase was struvite. Sugama &
Kukacka also considered that Mg3(PO4)2. 4H2O was another major
reaction product and that other hydrates, Mg(OH)2 and newberyite,
MgHPO4.3H2O, were also present. Again, Abdelrazig & Sharp (1985)

Figure 6.11 Scanning electron micrographs showing the microstructure of a cement formed
from magnesium oxide and ammonium hydrogenphosphate solutions (Sugama & Kukacka,
1983b).

232
Magnesium phosphate cements
argued that Mg3(PO4)2. 4H2O, MgHPO4. 3H2O and Mg(OH)2 were
unlikely to be present in this system. Thus, it would appear that the
reaction products of these cements are struvite and other phases yet to be
identified. During the course of the reaction it would appear that P-O-P
bridges hydrolyse into orthophosphates.

6.4.9 Cement formation with aluminium acid phosphate


Ando, Shinada & Hiraoka (1974) examined cements formed by the
reaction between magnesium oxide and concentrated aqueous solutions of
aluminium dihydrogen phosphate. Later, Finch & Sharp (1989) made a
detailed examination of the cement-forming reaction and reported that the
reaction yielded cements of moderate strength.
They considered that cement formation was the result of an acid-base
reaction leading to the formation of hydrates by a through-solution
mechanism, by nucleation and precipitation from porefluids.Two phases
were found in the matrix, one amorphous and the other crystalline. The
crystalline phase was newberyite. Finch & Sharp concluded that the
amorphous phase was a hydrated form of aluminium orthophosphate,
A1PO4, which almost certainly contained magnesium. They ruled out a
pure A1PO4.«H2O, for they considered that the reaction could not be
represented by the equation

2MgO + A1(H2PO4)3 A 2MgHPO4. 3H2O + A1PO4. «H2O [6.9]


because cement formation was poor when the MgO/Al(H2PO4)3 mole
ratio was 2:1. It followed that the amorphous phase must be a composition
within the Al2O3-MgO-P2O5-H2O system. This conclusion was in line
with results of an energy-dispersive X-ray microanalysis of the fracture
surface using a scanning electron microscope.
As we have seen in Section 6.2, there is some evidence for supposing that
zinc phosphate cements contain an amorphous aluminium phosphate or
zinc aluminophosphate phase. Also, as we shall see in Section 6.5,
amorphous aluminium phosphate is the binding matrix of dental silicate
cement.
Scanning electron micrographs of fracture surfaces revealed the
presence of both amorphous and crystalline phases which corresponded to
results from XRD analysis (Figure 6.12). What is of interest is that the
crystalline phase is MgHPO4. 3H2O and not Mg(H2PO4)2. 2H2O or

233
Phosphate bonded cements
Mg(H2PO4)2. 4H2O as was the case in the reaction between MgO and
simple phosphoric acid solutions. Inspection of the diagrams of
Belopolsky, Shpunt & Shulgina (1950) shows that newberyite is the stable
phase at lower concentrations of phosphate. Presumably, in the present
case, aluminium locks up some phosphate and so reduces the phosphate
available for the magnesium phosphate phase.
Finch & Sharp (1989) found the mole ratio of MgO to A1(H2PO4)3 to be
an important parameter that affected both the reaction rate and the nature
of the reaction products. The critical mole ratio was 2:1. When the ratio
was less than 2:1 cements were not formed at all, and when it was exactly
2:1 the paste set slowly and always remained tacky. Further increases in
the ratio caused cements to set faster with greater evolution of heat. Finch
& Sharp (1989) also found that this ratio affected the proportion of
crystalline phase to amorphous phase in the cement matrix. The proportion
of newberyite in the matrix reached a maximum when the
MgO/A1(H2PO4)3 ratio was 4:1 and decreased to a low level when the ratio
was 8:1.
When Finch & Sharp (1989) used solutions of lower water content they
found an unknown XRD pattern that was distinct from that of
MgHPO4.3H2O. This unidentified phase they dubbed hayesite and
speculated that it might be a lower hydrate, either MgHPO4.2H2O or
MgHPO4.H2O. Infrared spectroscopy showed that hayesite was less well

Figure 6.12 Microstructure of MgO-aluminium hydrogenphosphate cement (Finch &


Sharp, 1989).

234
Dental silicate cement
crystalline than newberyite, but did contain bands at c. 3400 cm"1 and
1640 cm"1, showing it to be hydrated.
A higher hydrate, phosphorresslerite, MgHPO4. 7H2O, was formed
when these cements were exposed to water and dried in air. On ageing, this
hydrate readily transformed back to newberyite.

6.4.10 Cements formed from magnesium titanates


Cements have been prepared from magnesium titanates (Mg2TiO4,
MgTiO3 and Mg2Ti2O5) and phosphoric acid (Sychev et ai, 1982; Sudakas,
Turkina & Chernikova, 1982). The reaction product, MgHPO4. 3H2O, is
crystalline when Mg2TiO4 is used and amorphous with the other
magnesium titanates. The amorphous product gives the stronger cements.

6.5 Dental silicate cement


6.5.1 Historical
Dental silicate cement was once the most favoured of all anterior (front)
tooth filling materials. Indeed, it was the only material available for the
important task of aesthetic restoration from the early 1900s to the mid
1950s, when the not very successful simple acrylic resins made their
appearance (Phillips, 1975). In the mid sixties there were some 40 brands
available (Wilson, 1969) and Wilson et al. (1972) examined some 17 of
these. Since that time the use of the cement has declined sharply. It is rarely
used and today only two or three major brands are on the market. The
reason for this dramatic decline after some 50 years of dominance is closely
linked with the coming of modern aesthetic materials: the composite resin
from the mid 1960s onwards (Bowen, 1962), and the glass-ionomer cement
(Wilson & Kent, 1971) from the mid 1970s.
Dental silicate cement was also variously known in the past as a
translucent, porcelain or vitreous cement. The present name is to some
extent a misnomer, probably attached to the cement in the mistaken belief
that it was a silicate cement, whereas we now know that it is a phosphate-
bonded cement. It is formed by mixing an aluminosilicate glass with an
aqueous solution of orthophosphoric acid. After preparation the cement
paste sets within a few minutes in the mouth. It is, perhaps, the strongest
of the purely inorganic cements when prepared by conventional methods,
with a compressive strength that can reach 300 MPa after 24 hours
(Wilson et al., 1972).

235
Phosphate bonded cements

The early history of the cement is obscure. Dreschfeld (1907) and


Sanderson (1908) attributed its invention to Fletcher. Fletcher (1878,1879)
certainly described cements formed from concentrated orthophosphoric
acid solutions and sintered mixtures of oxides which included SiO2, A12O3,
CaO and ZnO. One was reported by Fletcher (1879) as being slightly
translucent. These cements were not successful in clinical use.
Voelker (1916a) reported three early dental silicate cements which
appeared in 1895, 1897 and 1902; all proved inadequate. The first
successful material was developed by Steenbock (1903,1904) who explicitly
sought and formulated a translucent cement (Voelker, 1916a,b). It was
marketed by Ascher in 1904 as New Enamel Richters; Harvadid cement
followed in the same year. Thereafter development was rapid and eight
varieties were reported by Morgenstern in 1905. However, from their
chemical composition we doubt whether they were sufficiently translucent.
The liquids used in these early formulations were 50% solutions of
orthophosphoric acid, often containing aluminium and zinc. Chemical
analyses were published between 1904 and 1972 (Voelker, 1916a; Greve,
1913; Watts, 1915; Paffenbarger, Schoonover & Souder, 1938; Axelsson,
1964; W i l s o n s al, 1972).
The glasses used by Steenbock in his original compositions were mixtures
of calcium aluminosilicates and beryllium silicates; but, as Dreschfeld
reported in 1907, subsequent developments moved away from the use of
beryllium compounds. Published chemical analyses in the period to 1916
(Voelker, 1916a; Greve, 1913; Watts, 1915) confirmed Dreschfeld's
statement. In the following we shall refer to these as oxide glasses.
In 1908 Schoenbeck made the most important of all compositional
innovations when he introduced the use of fluoride-fluxed glasses into
dental silicate cement. In this discussion we shall refer to these as fluoride
glasses, although they are, in fact, mixed fluoride and oxide glasses. Over
the years fluoride glasses have progressively replaced the purely oxide ones.
Although materials based on fluoride-containing aluminosilicate glasses
were rare before 1920 (Watts, 1915; Greve, 1913), and Wright (1919)
ignored them in his studies, by 1938, Paffenbarger, Schoonover & Souder
(1938) reported that most dental silicate cements were of the Schoenbeck
type. This development is not surprising because, besides lowering the
temperature of fusion, fluoride confers greater translucency and strength
on the cement and has beneficial therapeutic effects. There have been no
major compositional innovations since.
Dental silicate cement is used exclusively for the aesthetic restoration of

236
Dental silicate cement

anterior (front) teeth. In this situation, unlike cements used for cemen-
tation, the cement is fully exposed to erosive attack by oral fluids. Thus,
considerable interest was shown in the physical, chemical and biological
properties of these cements in the years following their introduction. Early
workers in the field were Morgenstern (1905), Kulka (1907), Rawitzer
(1908, 1909), Proell (1913) and Poetschke (1916). These workers reported
a number of faults, finding the cements to be porous, prone to staining,
attacked by mouth acids, shrinking under drying conditions and lacking in
adhesion. These observations have been confirmed during the course of
time. In certain mouths, dental silicate cement stains, erodes and even
washes away. For this reason it has never been fully satisfactory and,
consequently, now that alternatives have become available, it has fallen
out of general use.
The first period of development ended with the research of Wright
(1919) who published the results of an extensive survey of cements
prepared from experimental SiO2-Al2O3-CaO glasses and orthophos-
phoric acid solutions containing aluminium phosphate. By this time the
main cement formulations had been established. Between 1919 and 1950
only minor improvements were attempted; these were of a technological
nature and unsuccessful.
After 1950 some serious attempts were made to improve dental silicate
cement. Notable were those of Manly et al. (1951) and Rockett (1968);
also, Pendry reported an acid-resistant cement containing indium (Pendry
& Cook, 1972; Pendry, 1973). None of these experimental materials went
into production. These attempts came too late in view of progress made in
other directions, and the picture of dental silicate cement remains one of an
essentially traditional material which has changed little since the original
developments made mainly by German chemists prior to 1914.
The main line of development now lies with its successor, the
glass-ionomer cement, which uses a similar glass, but in which phosphoric
acid is replaced by poly(acrylic acid); this cement is more resistant to acid
erosion and staining and has the great advantage of adhesion to tooth
material.

6.5.2 Glasses
The powders used in dental silicate cement are unusual in being ground
opal glasses rather than crushed crystalline clinker. The glassy nature of
the powder gives the set cement the unusual property of translucency

237
Phosphate bonded cements
which it shares with one other cement, its successor, the glass-ionomer
cement. These glasses are calcium aluminosilicates which, in materials
available since the 1940s, always contain fluoride. Essentially they are
based on the SiO2-Al2O3-CaF2 system.
The aluminosilicate glass has a dual role: it acts as afillerand is also the
source of ions required to gel phosphoric acid solutions. These glasses act
as a source of ions because they are decomposed by acids. This property is
dependent on the Al/Si ratio being sufficiently high, approaching 1:1. The
reasons and criteria for the decomposition of aluminosilicates are
examined fully in Section 5.9.2.
The glasses are similar to those used in glass-ionomer cements but their
reactivity towards acids has to be less, as orthophosphoric acid is a
stronger acid than the poly(alkenoic acid)s. The consequence is that the
Al/Si ratio, which determines reactivity, is lower than in the glass-ionomer
cement glasses.
Typical glass compositions are given in Table 6.6. The preparation of
these glasses is given in Section 5.9.2. The essential components are ground
silica (quartz), alumina and fluorite (which also acts as a flux). Cryolite is
added as an additional flux and minor amounts of aluminium phosphate
are present, which apparently improve the mixing qualities of the cement.
The ratio of alumina to silica controls the setting time of the cement.
Fluoride tends to slow setting while aluminium phosphate improves the
mixing of the paste.
The temperature of fusion is 1050 to 1300 °C, a somewhat lower range
than that for the glass-ionomer cement glasses. The melts are shock-cooled
and ground. The median particle size ranges from 8-6 to 11-5 |im. The
glasses are slightly opal in appearance and evidence from electron
microscopy shows that this arises from phase separation (Wilson et al.,
1972). The separated phase appears as droplets although it may be
spinodal. An example of one glass (Super Syntrex) showed the segregation
of larger droplets of uniform size c. 400 nm in diameter and smaller
droplets of 20 to 30 nm (Figure 6.13). Etching with phosphoric acid
showed that the droplets were selectively attacked by acids.

Effect of glass composition on cement properties


As we have indicated previously, two types of glass have been used in
dental silicate cements: the obsolete oxide glass and the modern fluoride
glass. Only four studies on glass composition and its relationship to cement
property have been published (Wright, 1919; Crepaz, 1951; Manly et al.,

238
Dental silicate cement

Table 6.6. Chemical composition of commercial dental silicate


cements {Wilson et al., 1972)

Powders Liquids

Species % by mass Species % by mass


SiO2 31-5-41-6 (35-9) H 3 PO 4 48-8-55-5 (65-9)
A12O3 27-2-29-1 (290) Al 1-6-2-5 (00)
CaO 7-7-9-0(6-1) Zn 4-2-9-1 (0-0)
Na 2 O 7-7-11-2(14-5) Mg 0(1-6)
F 13-3-22-0(15-2) Be 0 (0-65)
PA 3-0-5-3 (4-4)
ZnO 0-1-2-9 (0-3)
MgO 0-0-0-1 (00)
SrO 0-0-0-2 (00)
H2O 1-6-2-2

Based on four typical examples and one atypical example.


Composition of the atypical example is in parentheses.

o • O:

Figure 6.13 Electron micrograph of a single-stage replica of a dental silicate cement glass,
showing phase-separated droplets rich in calcium and fluoride: large droplets 400 nm in
diameter and small droplets 20 to 30 nm in diameter (Wilson et al., 1972).

239
Phosphate bonded cements
1951; Rockett, 1968) and these are concerned almost entirely with non-
fluoride glasses of the SiO2-Al2O3-CaO and SiO2-Al2O3-CaO-P2O5 types.
In their studies on SiO2-Al2O3-CaO glasses Manly et al. (1951) found
that the SiO2/CaO ratio controlled the setting rate of the cement pastes. If
this ratio was greater than 2-08 by mass then the glass powder-liquid paste
did not set; when it was less than 1-74 pastes set rapidly to form cements.
Between these extremes there were narrow compositional bands corres-
ponding to slow-setting cements (ratio = 1-95) and moderate-setting
cements (ratio = 1-77-1-89).
Since the 1940s all commercial dental silicate cements have used fluoride
glasses. Fluoride glasses yield more translucent cements than oxide glasses
because they have lower refractive indices. It is also probable, by analogy
with glass-ionomer cements, that they yield stronger cements (Section
5.9.4). Unfortunately, in the four studies cited above, the workers were
concerned almost entirely with non-fluoride glasses and made no sys-
tematic studies on fluoride glasses. Only Manly et al. (1951) made even a
cursory examination of these glasses, which are the basis of practical
cements. Consequently, none of these workers were able to improve on or
even equal the performance of the commercial examples. Nor were they
able to shed much light on fundamental compositional factors controlling
the setting of cements based on fluoride glasses.
It was left to Kent & Wilson (1968), in unpublished observations, to
discover that in glasses based on SiO2-Al2O3-CaF2 compositions the
Al/Si ratio controlled the rate at which the cement paste set. These
observations laid the foundation for the development of the glass-ionomer
cement, during which most of the work onfluorideglasses was done. This
topic is covered in detail in Section 5.9.2.
The degree of subdivision of the powder has considerable effect on the
properties of cements (Swanson, 1936; Charbeneau, 1961; Kent & Wilson,
1971). Kent & Wilson (1971) showed that the cements prepared from a
fine-grain powder when mixed at the same powder/liquid ratio as normal-
grain powder had greater strength, but set too fast for practical use and
lacked translucency (Table 6.7). Moreover, cements prepared from
normal-grain powder could be mixed to a higher powder/liquid ratio,
resulting in a notable increase in strength combined with an optimum
setting time.

240
Dental silicate cement

Table 6.7. Effect of particle size of powder on cement properties (Kent &
Wilson, 1971)

Powder

Fine Normal Normal

Surface area, um * 1-62 0-87 0-87


Powder: liquid ratio, g cm"3 3125 3125 400
Setting time (37 °C), 2-5 90 4-0
minutes
Wet compressive" 289 176 217
strength (24 h), MPa
Wet tensile" 14-7 6-2 13-6
strength (24 h), MPa

a
After storage for 24 hours in water at 37 °C.

6.5.3 Liquid
Dental silicate cement liquids are concentrated aqueous solutions of
orthophosphoric acid generally containing aluminium and zinc (Wilson,
Kent & Batchelor, 1968; Kent, Lewis & Wilson, 1971a,b; Wilson et al,
1972). The optimum orthophosphoric acid concentration is 48 to 55 % by
mass (Wilson et al., 1970a), although higher concentrations are en-
countered. Aluminium is present as phosphate complexes and zinc as a
simple ion (see Section 6.1.2). Examples are given in Table 6.6.

Effect of liquid composition on cement properties


The properties of dental silicate cements are affected both by the
concentration of phosphoric acid and by the presence of metal salts.
The effect of the concentration of orthophosphoric acid on cement
properties has long been known (Poetschke, 1916; Ray, 1934; Worner &
Docking, 1958; Wilson et al., 1970a). The setting time of a cement paste
increases as the orthophosphoric acid concentration increases; this effect is
particularly marked above 65% H3PO4 (Figure 6.14). There are several
reasons for this sharp change. Water is required to act as a reaction
medium and also to hydrate reaction products; a deficiency could therefore
retard or even arrest the reaction. There is also the possibility of a change
in structure of the orthophosphoric acid solution as its concentration
increases (see Section 6.1.1).

241
Phosphate bonded cements

Cement strength and resistance to aqueous attack are also critically


dependent on phosphoric acid concentration, and there is an optimum
concentration for the development of maximum strength and resistance to
aqueous attack (Wilson et ai, 1970a; Wilson, Paddon & Crisp, 1979). The
effect is particularly critical when the phosphoric acid liquid contains
aluminium and zinc.
This sensitivity to water has practical implications. A cement liquid is
stable in an atmosphere of 70 % relative humidity. It will gain water in
more humid atmospheres and lose it to drier ones, and this will adversely
affect cement properties (Paffenbarger, Schoonover & Souder, 1938;
Worner & Docking, 1958; Wilson et al., 1970a).
All commercial examples of phosphoric acid solutions used in these
cements contain metal ions, whose role has been discussed in Section 6.1.2.
In the case of the dental silicate cement, aluminium and zinc are the metals
added to liquids of normal commercial cements and have a significant
effect on cement properties (Table 6.8) (Wilson, Kent & Batchelor, 1968;
Kent, Lewis & Wilson, 1971a,b). Aluminium accelerates setting for it
forms phosphate complexes and is the principal cation of the phosphatic
matrix. Zinc retards setting for it serves to neutralize the acidic liquid - it

3OO

4O 5O 6O 7O 8O
liquid composition °/ow/w H3PC
Figure 6.14 Effect of liquid composition on the setting time and strength of dental silicate
cements (Wilson et al, 1970a).

242
Dental silicate cement

Table 6.8. Effect of metals contained in the phosphoric acid liquid on


cement phosphate properties (Wilson, Kent & Batchelor, 1968)

Al, Zn, Powder: liquid, Setting time, Compressive strength


% % g cm"3 minutes (wet, 24 hours), MPa

00 00 3-9 3-5 169


00 130 40 4-5 267
2-5 00 3-9 3-0 269
1-25 6-5 40 3-5 291

forms simple salts - and thus reduces the rate of extraction of cations from
the aluminosilicate powder. When both metals are present these opposing
effects tend to cancel out. Both metals, alone or in combination in the
liquid, serve to improve the strength of cements, a combination being most
effective.

6.5.4 Cement-forming reaction

The setting reaction of dental silicate cement was not understood until
1970. An early opinion, that of Steenbock (quoted by Voelker, 1916a,b),
was that setting was due to the formation of calcium and aluminium
phosphates. Later, Ray (1934) attributed setting to the gelation of silicic
acid, and this became the received opinion (Skinner & Phillips, 1960).
Wilson & Batchelor (1968) disagreed and concluded from a study of the
acid solubility that the dental silicate cement matrix could not be composed
of silica gel but instead could be a silico-phosphate gel. However, infrared
spectroscopy failed to detect the presence of P-O-Si and P-O-P bonds
(Wilson & Mesley, 1968).
The nature of the setting reaction was finally elucidated by Wilson et al.
(1970a), who established that formation of an aluminium phosphate gel
was responsible; although siliceous gel was also formed it merely coated
the partly reacted glass particles.
The following account is based mainly on the studies of Wilson and
coworkers, with some re-interpretation of experimental data. The com-
position of the cement used is given in Table 6.9. In brief, the reaction takes
place in several overlapping stages: extraction of ions from the glass,
migration of cations into the aqueous phase, precipitation of insoluble
salts as pH increases, leading to formation of an aluminium phosphate gel.

243
Phosphate bonded cements

Table 6.9. Chemical composition of the dental silicate cement


used in reaction and structural studies

Powders Liquids

Species % by mass Species % by mass


SiO2 41-6 H 3 PO 4 48-8
A12O3 28-2 Al 1-6
CaO 8-8 Zn 61
Na 2 O 7-7
F 13-3
P2O5 3-3
ZnO 0-3
H2O 2-2
Less O for 2F -5-6
Total 99-8

In the subsequent hardening phase, precipitation and hydration continue.


The set cement consists, essentially, of partly-reacted glass particles
embedded in an aluminium phosphate gel. The morphology of the filler
particles is one where a glass core is sheathed by silica gel.
The cement-forming reactions may be described as follows. On mixing
the powder and liquid, hydrogen ions from the phosphoric acid solution
attack the glass particles, which are decomposed to silicic acid (Wilson &
Batchelor, 1967a; Wilson & Mesley, 1968). Al 3+ , Ca 2+ , Na + , and F~ ions
are released (Wilson & Kent, 1970a), the pH of the aqueous phase increases
(Kent & Wilson, 1969) and, as infrared spectroscopy shows, H 3 PO 4 ionizes
to H 2 PO 4 (Wilson & Mesley, 1968). An electrical imbalance results and
under the influence of an electrostatic field cations migrate into the
aqueous phase where they accumulate. Most probably, aluminium and
fluoride form cationic A1F2+ and A1F2 complexes, whose existence has
been reported by Connick & Poulsen (1957), O'Reilly (1960) and Yamazaki
& Takeuchi (1967).
The most important cation is aluminium. In the absence of fluoride,
aluminium is present in solution as the Al(H2O)g+ complex which
hydrolyses to form complex multinuclear species such as [A12(OH)2]4+ and
[A113O4(OH)24(H2O)12]7+ (Aveston, 1965; Waters & Henty, 1977). There
are also two kinds of phosphate complexes (Akitt, Greenwood & Lester,
1971; O'Neill et al, 1982): those based on the H 3 PO 4 ligand, A1(H3PO4)3+

244
Dental silicate cement
and complexes of unknown protonation, A1(H3PO4)W, where n ^ 2; and
those where the ligand is H2PO4, A1(H2PO4)2+ and A1(H2PO4)J.
As pH increases and the ionization of H3PO4 to H2PO4 continues (Kent
& Wilson, 1969), the formation of those aluminophosphate complexes
based on the H2PO4 ligand is favoured. Further increases in the
concentration of aluminium and H2PO4 give rise to the formation of
binuclear complexes in which aluminium and phosphate are linked by
oxygen bridges (Akitt, Greenwood & Lester, 1971; O'Neill et ai, 1982).
Most probably, this process continues with the formation of multinuclear
complexes and networks based on Al-O-P linkages, which leads to
gelation.
The presence of fluoride complicates this picture but does not change its
essence. Using 31P, 19F and XH NMR Akitt, Greenwood & Lester (1971)
found numerous complexes in such solutions. There were 19F NMR peaks
associated with A1F2+ and A1F2 complexes and, if the fluoride content was
high, a peak corresponding to exchanging AlFjf"w)+, F", HF and HF2.
In addition there were peaks corresponding to fluorine-containing
aluminophosphate complexes which were similar to the aluminophosphate
complexes noted above but with the addition of one, two or three fluoride
ligands.
The insolubilization of ions during setting and hardening was followed
experimentally by Wilson & Kent (1970b) and is shown in Figure 6.15.
This figure shows the time-dependent variation of the concentration of
soluble ions during setting and hardening. There are two competing
processes which govern the concentration of ions: extraction from the
glass and removal by precipitation. The extractive process is illustrated by
the [Na]/time curve, as Na+ is not precipitated during cement formation.
Judging by the concentration of soluble sodium ions extraction is half
complete during mixing of the paste and over within 10 minutes, when
20 % of the glass powder is decomposed.
The progress of precipitation is revealed by the concentration/time
curves for zinc and phosphate, since both these species are present initially
in solution. There should be maxima for the soluble aluminium, calcium
and fluoride which are extracted from the glass, but because of the early
onset of precipitation these are not observed. Precipitation is accompanied
by an increase in pH; when it reaches 1-8, at which juncture 50% of both
zinc and phosphate have been precipitated, the cement paste gels (5
minutes after preparation).
Hardening continues after gelation, rapidly reaching 65 % of its final

245
Phosphate bonded cements

value within 30 minutes, and ceasing after about 72 hours. The pH


continues to rise, reaching 5-2 after 48 hours. Precipitation of aluminium
and calcium ions appears to be complete within an hour but zinc continues
to precipitate. Sodium and fluoride ions do not completely precipitate.

- IOO

30
CEMENT AGE (min)

Figure 6.15 The time-dependent variation of the concentration of soluble ions during the
setting and hardening of a dental silicate cement (Wilson & Kent, 1970b).

246
Dental silicate cement

Evidence from electrical conductivity experiments (Wilson & Kent,


1968) indicates that, even after hardening is apparently complete, the
reaction continues for at least 7 weeks; indeed it is known from the work
of Paffenbarger, Schoonover & Souder (1938) that the cement continues to
strengthen for at least a year.
The correlation of phosphate precipitation with decrease of conductivity
(Wilson & Kent, 1968), increase in pH (Kent & Wilson, 1969) and hardness
(Wilson et al., 1972) is shown in Figure 6.16. These results demonstrate the
relationship between the development of physical properties and the
underlying chemical changes, but there are no sharp changes at the gel
point. Evidence from infrared spectroscopy (Wilson & Mesley, 1968) and
electron probe microanalysis (Kent, Fletcher & Wilson, 1970; Wilson et
al., 1972) indicates that the main reaction product is an amorphous
aluminophosphate. Also formed in the matrix were fluorite (CaF2) and
sodium acid phosphates.
The fate of silicic acid is of some interest. Silicic acid polymerizes, by
condensation, and finally a silica gel is formed (Wilson & Mesley, 1968).
The insolubilization of silicic acid has been observed to parallel closely the
precipitation of phosphate (Wilson & Batchelor, 1967b) and is related to
an increase of pH within the cement (Kent & Wilson, 1969). A low
concentration of silicic acid must remain in the matrix. All this is in accord
with the known aqueous chemistry of silica.
Orthosilicic acid at concentrations above 100 ppm in solution condenses
to a dimer. At higher concentrations cyclic and polymeric species are
formed (Her, 1979; Andersson, Dent Glasser & Smith, 1982). These
processes are ones where silanol groups condense to form siloxane
linkages:

—OH + HO—Si^- = -^Si—O—Si^- + H2O

According to Vysotskii et al. (1974) gelation is the result of a complex


process, and polymerization does not necessarily ensure gelation. The
formation of polymeric particles is followed by their growth. These
particles are then linked by siloxane bonds to form branched chains.
Networks extend throughout the liquid phase, and finally gelation occurs,
provided the pH is below 7. Gelation is most rapid for pH = 5 to 6 and
minimal for pH = 1-5 to 3 (Her, 1979). The observations of Wilson &
Batchelor (1967b) are in accord with this theory.
The role of water is important, for it acts as a reaction medium and

247
Phosphate bonded cements

100r

500 50

40

30

20

10

Conductivity

20 40 OOmin 4h 24h
Figure 6.16 The relationship between the development of hardness and the underlying
physicochemical process: decrease in phosphate concentration, increase in pH, and decrease
in electrical conductivity (Wilson et al., 1972).

248
Dental silicate cement

serves to hydrate reaction products. During setting and hardening, water


becomes progressively bound to the matrix and in the fully hardened
cement the ratio of water of hydration (non-evaporable water) to loosely
bound (evaporable) water reaches 1:1 (Wilson, Paddon & Crisp, 1979).
A deficiency of water can inhibit the reaction. Kent & Wilson (1969)
noted that storage conditions affected the rate of reaction as measured by
the increase in pH over 18 hours at 37 °C. When water was allowed neither
to enter nor to leave the cement the pH rose to 5-15; when cured in water
the reaction was enhanced and the pH rose to 5-40; but when stored under
desiccating conditions the reaction was retarded and the pH only rose to
4-30.
A deficiency of water in the cement liquid has the same effect and this
occurs when the H3PO4 content exceeds 60%. Wilson & Mesley (1968)
noted that in a cement formed from a solution of 65 % H3PO4 there was
evidence of incomplete reaction even after 6 hours. We have noted in
Section 6.5.3 that there is a sharp decline in the rate of reaction when the
orthophosphoric acid concentration exceeds 65% H3PO4 (Figure 6.14).
The avidity of cements to absorb water from humid surroundings also
increases sharply when the phosphoric acid in the cement-forming liquid
exceeds 60%. It is difficult to avoid the conclusion that these two
phenomena are related and that a deficiency of water retards the cement-
forming reaction.

6.5.5 Structure
Information on the microstructure and molecular composition of the set
cement comes almost entirely from the optical, electron microscopic and
electron probe microscopic analysis of Kent, Fletcher & Wilson (1970),
Wilson et al. (1970a, 1972) and Brune & Smith (1982). There are some
differences which require resolution. Although under optical microscopy
the set cement appears as an irregular mosaic of angular, and seemingly
unattacked, glass particles connected by an apparently structureless
matrix, its appearance under the reflectance optical microscope is more
revealing (Figure 6.17). Here the glass particles are shown to be pitted with
hemispherical craters and to vary in size from 1 to 100 urn. The matrix is
seen as particulate. A stereoscan of a fractured surface reveals more detail
and shows an angular glass particle which has debonded from a particulate
matrix. The surface of the glass particle is spotted where selective acid

249
Phosphate bonded cements

attack has occurred at the site of phase-separated droplets (Figure 6.18).


Apart from this etching the glass particles appear unattacked, but this is
not the case.
A combination of infrared spectroscopy and electron probe micro-
analysis shows that the matrix of the set cement is amorphous aluminium
phosphate (Wilson et al., 1970a). Silica gel formed appears to sheathe the
partly reacted glass particles. Two crystalline phases have been detected in
the matrix: fluorite, CaF 2 , and augelite, A12(OH)3PO4 (Wilson et al., 1972).
However, the XRD pattern of the matrix did not quite correspond to that
of augelite in published powder diffraction data, and the possibility exists
that F replaces OH to give A12F3PO4; this would be in accordance with the
formation of fluorine-containing aluminophosphate complexes noted
previously.
The element distributions of Si, P, F, and Na within a dental silicate
cement were recorded by Kent, Fletcher & Wilson (1970) and Wilson et al.
(1972) who used cements from which fine particles had been removed, and

Figure 6.17 Phase contrast micrograph of a polished section of a dental silicate cement,
showing angular glass particles, size 1 to 100 urn, embedded in a featureless matrix (Wilson
et al, 1972).

250
Dental silicate cement

by Brune & Smith (1982) who used a normal powder. The element
distribution maps for Si, Al and P are shown in Figure 6.19, where white
highlights indicate the positive presence of an element. An optical
micrograph of the area of study shows the presence of a distinct outer layer
around each reacted glass particle (Figure 6.19a). Silicon is present only in
the particles and its concentration is enhanced at the boundary (Figure
6.196). This implies the existence of a layer of silica gel surrounding each
glass particle. The gel-like nature of the silica sheath can be clearly seen in
Figure 6.19a where it has detached from the glass core. Phosphorus is
found substantially only in the matrix (Figure 6.19c). Aluminium is found
both in the glass particles and in the matrix, showing that it has migrated.
A prominent feature is the depletion of Al at the boundary (Figure 6A9d).
Calcium, sodium and fluorine were also shown to have migrated from the
glass particles to the matrix.
These findings support the view that during the reaction ions are
extracted from the surface of the glass particles, migrate to the aqueous
phase where they form the matrix, and leave a silica gel relict. This explains
why the glass particles appear to be unattacked when examined under the
microscope. The presence of both Al and P in the cementing matrix and the

Figure 6.18 A stereoscan of a fracture surface of a dental silicate cement. The debonded glass
particle is to be identified by its pitted surface, the result of selective acid attack. Note the
particulate nature of the matrix (Wilson et al., 1972).

251
Phosphate bonded cements

knowledge that infrared spectroscopy had identified the presence of an


aluminium phosphate led to the conclusion that the matrix was an
aluminium phosphate hydrogel (50 % of the water is bound to the matrix).
Some of these conclusions may require revision, since recent findings of
Ellison & Warrens (1987) on the related glass-ionomer cement (Section
5.9.6) suggest that acid attack occurs throughout the body of the glass
particle and not just at the surface layer. In that case the silica gel layer is
not a relict but a zone of gelation. This view is more in accord with ideas
on the decomposition of aluminosilicate glasses.

Figure 6.19 Element distribution maps for a scanning electron micrograph of a dental silicate
cement: (a) optical micrograph of area of study, (b) Si distribution, (c) P distribution, (d) Al
distribution (Wilson et al, 1972).

252
Dental silicate cement
A close examination of Figure 6.19&, especially the two closely adjacent
particles formed by the splitting of one particle, reveals that the Si regions
appear to extend slightly beyond the boundaries of the glass particles. Thus
it is possible that the whole of the glass structure is slightly acid-degraded,
although remaining essentially glassy, and that silicic acid and ions are
released. These species migrate outwards, the ions into the aqueous phase,
while the majority of the silicic acid condenses to a shell of silica gel just
beyond the particle-matrix boundary.
This alternative hypothesis may explain the observational differences
between different workers. Brune & Smith (1982), unlike Wilson et al
(1972), found Si distributed throughout the cement but were uncertain
whether it was due to Si in the matrix or the degradation of fine particles
to silica gel. But Brune & Smith (1982) used a normal glass powder while
Wilson et al. (1972) removed fine particles to improve resolution. These
differing observations are reconciled if the silicic acid which is formed
migrates slightly before condensing to silica gel.
Summary
Overall, the formation of dental silicate cement can be represented in
outline as in Figure 6.20.

6.5.6 Physical properties


Dental silicate cement is used for the aesthetic restoration of anterior
(front) teeth because it is translucent and so can be made to colour-match
tooth enamel. It is prepared by introducing powder into the liquid
gradually in order to dissipate heat, although the exotherm is not so great
Calcium aluminosilicate glass
+
Phosphoric acid solution

Aluminophosphate Calcium Sodium Silica


complexes fluoride acid phosphates gel
(soluble)

Al phosphate
polynuclear complexes- Aluminium phosphate gel
Figure 6.20 Formation of dental silicate cement.

253
Phosphate bonded cements

Table 6.10. Properties of commercial dental silicate cements

Specification"
Refs. Value limits

Powder: liquid*, g cm"3 [1] 2-70-4-02 n


Working time (23 °C), [2] 3-6 n
minutes
Setting time (37 °C), [1,2] 3-25-7-0 3-8
minutes
Compressivec [1,2,3] 68-5-255 166 minimum
strength (24 h), MPa
Compressivec [4,5] 18-0-21-0 n
modulus (24 h), GPa
Flexuralc [2] 24-5 n
strength (24 h), MPa
Tensilec [6] 13-6 n
strength (24 h), MPa
Fracture toughness0 [7] 012-0-30 n
(24 h), MN m-f
Solubility & [1] 0-34-3-8 1-0 maximum
disintegration (24 h), %
Opacity, C 0 7 [1] 0-42-0-71 0-35-0-55

a
BS 3365/1: 1969 Specification for Dental Silicate Cement and Dental
Silicophosphate Cement. Part 1 Dental Silicate Cement.
6
For a consistency spread of 25 mm diameter for 0-5 cm3 of cement paste under
a load of 1-5 Kgf applied after 2 minutes at 23 °C.
c
After storage for 24 hours in water at 37 °C.
n no specification test.
[1] Wilson et aL, (1972); [2] 0ilo (1988); [3] Wilson (1975c); [4] Paddon &
Wilson (1976); [5] Wilson, Paddon & Crisp (1979); [6] Kent & Wilson (1971);
[7] Lloyd & Mitchell (1984).

as that of zinc phosphate cement (Crisp, Jennings & Wilson, 1978). The
cement is mixed very thickly and the powder/liquid ratio can be as high as
4 g cm"3 for good examples of this cement (Wilson et aL, 1972). At these
thick consistencies the working time is good; an isolated observation
indicates that it is 3-6 minutes at 23 °C (0ilo, 1988). Setting time (37 °C)
varies from 3-25 to 7-0 minutes (Wilson et aL, 1972; 0ilo, 1988). Properties
are summarized in Table 6.10.

254
Dental silicate cement

The dental silicate cements have brittle characteristics which are


immediately evident after set (Paddon & Wilson, 1976). The best examples
develop a compressive strength of about 250 MPa after 24 hours (Wilson
et al., 1972; Wilson, 1975c; 0ilo, 1988), which is higher than that recorded
for any other acid-base cement, including the glass polyalkenoate cement
(Table 6.10). Strength continues to increase for at least a year (Paffen-
barger, Sweeney & Isaacs, 1933). Compressive modulus is 18 to 21 GPa
after 24 hours (Paddon & Wilson, 1976; Wilson, Paddon & Crisp, 1979).
Modulus changes with time and for one example increased from 18-0 GPa
after 24 hours to 30-6 GPa after 30 days.
Little information is available on other tests of strength. Isolated
measurements give a flexural strength of 24-5 MPa (0ilo, 1988) and a
tensile strength of 13-6 MPa (Kent & Wilson, 1971). These values lie within
the range of those recorded for glass polyalkenoate cement. Translucency
is easily achieved as values for the inverse property of opacity show (Table
6.10).
Most of the properties of a dental silicate cement are affected by
preparative variables, particularly the powder/liquid ratio (Jorgensen,
1963; Wilson & Batchelor, 1967b). Increase in the powder/liquid ratio
accelerates set and increases strength and resistance to erosion (Figure
6.21). Temperature and, to a lesser extent, humidity during mixing have
some effect, but chiefly they affect setting time.
Although these cements have high compressive strength, their low
flexural and tensile strengths coupled with brittleness and lack of toughness
makes them suitable only for low-stress anterior (front teeth) restorations.

6.5.7 Dissolution and ion release


The dissolution and ion release from dental silicate cement have been the
most investigated characteristics; with good reason, for they are central to
its clinical performance. Erosion limits its life but release of fluoride has
important clinical consequences.

Ion release
In neutral solution when fully hardened, dental silicate cements are
resistant to aqueous attack. Before they have fully hardened, set cements
contain soluble reaction intermediates - soluble sodium salts, acid phos-
phates and fluorides - which render them vulnerable to attack even by
neutral solutions including saliva (Wilson, 1976).

255
Phosphate bonded cements

Irretrievable loss of matrix-forming cations and anions can result in


permanent damage to the cement surface. This is visible as milky or chalky
patches or even raised blisters. For this reason it is customary to protect,
temporarily, the freshly placed cement by varnish. Once hardened, attack
by neutral solutions causes failure only when a cement has been poorly
formulated and contains excessive amounts of soluble reaction products.
In this case osmotic effects can cause blistering or even disintegration
under the action of internal forces, as Figure 6.22 illustrates (Wilson &
Batchelor, 1967a).
The composition of the leachates does not correspond to the com-
position of the cement at all (Wilson & Batchelor, 1967a,b). The
predominant species eluted are the soluble sodium salts of phosphate and
fluorides, although sodium is only a minor constituent of the cement. For
one example of cement examined, the leachate contained 0-28 % sodium
and 0-20 % phosphate (expressed as a percentage of the amount of the
species contained in the cement). For the major constituents of the glass the
figures were 0-07 % fluoride, 0-02 % A12O3, 0-01 % SiO2 and 0-003 % CaO.

3.5 4.0
Powder/Liquid Ratio (g/ml)

Figure 6.21 The effect of powder/liquid ratio on setting time and compressive strength of a
dental silicate cement (Wilson & Batchelor, 1967b).

256
Dental silicate cement
The rate of elution declines sharply with time and the pattern of elution
changes. The acid phosphate ions, HgPO^ and HPO|~, are removed by
further reaction or by elution, and the release of phosphate changes from
predominant to minor. Thereafter, the rate of loss of phosphate is
governed by the phosphate concentration of the solution; indeed if the
phosphate concentration of the solution is sufficiently high the process is
reversed and the cement takes up phosphate. So, clearly, an ion exchange
phenomenon is involved (Kuhn & Wilson, 1985; Kuhn, Winter & Tan,
1982).
Elution of ions is accompanied by absorption of water and this can
amount to as much as 20% by mass in five days (Kuhn et al., 1982). The
extent of water uptake is affected by the ionic concentration of the solution.

Fluoride release
As the cements age, sodium, fluoride and silica become the major species
eluted, although the amounts involved are small. Fluoride is released in a
sustained fashion over a prolonged period (Wilson & Batchelor, 1967a; de

Figure 6.22 The effect of water on a poor example of a dental silicate cement. An osmotic
force causes blistering and disintegration (Wilson & Batchelor, 1967a).

257
Phosphate bonded cements
Freitas, 1968; Cranfield, Kuhn & Winter, 1982; Kuhn & Jones, 1982). This
is a biologically advantageous property. The release offluorideis governed
by the following expression:

Rate of release of fluoride = Af~* + B

The A term relates to a diffusion-controlled process and the B zero-order


term to an erosive process.
Kuhn & Jones (1982) examined various models for fluoride release and
showed that release did not fit the membrane and homogenous monolith
model. Instead, they concluded that the cement behaved as a porous
granular monolith, as described by Kydonieus (1980). The release of
fluoride appears to be an ion exchange phenomenon, as dental silicate
cement takes up rather than releases fluoride from solution if it is present
in sufficient concentration (Kuhn, Lesan & Setchell, 1983).
Fluoride release is biologically important. Since the early 1940s, fluoride
has been known to inhibit dental decay (Dean, Arnold & Elvove, 1942),
but the effect is not fully understood and several mechanisms have been
suggested (Levine, 1976). Tooth enamel, dentine and bone all possess a
mineral phase that has a hydroxyapatite-like substance which takes up
fluoride by replacement of the hydroxyl group (Hallsworth & Weatherall,
1969). In fact fluoride released by dental silicate cement is taken up by
adjacent tooth enamel (Halse & Hals, 1976). This apparently increases the
resistance of enamel to dissolution (McLundie & Murray, 1972) and, if the
acidogenic theory of caries is accepted (Levine, 1976) must have a
cariostatic effect. Maldonado, Swartz & Phillips (1978) have found that the
solubility of enamel in acid was reduced by 39 % when it was in contact
with a dental silicate cement. Moreover, the surface energy of fluorapatite
is lower than that of hydroxyapatite (Glanz, 1969) making the adhesion of
unwanted cariogenic substance, such as dental plaque, more difficult
(Rolla, 1977). For these reasons secondary caries is rarely observed under
dental silicate cement restorations (Bock, 1971; Hals, 1975) and in this
respect the cement is superior to composite resins and dental amalgams
(Updegraff, Change & Joos, 1971).
Aluminium ions released from the dental silicate cement are also
absorbed by hydroxyapatite and have a similar beneficial effect to that of
fluoride (Halse & Hals, 1976; Putt & Kleber, 1985). Thus, the dental
silicate cement confers protection against caries (dental decay) on
surrounding tooth material.

258
Dental silicate cement

Acid erosion
Dental silicate cement has always been regarded as suspect in service and
has a variable life-span (Paffenbarger, 1940; Wilson, 1969). It was early
suspected that this was associated with a susceptibility to acid attack
(Kulka, 1907; Rawitzer, 1909; Voelker, 1916b; Poetschke, 1916), a view
which has been confirmed by subsequent in vitro and in vivo work
(Norman, Swartz & Phillips, 1957, 1959; Jorgensen, 1963; Wilson &
Batchelor, 1968; Norman et al., 1969). Thus, although the cement is stable
in neutral media, such as normal saliva, it becomes progressively more
eroded under acid conditions (Wilson & Batchelor, 1968). This effect is
shown in Figure 6.23. Such conditions occur in stagnation regions of the
mouth. In these regions, dental plaque containing streptococci and

Figure 6.23 The effect of pH on the removal of ions from a dental silicate cement (Wilson &
Batchelor, 1968).

259
Phosphate bonded cements
lactobacilli degrades sugars, including plaque polysaccharides, to lactic
acid (Jenkins, 1965) and pH values as low as 4-0 have been recorded
(Stephan, 1940; Kleinburg, 1961). In fact preferential breakdown of dental
silicate cement has been observed in these regions (Henschel, 1949;
McLean & Short, 1969).
Tay et al. (1974, 1979) have studied the mechanism of erosion of the
dental silicate cement in service,findingthat grooving occurs at the margin
between the restoration and the tooth. Erosion exposes the cavity and
provides sites for the accumulation of food debris and bacteria which can
cause inflammation of the gingiva (Larato, 1971). It also leads to staining
of the restoration (Bock, 1971; Kent, Lewis & Wilson, 1973).
The extent of acid erosion depends on the nature of the acid; acids with
strong complexing function, such as citric acid, are particularly erosive
(Wilson & Batchelor, 1968; Stralfors & Eriksson, 1969). These acids are
found in citrus drinks.
Despite the failing of the dental silicate cement under acid conditions it
is more resistant to acid attack than all other dental cements with the
notable exception of the glass polyalkenoate cement (Norman, Swartz &
Phillips, 1959; Walls, McCabe & Murray, 1985; Beech & Bandyopadhyay,
1983; Kuhn, Setchell & Teo, 1984; Wilson et al., 1986a). These studies
have been confirmed by in vivo observations (Norman et al., 1969). A
clinical study carried out by Robinson (1971) over many years showed that
when carefully prepared and placed, the dental silicate cement was capable
of giving good performance. Many of the failures of this material must be
attributed to faulty preparation.

6.5.8 Biological aspects


The adverse pathological effects of dental silicate cement have been known
since Kulka (1911a,b). Since then many workers have observed that this
cement causes significant pulpal inflammation (McComb, 1982).
Manley (1936, 1943) reported that major histological changes occurred
in the pulp 24 hours after placing a silicate restoration, afindingconfirmed
by other workers (Zander, 1946; Brannstrom & Nyborg, 1960; Stanley,
Swerdlow & Buonocore, 1967; Qvist, 1975). The silicate cement also
inflames the gingiva (gum tissues) (Larato, 1971; Trivedi & Talim, 1973)
and demineralizes both dentine and enamel (Grieve, 1974).
At one time, irritation of the pulp was entirely attributed to the acidity

260
Dental silicate cement
of the cement which is most marked when it is freshly mixed (Manley,
1936; Harvey, Le Brocq & Rakowski, 1944; Roydhouse, 1961; Svare &
Meyer, 1965; Matsui et al., 1967). This theory is supported by observations
that tissue reactions become less marked with time (Svare & Meyer, 1965).
Moreover, it was observed that phosphoric acid penetrated dentine to a
considerable depth (Svare & Meyer, 1965; Swartz et al., 1968).
Despite this evidence, some workers have doubted this theory
(Roydhouse, 1961; Antonioli, 1969; Johnson et al., 1970). In particular,
Brannstrom and coworkers have strongly advocated an alternative theory,
that bacterial contamination causes pulpal damage (Brannstrom &
Nyborg, 1971; Bergvall & Brannstrom, 1971; Brannstrom & Vojinovic,
1976). They considered that virtually all the damage to pulp under a silicate
restoration was caused by bacterial infestation. This idea has been amply
confirmed by subsequent observations (Qvist, 1975; Brannstrom &
Nyborg, 1971;Mjor, 1977).
Watts (1979), while agreeing that bacterial contamination plays an
important role in causing irritation to tissues, showed that a silicate cement
even under germ-free conditions produced tissue damage. Of course, the
acidic dental silicate cement does not possess the antiseptic action of the
alkaline cements.
Cell culture tests show that dental silicate cement is strongly cytotoxic
- that is it severely damages cells - even after set (Spangberg et al., 1973).
This effect has been attributed to the hydrogen and fluoride ions present
(Helgeland & Leirskar, 1972, 1973; Tyas, 1979).
Another biological disadvantage is that dental silicate cement does not
bond to tooth material, and harmful substances and bacteria can percolate
between it and the tooth, giving rise to secondary caries and pulpal
irritation (Going, Massler & Dute, 1960). These effects are magnified when
dissolution of the cement occurs.
One advantageous biological property possessed by dental silicate
cement is the sustained release of fluoride; this has been discussed in
Section 6.5.7.

6.5.9 Conclusions
Dental silicate cement is solely used for restoring anterior (front) teeth. It
is probably the strongest purely inorganic cement and develops its strength
rapidly. Although satisfactory in areas of the mouth washed by saliva it is

261
Phosphate bonded cements
not quite up to the demands of more severe oral conditions. Its great
advantage is that it acts as a fluoride release agent and protects adjacent
tooth enamel.

6.5.10 Modified materials


A number of innovations made in the 1920s and 1930s may be noted.
Several attempts were made to reduce the dissolution of these cements in
oral fluids and their adverse effect on the pulp by inclusion of oils and
greases (Simon, 1929, 1932; Eberly, 1934). None have been considered
beneficial (Paffenbarger, Schoonover & Souder, 1938), a not surprising
result because the inclusion of hydrophobic substances is bound to
interfere in the setting of an aqueous cement.
Poetschke (1925) patented a dental silicate powder prepared by fusing
zinc silicate with calciumfluoride.This is a kind of silicophosphate cement
(Section 6.6). Thomsen (1931) attempted to formulate a water-setting
dental cement. Heynemann (1931) included lithium salts in the flux and
Brill (1935) included them in the liquid.
During this period a number of attempts at reinforcing these cements
were made. Fillers described include carborundum (Salzmann, 1930),
cellulosefibres(Schonbeck & Czapp, 1936) and even diamonds (Salzmann,
1930). None of these innovations found their way into commercial
materials (Paffenbarger, Schoonover & Souder, 1938).
More recently, Stanicioiu, Chinta & Hartner (1959) attempted to
reinforce the cement with glassfibres,but this was not successful. The most
serious study on the reinforcement of dental silicate cement was made by
J. Aveston (in Wilson et al., 1972). Silicon carbide whiskers, carbon fibres
and alumina powder were introduced into the cement mix. Unfortunately,
the glass powder/liquid ratio had to be reduced, and the strength gained by
reinforcement was thereby lost. It is clear that dental silicate cement cannot
be strengthened by fibre or particulate reinforcement.
Systematic attempts to formulate improved materials have met with no
success (Manly et al., 1951; Rockett, 1968). The last and, in some ways,
most promising attempt at improving the dental silicate cement was made
by Pendry (Pendry & Cook, 1972; Pendry, 1973) who improved its
resistance to acid by adding indium to both powder (5*8 %) and liquid
(5-65 %). The cement, however, lacked sufficient translucency, and by this
time the glass-ionomer cement had arrived with its advantages of
translucency and resistance to staining and acid attack.

262
Silicophosphate cement
Table 6.11. Specification properties of commercial luting silicophosphate
cements {Anderson & Pajfenbarger, 1962)

Value Specification limits"

Powder: liquid,6 g cm 3
2-00-2-70 n
Setting time (37 °C), 6-14
minutes
Film thickness 42-53 50
(2 min), urn
Compressivec 101-171 135 minimum
strength (24 h), MPa
Solubility & 0-7-2-3 1-0 maximum
disintegration (24 h), %

a
BS 3365/2: 1971 Specification for Dental Silicate Cement and Dental
Silicophosphate Cement. Part 2. Dental Silicophosphate Cement.
b
For a consistency spread of 25 mm diameter for 0-5 cm3 of cement paste under
a load of 220 gf applied after 2 minutes at 23 °C.
c
After storage for 24 hours in water at 37 °C.
n no specification test.

6.6 Silicophosphate cement


The silicophosphate cement has always been a minor and somewhat
obscure material. There have been no investigations into its setting
reaction and structure. Its chief uses are as a cementing agent and as a
temporary posterior filling material in dentistry. It can be regarded as a
hybrid of the dental silicate and zinc phosphate cements since the powder
is a physical mixture of an aluminosilicate glass and zinc oxide, the amount
of zinc oxide varying from 9-3 to 17*9% (Wilson, Crisp & Lewis,
1982).
The silicophosphate cement originated with the dental silicate cement,
for there is no doubt that early investigators experimented with mixtures of
aluminosilicate glass and zinc oxide (Fletcher, 1878,1879; Eberly, 1928). It
appears to have no particular advantages. As is often the case with hybrids,
it can combine the worst features as well as the best of the parents, and
often properties have intermediate values. Nevertheless, it continues to
have a small but persistent usage arising from its one advantage over the

263
Phosphate bonded cements

Table 6.12. Specification properties of commercial filling silicophosphate


cements {Wilson, 1975c)

Value Specification limits"

Powder: liquid,6 g cm 3
3-3-4-1 n
Setting time (37 °C), 3-5-5-75 3-6
minutes
Compressivec 179-209 170 minimum
strength (24 h), MPa
Solubility & 0-3-0-7 0-7 maximum
disintegration (24 h), %

a
BS 3365/2: 1971 Specification for Dental Silicate Cement and Dental
Silicophosphate Cement. Part 2. Dental Silicophosphate Cement.
b
For a consistency spread of 23 mm diameter for 0-5 cm3 of cement paste under
a load of 1-5 Kgf applied after 2 minutes at 23 °C.
c
After storage for 24 hours in water at 37 °C.
n no specification test.

zinc phosphate cement: it releasesfluorideand in certain clinical situations


this protection against dental caries is invaluable.
The flow properties are not as good as those of zinc phosphate cement
(Eames et ai, 1978; Hembree, George & Hembree, 1978) and film
thickness is greater (Table 6.11). Moreover, it does not have the
translucency of dental silicate cement (Wilson, 1975c).
The strength of silicophosphate cement lies between that of dental
silicate cement and zinc phosphate cement. Anderson & Paffenbarger
(1962) have reported properties of luting cements (Table 6.11) and Wilson
(1975c) those of filling materials (Table 6.12). Cameron, Charbeneau, &
Craig (1963) have confirmed these results. Housten & Miller (1968)
reported the properties for silicophosphate cements used for cementing
orthodontic brackets, where the consistency of the mix, and consequently
cement properties, lie between those of luting agents andfillingmaterials.
Silicophosphate cement acts as an agent for the sustained release of
fluoride, although different cements behave very differently (Wilson, Crisp
& Lewis, 1982). Silicophosphate cement has a durability in the mouth
similar to that of dental silicate cement. It is less resistant to oralfluidsthan
glass polyalkenoate cement, but more resistant than all other dental
cements, as is shown by both in vivo studies (Norman et aL, 1969; Ritcher
& Ueno, 1975; Clark, Phillips & Norman, 1977; Mitchem & Gronas, 1978;

264
References

Osborne et al., 1978) and the laboratory impinging jet method (Beech &
Bandyopadhyay, 1983; Wilson et ai, 1986a).
It is superior to the zinc phosphate cement for bonding orthodontic
bands to teeth (Clark, Phillips & Norman, 1977). It has greater durability
and there is less decalcification in adjacent tooth enamel. This latter
beneficial effect must arise from the release offluoridewhich is absorbed by
the enamel, so protecting it in a clinical situation where caries-producing
debris and plaque accumulate.

6.7 Mineral phosphate cements


Semler (1976) reported cements formed by reacting ground wollastonite
(CaSiO3) with phosphoric acid solution containing aluminium and zinc.
The setting times of these cements varied from 4 to 60 minutes. Best results
were obtained using a liquid of composition 69% H3PO4, 6-0% Zn,
2-0 % Al with a selected grade of wollastonite. The cement set in 4 minutes
and developed a compressive strength of 73 MPa in 24 hours compared
with a value of 20 MPa obtained with a fast-setting Portland cement. The
matrix of the wollastonite phosphate cement was observed to contain
silica, calcium and phosphate.
Serpentinite, Mg6Si4O10(OH)8, phosphate cements have been reported
by Ter-Grigorian et al (1982, 1984) and Zenaishvili, Bakradze & Chelidze
(1984). The reaction products are MgHPO4.3H2O and Mg(H2PO4)2,
which are transformed on heating to Mg2P2O7 and Mg3(PO4). Setting time,
strength and resistance to aqueous attack are all affected by the
concentration of phosphate used to prepare the cements. These materials
should be compared with the magnesium phosphate cements described in
Section 6.4.
Naturally occurring phosphate cements are also known (Krajewski,
1984).

References
Abdelrazig, B. E. I. & Sharp, J. H. (1985). A discussion of the papers on
magnesia-phosphate cement by T. Sugama and L. E. Kukacka. Cement &
Concrete Research, 15, 921-2.
Abdelrazig, B. E. I. & Sharp, J. H. (1988). Phase changes on heating ammonium
magnesium phosphate hydrates. Thermochimica Acta, 129, 197-215.
Abdelrazig, B. E. L, Sharp, J. H. & El-Jazairi, B. (1988). The chemical
composition of mortars made from magnesia-phosphate cement. Cement &
Concrete Research, 18, 415-25.

265
Phosphate bonded cements

Abdelrazig, B. E. L, Sharp, J. H. & El-Jazairi, B. (1989). Microstructure and


mechanical properties of mortars made from magnesia-phosphate cement.
Cement & Concrete Research, 19, 247-58.
Abdelrazig, B. E. I., Sharp, J. H., Siddy, P. A. & El-Jazairi, B. (1984). Chemical
reactions in magnesia-phosphate cements. Proceedings of the British Ceramic
Society, 35, 141-54.
Abraham. (1913). Pulpentod unter Silikatfullungen. Deutsche Monatsschrift fiir
Zahnheilkunde, 33, 532-6.
Abramovich, A., Macchi, R. L. & Ribas, L. M. (1976). Enamel corrosion
produced by zinc phosphate dental cement. Journal of Dental Research, 55,
107-10.
Akitt, J. W., Greenwood, N. N. & Lester, G. D. (1971). Nuclear magnetic
resonance and Raman studies of aluminium complexes formed in aqueous
solutions of aluminium salts containing phosphoric acids and fluoride ions.
Journal of the Chemical Society, A, 2450-7.
Allen, W. M., Sansom, B. F., Wilson, A. D., Prosser, H. J. & Groffman, D. M.
(1984). Release cements. British Patent GB 2, 123, 693 B.
Ames, W. B. (1892). Discussion - American Dental Association. Dental Cosmos,
34, 926-7.
Ames, W. B. (1893). Oxyphosphates. Dental Cosmos, 35, 869-75.
Anderson, J. N. & Paffenbarger, G. C. (1962). Properties of silicophosphate
cements. Dental Progress, 2, 72-5.
Andersson, K. R., Dent Glasser, L. S. & Smith, D. (1982). Polymerization and
colloid formation in silicate solutions. In Falcone, J. G. (ed.) Soluble Silicates.
ACS Symposium Series No. 194, Chapter 8. Washington, DC: American
Chemical Society.
Andrews, J. T. & Hembree, J. H. (1976). In vivo evaluation of the marginal
leakage of four inlay cements. Journal of Prosthetic Dentistry, 35, 532-7.
Ando, J., Shinada, T. & Hiraoka, G. (1974). Reactions of monoaluminium
phosphate with alumina and magnesia. Yogyo-Kyokai-Shi, 82, 644-9.
Antonioli, C. A. (1969). Etude sur la penetration des ions H + de differents
ciments de scellement a travers la dentine. Schweizische Monatsschrift fur
Zahnheilkunde, 79, 533-7.
Anzai, M., Hirose, H., Kikuchi, H., Goto, J., Azuma, F. & Higasaki, S. (1977).
Studies on soluble elements and solubility of dental cements. I. Solubilities of
zinc phosphate cement, carboxylate cement and silicate cement in distilled
water. Journal of the Nihon University School of Dentistry, 19, 26-39.
Arato, A. (1974). Physikalische Eigenschaften von verschiedenen Zementen fur
kieferothopadische Zwecke. Schweizische Monatsschrift fiir Zahnheilkunde,
54, 535-58.
Asch, W. & Asch, D. (1913). Silicates in Industry and Commerce, pp. 199-336.
London: Constable.
Aveston, J. (1965). The hydrolysis of the aluminium ion: ultracentrifugal and
acidity measurements. Journal of the Chemical Society, 4438-43.
Axelsson, B. (1964). Kemisk analys av dentala silikatcement och deras losliga
komponenter. Odontologisk Revy, 15, 150-68.

266
References

Axelsson, B. (1965). Kemisk analys av dentala zinkfosfatcement. Odontologisk


Revy, 16, 126-30.
Babin, J. B., Hurst, R. V. V. & Feary, T. (1978). Antibacterial activity of dental
cements. Journal of Dental Research, 37, Special Issue B, Abstract No. 214.
Beech, D. R. & Bandyopadhyay, S. (1983). A new laboratory method for
evaluating the relative solubility and erosion of dental cements. Journal of
Oral Rehabilitation, 10, 57-63.
Belopolsky, A. P., Shpunt, S. Ya. & Shulgina, M. N. (1950). Physicochemical
researches in the field of magnesium phosphates (the system MgO-P2O5-H2O
at 80 °C).Journal of Applied Science (USSR), 23, 873-84.
Bergvall, O. & Brannstrom, M. (1971). Measurement of the space between
composite resinfillingsand cavity walls. Swedish Dental Journal, 64,
217-26.
Bjerrum, N. & Dahm, C. R. (1931). Studies on aluminium phosphate. I.
Complex formation in acid solution. Zeitschrift fiir physikalische Chemie,
Bodenstein Festband 627-37 {Chemical Abstracts, 26, 666).
Bock, B. (1971). Klinische Nachuntersuchungen von Silikatzementfullungen.
Deutsche Zahndrtzliche Zeitschrift, 26, 665-71.
Bowen, R. L. (1962). Dentalfillingmaterial comprising vinyl silane treated fused
silica and a binder consisting of a reaction product of bisphenol and glycidyl
acrylate. US Patent 3,066,112.
Brannstrom, M. & Astrdm, A. (1972). The hydrodynamics of the dentine: its
possible relationship to dentinal pain. International Dental Journal, 22,
219-27.
Brannstrom, M. & Nyborg, H. (1960). Dentinal and pulpal responses.
Odontologisk Revy, 1, 37-50.
Brannstrom, M. & Nyborg, H. (1971). The presence of bacteria in cavities filled
with silicate cement and composite resin materials. Swedish Dental Journal,
64, 149-55.
Brannstrom, M. & Nyborg, H. (1974). Bacterial growth and pulpal changes and
inlays cemented with zinc phosphate cement and epoxylite CBA 9080. Journal
of Prosthetic Dentistry, 31, 556-65.
Brannstrom, M. & Vojinovic, O. (1976). Responses of the dental pulp to
invasion of bacteria around threefillingmaterials. Journal of Dentistry for
Children, 43, 83-9.
Brill, E. (1935). Improvements in or relating to dental cement. British Patent
430,624.
Brune, D. & Smith, D. (1982). Micro structure and strength properties of silicate
and glass-ionomer cements. Acta Odontologica Scandinavica, 40, 389-96.
Cameron, J. C, Charbeneau, G. T. & Craig, R. G. (1963). Some properties of
dental cements of specific importance in the cementation of orthodontic
bands. Angle Orthodontics, 33, 233-45.
Cartz, L., Servais, G. E. & Rossi, F. (1972). Surface structure of zinc phosphate
dental cements. Journal of Dental Research, 51, 1668-71.
Cassidy, J. E. (1977). Phosphate bonding then and now. American Ceramic
Society Bulletin, 56, 640-3.

267
Phosphate bonded cements

Charbeneau, G. T. (1961). Influence of specific surface on some properties of


silicate cement. Journal of Dental Research, 40, 763.
Clark, R. J., Phillips, R. W. & Norman, R. D. (1977). An evaluation of
silicophosphate cement. American Journal of Orthodontics, 71, 190-6.
Connick, R. E. & Poulsen, R. E. (1957). Nuclear magnetic resonance studies of
aluminium fluoride complexes. Journal of the American Chemical Society, 79,
5153-7.
Cranfield, M., Kuhn, A. T. & Winter, G. B. (1982). Factors relating to the rate
offluoride-ionrelease from glass-ionomer cement. Journal of Dentistry, 10,
333-41.
Crepaz, E. (1951). Constitution and reactivity against phosphoric acid of dental
cements. Chimica e Industria (Milan), 33, 137—40 (Chemical Abstracts, 45:
63211).
Crisp, S., Jennings, M. A. & Wilson, A. D. (1978). A study of temperature
changes occurring in setting dental cements. Journal of Oral Rehabilitation, 5,
139^4.
Crisp, S., O'Neill, I. K., Prosser, H. J., Stuart, B. & Wilson, A. D. (1978).
Infrared spectroscopic studies on the development of crystallinity in
dental zinc phosphate cements. Journal of Dental Research, 57,
245-54.
Crowell, W. S. (1929). Physical chemistry of dental cements. Journal of the
American Dental Association, 14, 1030-48.
van Dalen, E. (1933). Orienteerende onderzoekingen over tandcementen. Thesis.
Delft.
Darvell, B. W. (1984). Aspects of the chemistry of zinc phosphate cements.
Australian Dental Journal, 29, 2M-A.
Dean, H. T., Arnold, F. A. & Elvove, E. (1942). Domestic water and dental
caries. V. Public Health Reports (USA), 57, 1155-81.
Dimashkieh, M. R., Davies, E. H. & von Fraunhofer, J. A. (1974). A
measurement of the cement film thickness beneath full crown restorations.
British Dental Journal, 137, 281-4.
Dobrowsky, A. (1942). Uber die Herstellung eines neuen Zinkphosphatzementes
hochster Volumkonstanz. Die Chemische Technik, 15, 159-61.
Docking, A. R., Donnison, J. A., Newbury, C. R. & Storey, E. (1953). The
effect of orthodontic cement on tooth enamel. Australian Journal of Dentistry,
57, 139^9.
Dollimore, D. & Spooner, P. (1971). Sintering studies on zinc oxide.
Transactions of the Faraday Society, 67, 2750-9.
Dreschfeld, H. T. (1907). Chemistry of translucent cements. British Dental
Journal, 28, 1061-9.
Eames, W. B., Monroe, S. D., Roan, J. D. & O'Neal, S. J. (1977). Proportioning
and mixing of cements: a comparison of working times. Operative Dentistry,
2, 97-104.
Eames, W. B., O'Neal, S. J., Monteiro, J., Miller, C, Roan, R. D. & Cohen,
K. S. (1978). Techniques to improve the seating of castings. Journal of the
American Dental Association, 96, 432-7.

268
References

Earnshaw, R. (1960a). Investments for casting cobalt-chromium alloys. Part I.


British Dental Journal 108, 389-96.
Earnshaw, R. (1960b). Investments for casting cobalt-chromium alloys. Part II.
British Dental Journal 108, 429-40.
Eberly, J. A. (1934). Development of silicate cement tending to eliminate pulp
irritation. Dental Cosmos, 76, 419-24.
Eberly, N. E. (1928). Dental Cement, US Patent 1,671,104.
Eberly, N. E., Gross, C. V. & Crowell, W. S. (1920). The system zinc oxide,
phosphorus and water at 25° and 37°. Journal of the American Chemical
Society, 42, 1433-9.
Eichner, K., Lautenschlager, E. P. & von Radnoth, M. (1968). Investigation
concerning the solubility of dental cements. Journal of Dental Research, 47,
280-5.
El-Jazairi, B. (1982). Rapid repair of concrete pavings. Concrete, 12-15.
Ellison, S. & Warrens, C. (1987). Solid-state nmr study of aluminosilicate
glasses and derived dental cements. Report of the Laboratory of the
Government Chemist.
Elmore, K. L., Mason, C. M. & Christensen, J. H. (1946). Activity of
orthophosphoric acid in aqueous solutions at 25 °C from vapour pressure
measurements. Journal of the American Chemical Society, 68, 2528-32.
Eubank, W. R. (1951). Calcination studies of magnesium oxides. Journal of the
American Ceramic Society, 34, 225-9.
Evans, G. (1892). Discussion - American Dental Association. Dental Cosmos,
34, 927.
Finch, T. & Sharp, J. H. (1989). Chemical reactions between magnesia and
aluminium orthophosphate to form magnesia-phosphate cements. Journal of
Materials Science, 24, 4379-86.
Fleck, H. (1902). The chemistry of oxyphosphates. Dental Items of Interest,
906-35.
Fletcher, T. (1878). Compound for filling decayed teeth etc. British Patent 3028.
Fletcher, T. (1879). New filling. British Journal of Dental Science, 22, 74.
Frankel, E. (1913). Uber Silikatzement und Pulpatod. Deutsche Monatsschrift
fur Zahnheilkunde, 33, 529-32.
de Freitas, J. R. (1973). The long term solubility of a stannous fluoride-zinc
phosphate cement. Australian Dental Journal, 18, 167-73.
Gaylord, E. S. (1889). Oxyphosphates of zinc. Archives of Dentistry, 33, 364^80.
Genge, J. A. R., Holroyd, A., Salmon, J. E. & Wall, J. G. L. (1955). Phosphoric
acid as a complexing agent in ion-exchange chromatography. Chemistry in
Industry, 357-8.
Genge, J. A. R. & Salmon, J. E. (1959). Aluminium phosphates. Part III.
Complex formation between tervalent metals and orthophosphoric acid.
Journal of the Chemical Society, 1459-63.
Glanz, P.-O. (1969). On wettability and adhesiveness. Odontologisk Revy, 20,
Supplement 20.
Glasson, P. R. (1963). Reactivity of lime and related oxides. VIII. Production of
activated lime and magnesia. Journal of Applied Chemistry, 13, 111-19.

269
Phosphate bonded cements

Going, R. E., Massler, M. & Dute, H. L. (1960). Marginal penetration of dental


restorations as studied by crystal violet dye and 131 I. Journal of the American
Dental Association, 61, 285-300.
Greve, H. Chr. (1913). Theoretische und practische Studien iiber Zahnzemente.
Deutsche Monatsschrift fur Zahnheilkunde, 33, 346-58.
Grieve, A. R. (1974). The effect of silicate cement on enamel and dentine.
Scandinavian Journal of Dental Research, 82, 510-16.
Griffith, J. R. & Cannon, R. W. S. (1974). Cementation - materials and
techniques. Australian Dental Journal, 19, 93-9.
Gulabivala, K., Setchell, D. J. & Davies, E. H. (1987). An application of the jet
test method for the evaluation of disintegration of dental luting cements in
marginal gaps analogous to those of crowns and bridges. Clinical Materials,
2, 221-30.
Gursin, A. V. (1965). A study of the effect of stannous fluoride incorporated in
a dental cement. Journal of Oral Therapeutics and Pharmacology, 1, 630-6.
Halla, F. & Kutzeilnigg, A. (1933). Zur Kenntnis des Zinkphosphatzements.
Zeitschrift fur Stomatologie, 31, 177-81.
Hallsworth, A. S. & Weatherall, J. A. (1969). The microdistribution, uptake and
loss of fluoride in human enamel. Caries Research, 3, 109-18.
Hals, E. (1975). Histology of natural secondary caries associated with silicate
cement restorations. Archives of Oral Biology, 20, 291-6.
Halse, A. & Hals, E. (1976). Electron probe microanalysis of secondary carious
lesions adjacent to silicate fillings. Calciferous Tissues Research, 21, 183-93.
Hannah, C. M. & Smith, D. C. (1971). Tensile strengths of selected dental
restorative materials. Journal of Prosthetic Dentistry, 26, 314-23.
Harper, F. C. (1967). Effect of calcination temperature on the properties of
magnesium oxides for use in magnesium oxychloride cements. Journal of
Applied Chemistry, 17, 5-10.
Harvey, W., Le Brocq, L. F. & Rakowski, L. (1944). The acidity of dental
cements. British Dental Journal, 11, 61-9.
Helgeland, K. & Leirskar, J. (1972). A further testing of the effect of dental
materials on growth and adhesion of animal cells in vitro. Scandinavian
Journal of Dental Research, 80, 206-12.
Helgeland, K. & Leirskar, J. (1973). Silicate cement in a cell culture system.
Scandinavian Journal of Dental Research, 81, 251-9.
Hembree, J. H., George, T. A. & Hembree, M. E. (1978). Film thickness of
cements beneath complete crowns. Journal of Prosthetic Dentistry, 39, 533-5.
Henschel, C. J. (1943). The effect of mixing surface temperatures on dental
cementation. Journal of the American Dental Association, 30, 1582-9.
Henschel, C. J. (1949). Observations concerning the in vivo disintegration of
silicate cement restorations. Journal of Dental Research, 28, 528.
Heynemann, H. (1931). Verfahren zur Herstellung von Zahncementen. German
Patent 520,138.
Higashi, S., Nakazawa, S., Yamanaka, A., Jyoshin, K. & Yamaga, R. (1963).
On the properties of the hydraulic zinc phosphate cement during and after the
setting process. Journal of the Osaka University Dental School, 3, 77-88.

270
References

Higashi, S., Morimoto, K., Satomura, A., Horie, K., Anzai, M., Kimura, K.,
Takeda, K. & Miyazima, T. (1969a). Studies on water-settable cements. Part
9. Examination of the physical properties of water-settable tertiary zinc
phosphate. Journal of the Nikon University School of Dentistry, 11, 60-4.
Higashi, S., Morimoto, K., Satomura, A., Horie, K., Anzai, M., Oitate, M.,
Hirata, Y. & Haga, H. (1969b). Studies on water-settable cements. Part 10.
Manipulation of the newly made A type zinc phosphate cement and its
compressive strength. Journal of the Nihon University School of Dentistry, 11,
105-8.
Higashi, S., Yamaga, R., Satomura, A., Anzai, M., Nishijima, K., Kubo, M. &
Kawakami, R. (1972). Studies on water-settable cements. Part 11. Preparatory
experiments on water-settable phosphate cement and properties of water-
settable primary calcium phosphate. Journal of the Nihon University School of
Dentistry, 14, 58-63.
Hinkins, J. E. & Acree, S. F. (1901). The disintegration of cement fillings. Dental
Cosmos, 43, 581-97.
Hoard, R. J., Caputo, A. A., Contino, R. M. & Koenig, M. E. (1978).
Intracoronal pressure during crown cementation. Journal of Prosthetic
Dentistry, 40, 520-30.
Holroyd, A. & Salmon, J. E. (1956). Ion-exchange studies of phosphates. Part I.
Ion-exchange and pH-titration for detection of complex formation. Journal of
the Chemical Society 269-72.
Housten, W. J. B. & Miller, M. W. (1968). Cements for orthodontic use. The
Dental Practitioner, 19, 104-9.
Her, R. K. (1979). The polymerization of silica. The Chemistry of Silica. New
York: Wiley-Inter science Chapters 3 and 6.
Jameson, R. F. & Salmon, J. E. (1954). Aluminium phosphates: Phase-diagram
and ion-exchange studies of the system aluminium oxide-phosphoric
oxide-water at 25 °C. Journal of the Chemical Society, 4013-17.
Jendresen, M. D. (1973). New dental cements and fixed prosthodontics. Journal
of Prosthetic Dentistry, 30, 684-8.
Jenkins, G. N. (1965). The equilibrium between plaque and enamel in relation to
dental caries. In Wolstenholme, G. E. W. & O'Connor, M. (eds.) Caries
Resistant Teeth, pp. 192-210. London: Churchill.
Johnson, R. H., Christensen, G. J., Stigers, R. W. & Laswell, H. R. (1970).
Pulpal irritation due to the phosphoric acid component of silicate cement.
Oral Surgery, 29, 447-54.
Jorgensen, K. D. (1960). Factors affecting the film thickness of zinc phosphate
cements. Ada Odontologica Scandinavica, 18, 479-90.
Jorgensen, K. D. (1963). On the solubility of silicate cements. Acta Odontologica
Scandinavica, 21, 141-58.
Jorgensen, K. D. & Peterson, G. F. (1963). The grain size of zinc phosphate
cements. Acta Odontologica Scandinavica, 21, 255-70.
Kato, K., Shiba, M., Nakamura, M. & Ariyoshi, T. (1976). Report of the
Institute of Medical and Dental Engineering (Tokyo Medical and Dental
University), 10,45-61.

271
Phosphate bonded cements

Kendzior, G. M., Leinfelder, K. F. & Hershey, H. G. (1976). The effect of cold


temperature mixing on the properties of a zinc phosphate cement. Angle
Orthodontics, 46, 345-50.
Kent, B. E., Fletcher, K. E. & Wilson, A. D. (1970). Dental silicate cements: XI.
Electron probe studies. Journal of Dental Research, 49, 86-92.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1971a). Dental silicate cements:
XIII. The crazing and dulling of the surface. Journal of Dental Research, 50,
393-9.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1971b). Dental silicate cements:
XIV. Crazing, cement properties and liquid composition. Journal of Dental
Research, 50, 400-4.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1973). The properties of a
glass-ionomer cement. British Dental Journal, 135, 322-6.
Kent, B. E. & Wilson, A. D. (1968). Unpublished observations.
Kent, B. E. & Wilson, A. D. (1969). Dental silicate cements: VIII. Acid-base
aspect. Journal of Dental Research, 48, 412-18.
Kent, B. E. & Wilson, A. D. (1971). Dental silicate cements: XV. Effect of the
particle size of the powder. Journal of Dental Research, 50, 1616-20.
Kingery, W. D. (1950a). Fundamental study of phosphate bonding in
refractories. I. Literature review. Journal of the American Ceramic Society, 33,
239-41.
Kingery, W. D. (1950b). Fundamental study of phosphate bonding in
refractories. II. Cold setting properties. Journal of the American Ceramic
Society, 33, 242-7.
Kleinburg, I. (1961). Studies on dental plaque. I. The effect of different
concentrations of glucose on the pH of dental plaque in vivo. Journal of
Dental Research, 40, 1087-110.
Komrska, J. & Satava, V. (1970). Die chemischen Prozesse bei der Abbindung
von Zinkphosphatzementen. Deutsche Zahndrtzliche Zeitschrift, 25, 914-21.
Krajewski, K. P. (1984). Early diagenetic phosphate cements in the albian
condensed glauconitic limestone of the Tatra mountains, Western
Carpathians. Chemical Abstracts, 10, 114382.
Kuhn, A. T. & Jones, M. P. (1982). A model for the dissolution and fluoride
release from dental cements. Biomaterials, Medical Devices and Artificial
Organs, 10, 281-93.
Kuhn, A. T., Lesan, W. & Setchell, D. (1983). Some further factors affecting the
solubility of silicate cements. Journal of Materials Science Letters, 2, 191-4.
Kuhn, A. T., Setchell, D. J. & Teo, C. K. (1984). An assessment of the jet
method for solubility measurements of dental cements. Biomaterials, 5, 161-8.
Kuhn, A. T., Tan, W. K., Davies, E. H. & Winter, G. B. (1982). Water uptake
and loss in silicate cements. Journal of Materials Science Letters, 17, 3263-7.
Kuhn, A. T. & Wilson, A. D. (1985). The dissolution mechanisms of silicate and
glass-ionomer dental cements. Biomaterials, 6, 378-82.
Kuhn, A. T., Winter, G. B. & Tan, W. K. (1982). Dissolution rates of silicate
cements. Biomaterials, 3, 136-44.
Kulka, M. (1907). Ueber die wichtigsten mechanischen and einige chemische

272
References

Eigenschaften der Silikat- und Zinkphosphatzemente. O ester reichisch-


Ungarische Vierteljahrschrift fiir Zahnheilkunde, 23, 568-602.
Kulka, M. (1911a). The possible chemical or pathological effects of cement
filling. Dental Items of Interest, 33, 360-74.
Kulka, M. (1911b). Ueber die Moglichkeit chemischer bzw. pathologischer
Wirkungen von Zementfiillungen. Oesterreichisch-Ungarische
Vierteljahrschrift fiir Zahnheilkunde, 27, 43-60.
Kydonieus, A. F. (1980). Controlled Release Technologies. West Palm Beach,
Florida: CRC Press.
Larato, D. C. (1971). Influence of silicate cements on gingiva. Journal of
Prosthetic Dentistry, 26, 186-8.
Laswell, H. R., Welk, D. A., Dilts, W. E. & Thompson, J. A. (1971). Evaluation
of a water activated zinc phosphate dental cement. Journal of the Tennessee
Dental Association, 51, 14—17.
Levine, R. S. (1976). The action of fluoride in caries prevention. A review of
current concepts. British Dental Journal, 140, 9-14.
Lloyd, C. H. & Mitchell, L. (1984). The fracture toughness of tooth coloured
restorative materials. Journal of Oral Rehabilitation, 11, 257-72.
McComb, D. (1982). Tissue reactions to silicate, silicophosphate, glass ionomer
cements and restorative materials. In Smith, D. C. & Williams, D. F. (eds.)
Biocompatibility of Dental Materials. Volume III. Biocompatibility of Dental
Restorative Materials, Chapter 4. Boca Raton: CRC Press Inc.
McLean, J. W. & von Fraunhofer, J. A. (1971). The estimation of cement film
thickness by an in vivo technique. British Dental Journal, 131, 107-11.
McLean, J. W. & Short, I. G. (1969). Composite anterior filling materials. A
clinical and physical approach. British Dental Journal, 111, 9-18.
McLundie, A. C. & Murray, F. D. (1972). Silicate cements and composite resins
- a scanning electron microscope study. Journal of Prosthetic Dentistry, 27,
544-51.
Maldonado, A., Swartz, M. L. & Phillips, R. W. (1978). An in vitro study of
certain properties of a glass-ionomer cement. Journal of the American Dental
Association, 96, 785-91.
Manley, E. B. (1936). A preliminary investigation into the reaction of the pulp
to various filling materials. British Dental Journal, 50, 321-31.
Manley, E. B. (1943). Pulp reaction to dental cements. Proceedings of the Royal
Society of Medicine, 36, 488-99.
Manly, R. S., Baker, C. F., Miller, P. N. & Welch, F. E. (1951). The effect of the
composition of liquid and powder on the physical properties of silicate
cements. Journal of Dental Research, 30, 145-56.
Manly, R. S. & Harrington, D. P. (1959). Solution rate of tooth enamel in an
acetate buffer. Journal of Dental Research, 38, 910-19.
Manston, R. & Gleed, P. T. (1985). Reaction cements as materials for the
sustained release of trace elements into the digestive tract of cattle and sheep.
II. Release of cobalt and selenium. Journal of Veterinary Pharmacology
Therapeutics, 8, 374-81.
Manston, R., Sansom, B. F., Allen, W. M., Prosser, H. J., Groffman, D. M.,

273
Phosphate bonded cements

Brant, P. J. & Wilson, A. D. (1985). Reaction cements as materials for the


sustained release of trace elements into the digestive tract of cattle and sheep.
I. Copper release. Journal of Veterinary Pharmacology Therapeutics, 8,
368-73.
Matkovic, B., Popvic, S., Rogic, V., Zunic, T. & Young, J. F. (1977). Reaction
products in magnesium oxychloride cement pastes. System
MgO-MgCl 2 -H 2 O. Journal of the American Ceramic Society, 60, 504-7.
Matsui, A., Buonocore, M. G., Sayegh, F. & Yamaki, M. (1967). Reactions to
implants of conventional and new dental restorative materials. Journal of
Dentistry for Children, 34, 316-22.
Mesu, F. P. (1982). Degradation of luting cements measured in vitro. Journal of
Dental Research, 61, 665-72.
Mesu, F. P. & Reedijk, T. (1983). Degradation of luting cements measured in
vitro and in vivo. Journal of Dental Research, 62, 1236-40.
Miller, W. D. (1881). Concerning oxyphosphate cements. Dental Cosmos, 33,
14-22.
Mitchem, J. C. & Gronas, D. G. (1978). Clinical evaluation of cement solubility.
Journal of Prosthetic Dentistry, 40, 453-6.
Mitchem, J. C. & Gronas, D. G. (1981). Continued evaluation of the clinical
solubility of luting cements. Journal of Prosthetic Dentistry, 45, 289-91.
Mjor, I. A. (1977). Histological demonstration of bacteria subjacent to dental
restorations. Scandinavian Journal of Dental Research, 85, 169-74.
Morgenstern, M. (1905). Untersuchungen der Silikat- und
Zinkphosphatzemente unter besonderer Berucksichtigung ihrer physikalischen
Eigenschaften. Oesterreichisch-Ungarische Vierteljahrschrift filr Zahnheilkunde,
21, 514-36.
Morris, J. H., Perkins, P. G., Rose, A. E. A. & Smith, W. E. (1977). The
chemistry and binding properties of aluminium phosphates. Chemical Society
Reviews, 6, 173-94.
Myers, C. L., Drake, J. T. & Brantley, W. A. (1978). A comparison of
properties for zinc phosphate cements mixed on room temperature and frozen
slabs. Journal of Prosthetic Dentistry, 40, 409-12.
Nakazawa, S., Jyoshin, K., Tsutsumi, S. & Yamaga, R. (1965). Studies on the
hydraulic zinc phosphate cement containing Mg(H 2 PO 4 ) 2 • 2H 2 O. Journal of
the Osaka University Dental School, 5, 61-71.
Neiman, R. & Sarma, A. C. (1980). Setting and thermal reactions of phosphate
investments. Journal of Dental Research, 59, 1478-85.
Norman, R. D., Swartz, M. L. & Phillips, R. W. (1957). Studies on the
solubility of certain dental materials. Journal of Dental Research, 36,
977-85.
Norman, R. D., Swartz, M. L. & Phillips, R. W. (1959). Additional studies on
the solubility of certain dental materials. Journal of Dental Research, 38,
1025-37.
Norman, R. D., Swartz, M. L., Phillips, R. W. & Sears, R. (1970). Properties of
cements mixed from liquids with altered water content. Journal of Prosthetic
Dentistry, 24, 410-17.

274
References

Norman, R. D., Swartz, M. L., Phillips, R. W. & Virmani, R. (1969). A


comparison of the intraoral disintegration of three dental cements. Journal of
the American Dental Association, 78, 777-82.
O'Hara, M. J., Duga, J. J. & Sheets, H. D. (1972). Studies in phosphate
bonding. American Ceramic Society Bulletin, 51, 590-5.
0ilo, G. (1978). Extent of slits at the interface between luting cements and
enamel dentin and amalgam. Acta Odontologica Scandinavica, 36, 257-61.
0ilo, G. (1988). Characterization of glass-ionomer filling materials. Dental
Materials, 4, 129-33.
0ilo, G. & Espevik, S. (1978). Stress/strain behaviour of some dental luting
cements. Acta Odontologica Scandinavica, 36, 45-9.
O'Neill, I. K., Prosser, H. J., Richards, C. P. & Wilson, A. D. (1982). NMR
spectroscopy of dental materials. I. 31P studies on phosphate-bonded cement
liquids. Journal of Biomedical Materials Research, 16, 39-49.
O'Reilly, D. E. (1960). NMR chemical shifts of aluminium: experimental data
and variational calculation. Journal of Chemical Physics, 32, 1007-12.
Osborne, J. W., Swartz, M. L., Goodacre, C. J., Phillips, R. W. & Gale, E. M.
(1978). A method for assessing the clinical solubility and disintegration of
luting cements. Journal of Prosthetic Dentistry, 40, 413-17.
Paddon, J. M. & Wilson, A. D. (1976). Stress relaxation studies on dental
materials. 1. Dental cements. Journal of Dentistry, 4, 183-9.
Paffenbarger, G. C. (1940). Silicate cements: an investigation by a group of
practising dentists under the direction of the ADA Research Fellowship at the
National Bureau of Standards. Journal of the American Dental Association,
27, 1611-22.
Paffenbarger, G. C, Schoonover, I. C. & Souder, W. (1938). Dental silicate
cements: Physical and chemical properties and a specification. Journal of the
American Dental Association, 25, 32-87.
Paffenbarger, G. C, Sweeney, S. J. & Isaacs, A. (1933). A preliminary report on
zinc phosphate cements. Journal of the American Dental Association, 20,
1960-82.
Palmer, S. B. (1891). Zinc phosphates. Dental Cosmos, 33, 364-80.
Peirce, C. B. (1879). Filling materials of oxide of zinc and glacial phosphoric
acid. Dental Cosmos, 21, 696-7.
Pendry, N. R. (1973). Siliceous dental cement. British Patent 1,314,090.
Pendry, N. R. & Cook, D. J. (1972). Improvements in or relating to dental
cement. British Patent 1,270,195.
Pennsylvania Association of Dental Surgeons. (1879). Dental Cosmos, 21,
695-7.
Phillips, R. W. (1975). Polymeric material for anterior restorations. In von
Fraunhofer, J. A. (ed.) Scientific Aspects of Dental Materials, Chapter 7.
London and Boston: Butterworths.
Plant, C. G., Jones, I. H. & Wilson, H. J. (1972). Setting characteristics of lining
and cementing materials. British Dental Journal, 133, 21-4.
Plant, C. G. & Tyas, M. (1970). Lining materials with special reference to
Dropsin. British Dental Journal, 128, 486-91.

275
Phosphate bonded cements

Plant, C. G. & Wilson, H. J. (1970). Early strength of lining materials. British


Dental Journal 129, 269-74.
Pluim, L. J. & Arends, J. (1981). In vivo solubility of dental cements. Journal of
Dental Research, 60, 1213, Abstract No. 61.
Pluim, L. J. & Arends, J. (1987). The relationship between salivary properties
and in vivo solubility of dental cements. Dental Materials, 3, 13-18.
Pluim, L. J., Arends, J. & Havinga, P. (1985). Comparison of in vivo and in vitro
solubility of 6 luting cements. Journal of Dental Research, 64, 714, Abstract
No. 98.
Pluim, L. J., Arends, J., Havinga, P., Jongebloed, W. J. & Stokroos, I. (1984).
Quantitative cement solubility experiment in vivo. Journal of Oral
Rehabilitation, 11, 171-9.
Poetschke, P. (1916). Physical properties of dental cements. Journal of Industrial
& Engineering Chemistry, 8, 302-9.
Poetschke, P. (1925). Dental cement. US Patent 1,552,341.
Popovics, S., Rajendran, N. & Penko, M. (1987). ACI Materials Journal, 84,
64-73.
Powers, J. M., Farah, J. W. & Craig, R. G. (1976). Modulus of elasticity and
strength properties of dental cements. Journal of the American Dental
Association, 92, 588-91.
Powis, D. R., Prosser, H. J., Shortall, A. C. & Wilson, A. D. (1988). Long-term
monitoring of microleakage of composites. Part I: radiochemical diffusion
technique. Journal of Prosthetic Dentistry, 60, 304-7.
Powis, D. R., Prosser, H. J. & Wilson, A. D. (1988). Long-term monitoring of
microleakage of dental cements by radiochemical diffusion. Journal of
Prosthetic Dentistry, 59, 651-7.
Proell. (1913). Experimentelle Untersuchungen iiber die Ursache des
Zahnpulpatodes unter Silikaphosphatzementen nebst theoretisch-praktischen
Studien iiber Zemente und andere Fullmaterialen. Deutsche Monatsschrift filr
Zahnheilkunde, 33, 81-122 (English abstract: Experimental investigations
regarding the cause of death of the dental pulp under silicate cements with
theoretical and practical studies on cements and other filling materials. The
Dental Cosmos, 750).
Prosen, E. M. (1939). Refractory material for use in making dental casting. US
Patent 2,152,152.
Prosen, E. M. (1941). Refractory material suitable for use in casting dental
investments, models, etc. US Patent 2,209,404.
Prosser, H. J., Wilson, A. D., Groffman, D. M., Brookman, P. J., Allen, W. M.,
Gleed, P. T., Manston, R. & Sansom, B. F. (1986). The development of
acid-base reaction cements as formulations for the controlled release of trace
elements. Biomaterials, 7, 109-12.
Putt, M. S. & Kleber, C. J. (1985). Dissolution studies of human enamel treated
with aluminium solutions. Journal of Dental Research, 64, 437-40.
Qvist, V. (1975). Pulp reactions in human teeth to tooth coloured filling
materials. Scandinavian Journal of Dental Research, 83, 54-66.
Rawitzer, J. (1908). Wie miissen Silicatzemente zusammengezetzt sein damit ein

276
References

Absterben der Pulpa unter ihnen vollkommen ausgeschlossen ist?


Correspondenz-Blatt fur Zahndrzte, 37, 340-5.
Rawitzer, J. (1909). Chemie der Silikatphosphatzemente. Deutsche Monatsschrift
fur Zahnheilkunde, 27, 269-81.
Ray, K. W. (1934). The behavior of siliceous cements. Journal of the American
Dental Association, 21, 237-51.
Ray, N. H. (1979). The structure and properties of inorganic polymeric
phosphates. British Polymer Journal, 11, 163-77.
Richter, W. A. & Ueno, H. (1975). Clinical evaluation and dental clinical
durability. Journal of Prosthetic Dentistry, 33, 294-7.
Robinson, A. D. (1971). The life of a filling. British Dental Journal, 130, 206-8.
Rockett, T. J. (1968). Experimental silicate cements. International Association of
Dental Research, 46th General Meeting, Abstract No. 433.
Rolla, G. (1977). Effects of fluoride on initiation of plaque formation. Caries
Research, 11, 243-61.
Rollins, W. H. (1879). A contribution to the knowledge of cements. Dental
Cosmos, 21, 574-6.
Rostaing di Rostagni, C. S. (1878). Verfahrung zur Darstellung von Kitten fur
zahnarztliche und ahnliche Zwecke, bestehend von Gemischen von
Pyrophosphaten des Calciums oder Bariums mit den Pyrophosphaten des
Zinks oder Magnesiums. German Patent 6015 (Berlin). Also in (1881)
Correspondenz-Blatt fur Zahndrzte, 10, 67-9.
Roydhouse, R. H. (1961). Silicate cements and acid production. Journal of
Dental Research, 40, 258-63.
Salmon, J. E. & Terrey, J. H. (1950). The system zinc oxide-phosphoric
acid-water at temperatures between 25° and 100°. Journal of the Chemical
Society, 2813-24.
Salmon, J. E. & Wall, J. G. L. (1958). Aluminium phosphates. Part II. Ion-
exchange and pH-titration studies of aluminium phosphate complexes in
solution. Journal of the Chemical Society, 1128-34.
Salzmann, W. (1930). Verbesserung von Zahnzementen. German Patent 553,425.
Sanderson, R. S. (1908). Silicate cements. The Dental Record, 28, 74-6.
Savignac, J. R., Fairhurst, C. W. & Ryge, G. (1965). Strength, solubility and
disintegration of zinc phosphate cement with clinically determined
powder/liquid ratio. Angle Orthodontics, 35, 126-30.
Schoenbeck, F. (1908). Process for the production of tooth cement. US Patent
897,160.
Schonbeck, F. & Czapp, E. (1936). Verfahren zur Herstellung von
Zahnzementen. German Patent 635,310.
Selvaratnam, M. & Spiro, M. (1965). Transference numbers of orthophosphoric
acid and the limiting equivalent conductance of the H2PO~ ion in water at
25 °C. Transactions of the Faraday Society, 61, 360-73.
Semler, C. E. (1976). A quick-setting wollastonite phosphate cement. American
Ceramic Society Bulletin, 55, 983-8.
Servais, G. E. & Cartz, L. (1971). Structure of zinc phosphate dental cement.
Journal of Dental Research, 50, 613-20.

277
Phosphate bonded cements

Sidler, P. & Strub, J. R. (1983). In-vivo Untersuchung der Loslichkeit und des
Abdichtungsvermogens von drei Befestigungszementen. Deutsche
Zahndrztliche Zeitschrift, 38, 564-71.
Simmons, F. F., D'Anton, E. W. & Hudson, D. C. (1968). Property studies of a
hydrophosphate cement. Journal of the American Dental Association, 76,
337-9.
Simon, O. (1929). Improvements in and relating to dental cements. British
Patent 333,325.
Simon, O. (1932). Process of producing cements for teeth. US Patent 1,886,982.
Skibell, R. B. & Shannon, I. L. (1973). Addition of stannous fluoride to
orthodontic cement. International Journal of Orthodontics, 11, 131-5.
Skinner, E. W. & Phillips, R. W. (1960). Silicate cement. In The Science of
Dental Materials, Chapter 15. Philadelphia and London: W. B. Saunders
Company.
Smith, D. C. (1977). Past, present and future of dental cements. In Craig, R. G.
(ed.) Dental Materials Review. Ann Arbor: University of Michigan Press.
Smith, D. C. (1982). Composition and characteristics of dental cements. In
Smith, D. C. & Williams, D. F. (eds.) Biocompatibility of Dental Materials.
Volume II. Biocompatibility of Preventive Dental Materials and Bonding
Agents, Chapter 8. Boca Raton: CRC Press Inc.
Sorrell, C. A. & Armstrong, C. R. (1976). Reactions and equilibria in
magnesium oxychloride cements. Journal of the American Ceramic Society, 59,
51-4.
Spangberg, L., Rodrigues, H., Langeland, L. & Langeland, K. (1973). Biological
effects of dental materials. 2. Toxicity of anterior tooth restorative materials
on HeLa cells in vitro. Oral Surgery, Oral Medicine, Oral Pathology, 36,
713-24.
Stanicioiu, Gh., Chinta, Gh. & Hartner, Al. (1959). Possibilities of improving
silicate cements. Stomatologia, 6, 369-73 {Chemical Abstracts, 54, 5983g).
Stanley, H. R., Swerdlow, H. & Buonocore, M. G. (1967). Pulp reactions to
anterior restorative materials. Journal of the American Dental Association, 75,
132-41.
Steenbock, P. (1903). Verfahren zur Herstellung einer durch Behandlung mit
Phosphorsauren und deren Salzen erhartenden Masse. Kaiserliches Patent
174,558.
Steenbock, P. (1904). Improvements in and relating to the manufacture of a
material designed for the production of cement. British Patent 15,181.
Steinke, R., Newcomer, P., Komarneni, S. & Roy, R. (1988). Dental cements:
investigation of chemical bonding. Materials Research Bulletin, 23, 13-22.
Stralfors, A. & Eriksson, S. E. (1969). The rate of dissolution of dental silicate
cement. Odontologisk Tidskrift, 11, 185-210.
Stephan, R. M. (1940). Changes in the hydrogen-ion concentration on tooth
surfaces and in carious lesions. Journal of the American Dental Association,
27, 718-23.
Sudakas, G. L., Turkina, L. I. & Chernikova, A. A. (1982). Properties of
phosphate binders. Chemical Abstracts, 96, 202,472.

278
References

Sugama, T. & Kukacka, L. E. (1983a). Magnesium monophosphate cements


derived from diammonium phosphate solutions. Cement & Concrete Research,
13, 407-16.
Sugama, T. & Kukacka, L. E. (1983b). Characteristics of magnesium
polyphosphate cements derived from ammonium polyphosphate solutions.
Cement & Concrete Research, 13, 499-506.
Svare, C. W. & Meyer, M. W. (1965). Available acidity of silicate cements.
Journal of the American Dental Association, 70, 354-61.
Sveshnikova, V. N. & Zaitseva, S. N. (1964). Aluminophosphates as
polyelectrolytes. Russian Journal of Inorganic Chemistry, 9, 672-5.
Swanson, E. W. (1936). Effect of particle size on the physical properties of
silicate cements. Journal of the American Dental Association, 23, 1620-31.
Swartz, M. L., Niblack, B. F., Alter, E. A., Norman, R. D. & Phillips, R. W.
(1968). In vivo studies on the penetration of dentin by constituents of silicate
cement. Journal of the American Dental Association, 76, 573-8.
Sychev, M. M., Medvedeva, I. N., Biokov, V. A. & Krylov, O. S. (1982). Effect
of reaction kinetics and morphology of neoformation on the properties of
phosphate cements based on magnesium titanates. Chemical Abstracts, 96,
222252e.
Takeda, S. et al. (1979). Thermal changes of binder in phosphate-bonded
investment. Shika Igaku, 42, 429-36.
Tay, W. M., Cooper, I. R., Morrant, G. A., Borlace, H. R. & Bultitude, F. W.
(1979). An assessment of anterior restorations in vivo using the scanning
electron microscope. Results after three years. British Dental Journal, 146,
71-6.
Tay, W. M , Morrant, G. A., Borlace, H. R. & Bultitude, F. W. (1974). An
assessment of anterior restorations in vivo using the scanning electron
microscope. Results after one year. British Dental Journal, 137, 463-71.
Ter-Grigorian, M. S., Beriya, V. V., Zedginidze, E. N. & Sychev, M. M. (1984).
Problem of the setting of serpentinite-phosphate cement. Chemical Abstracts,
101, 156469.
Ter-Grigorian, M. S., Zedginidze, E. N., Sychev, M. M., Papuashvili, S. M.,
Teideishvili, L. K. & Dateshidze, R. I. (1982). Study of serpentinite-phosphate
cement during heat treatment at 110-1200 °C. Chemical Abstracts, 96, 23920.
Theuniers, G. (1984). Een onderzock naar een duurzame afdichting door kroon-
en brugcementen. Thesis. Leuvan.
Thomsen, J. C. (1931). Dental cement and process making the same. US Patent
1,792,200.
Trivedi, S. C. & Talim, S. T. (1973). The response of human gingiva to
restorative materials. Journal of Prosthetic Dentistry, 29, 73-80.
Tuenge, R., Sugel, I. A. & Izutsu, K. T. (1978). Physical properties of a zinc
phosphate cement prepared on a frozen slab. Journal of Dental Research, 57,
593-6.
Tyas, M. J. (1979). The effects of silicate cement on mitochondria and lyosomes
of cultured cells assessed by quantitative enzyme histochemistry. Journal of
Oral Rehabilitation, 6, 55-60.

279
Phosphate bonded cements

Updegraff, D. M., Change, R. W. H. & Joos, R. W. (1971). Antibacterial


activity of dental restorative materials. Journal of Dental Research, 50, 382-7.
Van Wazer, J. R. (1958). Properties and Chemistry of Phosphorus and its
Compounds, pp. 486-91. New York: Interscience Publishers Inc.
Van Wazer, J. R. & Callis, C. F. (1958). Metal complexing by phosphates.
Chemical Reviews, 58, 1011^6.
Vashkevich, N. K. & Sychev, M. M. (1982). Setting of copper phosphate
cements in the presence of organic polymers. Chemical Abstracts, 97,
23921.
Vieira, D. F. & De Arujo, P. A. (1963). Estudo a cristizacao de cemento de
fosfato de zinco. Revista da Faculdade Odontologia da Universidade de Sao
Paolo, 1, 127-31.
Voelker, C. C. (1916a). The place of silicates in dentistry. The Dental Summary,
36, 177-200.
Voelker, C. C. (1916b). Dental silicate cements in theory and practice. The
Dental Cosmos, 36, 1098-111.
Vysotskii, Z. Z., Galinskaya, V. I., Kolychev, V. I., Strelko, V. V. & Strazhesko,
D. N. (1974). The role of polymerization and depolymerization reactions of
silicic acid. In Strazhesko, D. N. (ed.) Adsorption and Adsorbents, vol. 1, p. 75.
New York: Wiley.
Walls, A. W. G., McCabe, J. F. & Murray, J. J. (1985). An erosion test for
dental cements. Journal of Dental Research, 64, 1100-4.
Ware, A. L. (1971). Properties of cements for orthodontic bonding. Australian
Orthodontics Journal, 2, 254-61.
Waters, D. N. & Henty, M. S. (1977). Raman spectra of aqueous solutions of
hydrolysed aluminium(III) salts. Journal of the Chemical Society: Dalton
Transactions, 243-5.
Watts, A. (1979). Bacteria contamination and the toxicity of silicate and zinc
phosphate cements. British Dental Journal, 146, 7-13.
Watts, A. S. (1915). Dental porcelains. Transactions of the American Ceramic
Society, 17, 190-9.
Wege. (1908). Zur Frage betr. die Ursache des Absterbens der Pulpa unter
Silikatzementen, sowie einige Worte iiber Phenakit. Deutsche Zahndrztliche
Wochenschrift 346-58.
Wei, S. H. Y. & Sierk, D. L. (1971). Fluoride uptake by enamel from zinc
phosphate cement containing stannous fluoride. Journal of the American
Dental Association, 83, 621-4.
Williams, J. I., Gates, G. L., Hembree, J. H. & MacKnight, J. P. (1979). The
frozen-aluminium-slab mixing technique: its effect on zinc phosphate cements.
Journal of Dentistry for Children, 46, 398-403.
Williams, P. D. & Smith, D. C. (1971). Measurement of the tensile strength of
dental restorative materials by use of a diametral compressive strength test.
Journal of Dental Research, 50, 436-42.
Wilson, A. D. (1969). A survey of dental practice in the use of silicate cements.
Ministry of Technology Report. British Dental Journal, 127, 7 (abstract).
Wilson, A. D. (1975a). Dental cements - general. In von Fraunhofer, J. A. (ed.)

280
References

Scientific Aspects of Dental Materials, Chapter 4. London and Boston:


Butterworths.
Wilson, A. D. (1975b). Zinc oxide dental cements. In von Fraunhofer, J. A. (ed.)
Scientific Aspects of Dental Materials, Chapter 5. London and Boston:
Butterworths.
Wilson, A. D. (1975c). Dental cements based on ion-leachable glasses. In von
Fraunhofer, J. A. (ed.) Scientific Aspects of Dental Materials, Chapter 6.
London and Boston: Butterworths.
Wilson, A. D. (1976). Specification test for the solubility and disintegration of
dental cements: a critical evaluation of its meaning. Journal of Dental
Research, 55, 721-9.
Wilson, A. D. (1978). The chemistry of dental cements. Chemical Society
Reviews, 7, 265-96.
Wilson, A. D., Abel, G. & Lewis, B. G. (1974). The solubility and disintegration
test for zinc phosphate dental cements. British Dental Journal, 137, 313-17.
Wilson, A. D., Abel, G. & Lewis, B. G. (1976). The 'solubility and
disintegration' test for zinc phosphate dental cements: the use of small
specimens. Journal of Dentistry, 4, 28-32.
Wilson, A. D. & Batchelor, R. F. (1967a). Dental silicate cements. I. The
chemistry of erosion. Journal of Dental Research, 46, 1075-85.
Wilson, A. D. & Batchelor, R. F. (1967b). Dental silicate cements. II.
Preparation and durability. Journal of Dental Research, 46, 1425-32.
Wilson, A. D. & Batchelor, R. F. (1968). Dental silicate cements. III.
Environment and durability. Journal of Dental Research, 47, 115-20.
Wilson, A. D., Crisp, S. & Lewis, B. G. (1982). The aqueous erosion of
silicophosphate cements. Journal of Dentistry, 10, 187-97.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986a). An
evaluation of the significance of the impinging jet method for measuring the
acid erosion of dental cements. Biomaterials, 1, 55-60.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986b). A study
of variables affecting the impinging jet method for measuring the erosion of
dental cements. Biomaterials, 1, 217-20.
Wilson, A. D. & Kent, B. E. (1968). Dental silicate cements. V. Electrical
conductivity. Journal of Dental Research, 44, 463-70.
Wilson, A. D. & Kent, B. E. (1970a). Dental silicate cements. IX.
Decomposition of the powder. Journal of Dental Research, 49, 7-13.
Wilson, A. D. & Kent, B. E. (1970b). Dental silicate cements. X. The
precipitation reaction. Journal of Dental Research, 49, 21-6.
Wilson, A. D. & Kent, B. E. (1971). The glass-ionomer cement, a new
translucent cement for dentistry. Journal of Applied Chemistry and
Biotechnology, 21, 313.
Wilson, A. D., Kent, B. E. & Batchelor, R. F. (1968). Dental silicate cements.
IV. Phosphoric acid modifiers. Journal of Dental Research, 47, 233-43.
Wilson, A. D., Kent, B. E., Batchelor, R. F., Scott, B: G. & Lewis, B. G.
(1970a). Dental silicate cements. XII. The role of water. Journal of Dental
Research, 49, 307-14.

281
Phosphate bonded cements

Wilson, A. D., Kent, B. E., Clinton, D. & Miller, R. P. (1972). The formation
and microstructure of the dental silicate cement. Journal of Materials Science,
7, 220-38.
Wilson, A. D., Kent, B. E. & Lewis, B. G. (1970). Zinc phosphate cements:
chemical study of in vitro durability. Journal of Dental Research, 49,
1049-54.
Wilson, A. D., Kent, B. E., Mesley, R. F., Miller, R. P. Clinton, D. & Fletcher,
K. E. (1970b). Formation of dental silicate cement. Nature, 225, 272-3.
Wilson, A. D. & Lewis, B. G. (1980). The flow properties of dental cements.
Journal of Biomedical Materials Research, 14, 383-91.
Wilson, A. D. & Mesley, R. F. (1968). Dental silicate cements. VI. Infrared
studies. Journal of Dental Research, 47, 644-52.
Wilson, A. D., Paddon, J. M. & Crisp, S. (1979). The hydration of dental
cements. Journal of Dental Research, 58, 1065-71.
Windeler, A. S. (1978). The use of film thickness to measure working time of
zinc phosphate cements. Journal of Dental Research, 57, 697-701.
Windeler, A. S. (1979). Powder enrichment effects on film thickness of zinc
phosphate. Journal of Prosthetic Dentistry, 42, 299-303.
Wisth, P. J. (1972). The ability of zinc phosphate and hydrophosphate cements
to seal space bands. Angle Orthodontics, 42, 395-8.
Worner, H. K. (1940). The properties of commercial zinc phosphate cements.
Australian Journal of Dentistry, 44, 123-41.
Worner, H. K. & Docking, A. R. (1958). Dental materials in the tropics.
Australian Dental Journal, 3, 215-29.
Wright, J. W. (1919). A study of some dental cements. Journal of Dental
Research, 1, 35-60.
Wygant, J. F. (1958). Cementitious bonding in ceramic fabrication. In Kingery,
W. D. (ed.) Ceramic Fabrication Processes, pp. 171-88. New York: John
Wiley & Sons.
Yamano, C. (1968). Effect of NaF-phosphate cement on enamel of human
tooth. Journal of the Osaka University Dental School, 13, 123-37.
Yamazaki, T. & Takeuchi, M. (1967). The Chemical Society of Japan, Industrial
Chemistry Section, 10, 656.
Zander, H. A. (1946). The reaction of dental pulps to silicate cements. Journal of
the American Dental Association, 33, 1233^43.
Zenaishvili, N. V., Bakradze, E. G. & Chelidze, D. V. (1984). Serpentinite-
phosphate mortar containing iron and boron. Chemical Abstracts, 101,
156519r.
Zhuravlev, V. F., Volfson, S. L. & Sheveleva, B. I. (1950). The processes that
take place in the roasting of zinc-phosphate dental cement. Journal of Applied
Chemistry (USSR), 23, 121-8.

282
7 Oxysalt bonded cements

7.1 Introduction
Oxysalt bonded cements trace their origin to studies by Sorel in the third
quarter of the nineteenth century. The first of these cements which he
studied was zinc oxychloride (Sorel, 1855). Later he described a series of
magnesia-based cements which included both the magnesium oxychloride
and magnesium oxysulphate types (Sorel, 1867). Of this second group, the
magnesium oxychlorides in their hydrated form have been shown to have
larger values of modulus of elasticity, microhardness and compressive
strength than does Portland cement for a wide range of porosites (Beaudoin
& Ramachandran, 1975). The magnesium oxysulphate cements have
properties that have led to their being considered for nuclear applications,
since they have good fire resistance, low thermal conductivity and above
all, in marked contrast to the related oxychloride cements, no potential to
initiate corrosion of the reinforcing steel (Beaudoin & Ramachandran,
1978). These cements are also employed as binders in lightweight panels, in
insulating materials and in architectural applications (Urwongse & Sorrell,
1980b).
Oxysalt bonded cements are formed by acid-base reactions between a
metal oxide in powdered solid form and aqueous solutions of metal
chloride or sulphate. These reactions typically give rise to non-homo-
geneous materials containing a number of phases, some of which are
crystalline and have been well-characterized by the technique of X-ray
diffraction. The structures of the components of these cements and the
phase relationships which exist between them are complex. However, as
will be described in the succeeding parts of this chapter, in many cases there
is enough knowledge about these cements to enable their properties and
limitations to be generally understood.

283
Oxysalt bonded cements

7.1.1 Components of oxysalt bonded cements


The three major types of oxysalt bonded AB cement are the zinc
oxychloride, the magnesium chloride and the magnesium oxysulphate
cements. The bases employed, therefore, are either zinc oxide or mag-
nesium oxide, both of which readily undergo hydration in aqueous
solution, behaving as M(OH)2 species and acting as a source of hydroxyl
ions. They are thus both clearly bases in the Bronsted-Lowry sense.
By contrast, the acidity of the metal salts used in these cements has a less
clear origin. All of the salts dissolve quite readily in water and give rise to
free ions, of which the metal ions are acids in the Lewis sense. These ions
form donor-acceptor complexes with a variety of other molecules,
including water, so that the species which exists in aqueous solution is a
well-characterized hexaquo ion, either Mg(OH2)g+ or Zn(OH2)g+. How-
ever, zinc chloride at least has a ternary rather than binary relationship
with water and quite readily forms mixtures of ZnO-HCl-H2O (Sorrell,
1977). Hence it is quite probable that in aqueous solution the metal salts
involved in forming oxysalt cements dissolve to generate a certain amount
of mineral acid, which means that these aqueous solutions function as
acids in the Bronsted-Lowry sense.

7.1.2 Setting of oxysalt bonded cements


The nature of the solidification process in these cements has received little
attention. Rather, attention has focussed on the crystalline components
that form in cements which have been allowed to equilibrate for some
considerable time; the nature of such phases is now quite well understood.
Gelation is reasonably rapid for these cements and occurs within a
significantly shorter time than does development of crystalline phases. The
conclusion may be drawn that initial cementition is not the same as
crystallization, but must occur with the development of an essentially
amorphous phase. Reactions can continue in the amorphous gelled phase,
but are presumably limited in speed by the low diffusion rates possible
through such a structure. However, reactions are able to proceed
substantially to completion, since in many cases X-ray diffraction has
demonstrated almost quantitative conversion of the parent compounds
to complex crystalline mixed salts, though several days or weeks of
equilibration are required to bring this about.

284
Zinc oxychloride cements
7.2 Zinc oxychloride cements
7.2.1 History
These cements were the earliest of the oxysalt bonded cements to be
prepared (Sorel, 1855) and their chemistry has been the subject of
numerous investigations over the years. There are considerable difficulties
associated with such investigations. Not only does the cement contain a
complex mixture of different crystalline precipitates but it is unaffected by
boiling water and dissolves only slowly in strong acids. Consequently
separation or analysis of any of the phases which may be present is difficult.
Nonetheless, as early as 1925 at least 17 crystalline compounds were
claimed to occur in the zinc oxychloride cement (Mellor, 1925).
Of the early studies, two are still of value in providing an outline of the
possible phase compositions and equilibria existing in this material. The
first, by Droit (1910), concentrated on measuring the solubility of zinc
oxide in zinc chloride solutions at 18 °C, while the second, by Holland
(1930), was based on the analysis of saturated solutions and moist residues
equilibrated at 25 °C and 50 °C. Droit assigned the compositions
4ZnO. ZnCl2. 6H2O (4:1:6) and 2ZnO. 2ZnCl2. 3H2O (2:2:3 or 1:1:1-5)
to solids which he found in equilibrium. The former phase he described as
amorphous and reported that five of the six water molecules were lost at
200 °C.The last remaining molecule of water was not lost until a much
higher temperature, and then both HC1 and ZnCl2 were lost as well. Droit
described the 1:1:1-5 phase as microcrystalline and reported that it lost
one of its water molecules at 230 °Cwhile the remaining water, together
with HC1, was lost at a higher temperature.
Holland, by contrast, reported the existence of three well-defined phases
in this system, corresponding to ZnO:ZnCl2:H2O ratios of 5:1:8, 1:1:1
and 1:1:2 respectively (Holland, 1930). More recently it was pointed out
that these claims lack any unequivocal support such as X-ray charac-
terization of the phases, so that their validity must remain in doubt
(Sorrell, 1977).
A number of other workers have reported the existence of an additional
phase, corresponding to 4:1:5 (Feitknecht, 1930; Hayek, 1932; Aspelund,
1933), which may or may not be associated with a 1:1:1 phase. Feitknecht
(1933) also carried out a detailed study of a phase described as 4:1:4 using
X-ray diffraction and concluded that the material had a layer structure
with interspersed water molecules. Such a structure would permit the
addition and removal of water molecules without altering the interlayer

285
Oxysalt bonded cements
distance. Hence all of the phases based on a 4:1 ratio of ZnO to ZnCl2 may
be closely related and readily interconverted, depending on the precise
conditions of cementition.
Droit's original 4:1:5 phase has been studied by X-ray diffraction
(Nowacki & Silverman, 1961, 1962) and found to have a rhombohedral
layer structure. The 1:1:1 phase was also found to have a layer structure,
which consisted of pseudohexagonal layers of zinc atoms separated by
ordered layers comprising oxygen and chlorine atoms (Feitknecht, Ostwald
& Forsberg, 1959). This fundamental structure was apparently found for
both of the crystalline modifications in which this phase has been found to
occur, namely the monoclinic and the orthorhombic (Sorrell, 1977).

7.2.2 Recent studies


Sorrell (1977) further elucidated the structure and phase relationships in
zinc oxychloride cements and produced a phase diagram for the system
(Figure 7.1). Sorrell prepared his solutions of aqueous zinc chloride in one
of two ways: either by dissolving reagent-grade ZnCl2 in distilled water, or
by dissolving zinc metal cut from an ingot in aqueous hydrochloric acid
followed by boiling to low volume to remove the excess acid. Cement
samples were prepared by reacting either aqueous HC1 or aqueous ZnCl2
solutions with zinc oxide powder and sealing them in polyethylene
containers to equilibrate for at least four days before examining them.
Cements were analysed by X-ray diffractometry using Cu Ka radiation.
Powdered cement samples were also examined by quantitative differential
thermal gravimetry (DTG) from 25 °C to 715 °C at a heating rate of 10 °C
per minute.
X-ray analysis of the various samples that were produced indicated that
the system ZnO-ZnCl2-H2O includes four crystalline phases, two of
which, ZnO and ZnCl2.1-5H2O, are essentially the starting materials.
Sorrell also found the 4:1:5 phase, reported by Droit, with an identical X-
ray powder diffraction pattern to that reported by Nowacki & Silverman
(1961, 1962), and a 1:1:2 phase. Since neither the 1:1:2 nor the 4:1:5
phase lost or gained weight on exposure to air at about 50% relative
humidity and 22 °C and no changes developed in the X-ray diffraction
pattern following this exposure, he concluded that the previously reported
1:1:1 phase cannot be formulated from mixtures of ZnO and aqueous
ZnCl2.

286
Zinc oxychloride cements
The cementition reaction between zinc oxide powder and aqueous zinc
chloride was found to be both rapid and extremely exothermic. Although
at least four days equilibration was allowed before examining any of the
cements in detail, Sorrell found evidence that reaction was complete within
20 to 30 minutes and occurred without observable development of
intermediate phases. He also found that, as the concentration of reactants
was increased, so reaction rate increased until, at sufficiently high
concentrations, reaction occurred too quickly to allow proper mixing of
the reactants. Preheating the zinc oxide at 900 °C for 16 hours was found
to slow the reaction down, but only slightly.
Sorrell showed that the two discrete phases could be readily and
reversibly interconverted. For example, the 1:1:2 phase was found to react

ZnO

90,
O ZnO ond ZnClg Solution*

• ZnO and MCI Solution!

10 20 30 40 50 60 60 90
H2O ZnCU
Wt%ZnCI2
Figure 7.1 Phase relationships in the ZnO-ZnQ2-H2O system at room temperature (Sorrell,
1977).

287
Oxysalt bonded cements

with water to generate the 4:1:5 phase so rapidly that it was not possible
to record an X-ray diffraction pattern for the mixture. Samples whose
composition fell into the triangle bounded by the phases 4:1:5,1:1:2 and
pure water dried in air to give a solid mixture of the two discrete phases. By
contrast, samples containing less ZnO did not dry out. Thus all changes
caused by variation in composition proved to be predictable from the
phase diagram and at no time were there any departures from this, such as
formation of Zn(OH) 2 .
Thermogravimetric analysis was carried out on both the 4:1:5 and the
1:1:2 phases (Sorrell, 1977). The latter was found to undergo a two-step
dissociation beginning at about 230 °C. Above 275 °C, the melting point of
ZnCl 2 , weight loss corresponding to removal of half of the constituent
water occurred. Above c. 680 °C weight loss ceased, having reached a level
corresponding to complete loss of ZnCl 2 and water. Attempts were made
to complement this thermogravimetric study by X-ray analysis on the
quenched samples but these were unsuccessful due to the deliquescent
nature of the dissociation products. Above 275 °C, however, the sample
existed as a solid mixed with a liquid and this crystallized on cooling as it
absorbed water from the atmosphere. The X-ray patterns which were made
shortly after quenching showed the presence of zinc oxide, and repeated
scanning indicated a rapid reaction to regenerate the 1:1:2 phase.
The 4:1:5 phase was shown by thermogravimetric analysis to dissociate
at about 160 °C to zinc oxide and the 1:1:2 phase, a process which was
verified using X-ray diffraction (Sorrell, 1977). Once the 1:1:2 phase was
formed it underwent characteristic dissociation at temperatures above
160 °C.
The equilibrium relationships found by Sorrell (1977) were valid only for
room temperature (22 + 2 °C) and, because samples were allowed to cure in
sealed containers, for equilibrium water vapour pressures determined by
the assembly of phases present. The phases which exist under such
conditions were quite unequivocally found to be 4:1:5 and 1:1:2. However
Sorrell pointed out that it is entirely possible that lower hydration states of
either phase could be stable at higher temperatures or lower humidities. In
particular the 4:1:4 phase (Feitknecht, 1933) may well be such a phase,
particularly as one of the five waters of hydration is known to be held only
loosely in the structure. Indeed, Sorrell reported that he observed a slight
shoulder on the larger dehydration peak of the DTG curve of the 4:1:5
phase that might be assigned to the loss of this first water molecule. He did
not, however, succeed in isolating or characterizing a 4:1:4 phase.

288
Zinc oxychloride cements
Similarly Sorrell (1977) had no success in attempts to isolate the 1:1:1
phase reported by Forsberg & Nowacki (1959), though he conceded that it,
too, might exist as a lower hydration state of his well-defined 1:1:2
phase. Unfortunately Forsberg and Nowacki did not provide details of
the alleged 1:1:1 phase, so comparison of their results with SorrelPs was
not possible.
To summarize these findings, the thermal decomposition of the two
ternary phases was found to be complex (Sorrell, 1977). On the basis of
thermogravimetric data alone, the loss of water appeared to occur in such
a way that both a 4:1:4 and a 1:1:1 phase might result. However, what
actually occurred was that a mixture of ZnCl2 and ZnO was obtained as a
residue; such a result is typical of what is found for aqueous solutions of
divalent metal chlorides that have been evaporated to dryness. An
additional complexity was the possible reoxidation of zinc as the mixtures
containing ZnCl2 were evaporated, thus leading to larger amounts of ZnO
in the residue than were originally present in the original mixture. Overall,
the thermogravimetric studies gave results that were extremely difficult to
interpret, largely because of the highly complicated set of relationships that
can exist in this system.
In very dilute HC1 solutions, specifically those with a pH above 5-48, the
4:1:5 phase was found to be insoluble. By contrast, addition of
concentrated HC1 to the 4:1:5 phase was shown to lead to formation of the
1:1:2 phase (Sorrell, 1977). Below 35wt% HC1, the 4:1:5 phase was
found to dissolve congruently. Since the 1:1:2 phase was also found to
dissolve congruently in hydrochloric acid solutions with concentrations
above 23 wt %, it follows that there is a range of concentrations over which
both phases are soluble in aqueous HC1. This behaviour explains why the
zinc oxychlorides have proved to be unsatisfactory in attempts to use them
as dental cements. The preparation of such cements from concentrated
aqueous solutions of ZnCl2 results in the formation either of the 1:1:2
phase alone or of mixtures of the 4:1:5 and 1:1:2 phases, neither of which
is stable in the presence of water. Preparing dental cements from less
concentrated solutions also results in the formation of mixed phases,
unless the bulk composition has excessive amounts of ZnO present. In
these latter cases the cement stability is acceptable but it lacks both a
workable consistency and a reasonable working time.
SorrelFs study of zinc oxychloride cement (1977), in addition to making
an important contribution to our understanding of the nature of this
material, also highlighted a more general feature of the chemistry of zinc

289
Oxysalt bonded cements
chloride. Since the 4:1:5 phase separated from dilute solutions, the
ZnCl2-H2O system was shown not to be binary. Instead, it represents a
section through the ZnO-HCl-H2O ternary system. Moreover, zinc
chloride is known to be one of the most soluble of all substances and is very
difficult to prepare in a completely anhydrous condition (Greenwood &
Earnshaw, 1984). This feature also presumably derives from the nature of
the phase relationships that develop in the ZnO-HCl-H2O system.

7.3 Magnesium oxychloride cements


7.3.1 Uses
Magnesium oxychloride cements are widely used for the fabrication of
floors. They find application for this purpose because of their attractive
appearance, which resembles marble, and also because of their acoustic
and elastic properties and their resistance to the accumulation of static
charge. They have also been used for plastering walls, both interior and
exterior; for exterior walls the cement often includes embedded stone
aggregate (Sorrell & Armstrong, 1976). However, there have been
problems with this latter application, since the base cement has been found
to be dimensionally unstable and, in certain circumstances, to release
corrosive solutions and show poor weather resistance.

7.3.2 Calcination of oxide


The quality of magnesium oxychloride cements is highly dependent on the
reactivity of the magnesium oxide used in their preparation. Typically,
such oxides are prepared by calcination of the basic carbonate (Eubank,
1951; Harper, 1967), but their reactivity varies according to the conditions
under which such calcination is carried out. As the reactivity alters so does
the amount of oxide that can be incorporated into a cement relative to the
amount of aqueous MgCl2 (Harper, 1967).
A detailed study of the effect of calcination conditions on properties of
the resulting oxide was carried out by Harper (1967), with the results
shown in Table 7.1 Oxides prepared from basic carbonate at 600 °C and
700 °C give weak cements because the reactivity of the oxide is so high that
reaction occurs during mixing and the cement is broken in the process. The
low strengths of these cements contrasted with the much higher strengths
obtained from oxides that had been calcined at higher temperatures. For

290
Magnesium oxychloride cements
Table 7.1. Compressive strengths of magnesium oxy chloride cements made
from basic carbonate (Harper, 1967)

Compressive strength, MPa


Calcination MgO/MgCl 2 , mo 1
temperature, Age,
°C days 6 7 8 9 10

7 7-6
600 14 7-4
28 8-3
7 15-9
700 14 16-3
28 17-3
7 49-6 55-3 57-4
800 14 55-9 631 63-5
28 57-5 60-0 50-0
7 47-7 58-2 67-5 73-8 78-1
900 14 48-6 68-6 691 71-7 690
28 55-6 68-8 65-8 67-8 66-6
7 60-5 71-7 77-8 73-6 80-0
1000 14 65-9 85-7 87-5 87-0 68-4
28 83-2 90-4 80-6 88-2 88-5

example, oxides prepared at 1000 °C were much less reactive, which


allowed time for adequate mixing with the aqueous magnesium chloride
and thus allowed the strength to develop to a greater extent. Such cements
went on increasing their strength for at least 28 days after mixing. These
lower-reactivity oxides could also be added in greater quantities to the
MgCl2 solution, thus allowing for greater variation in the composition of
the cement.

7.3.3 Setting chemistry


There have been a number of studies aimed at understanding the chemistry
of the curing and setting of magnesium oxychloride cements and at
identifying the phases that are present in the final material. Investigations
in the first half of the twentieth century revealed that cement formation in
the MgO-MgCl2-H2O system involves gel formation and crystallization of

291
Oxysalt bonded cements
ternary oxychloride phases of uncertain composition (Robinson &
Waggaman, 1909; Bury & Davies, 1932). In the years around 1950, the
system was studied by Walter-Levy and coworkers; they succeeded in
identifying crystalline phases and determining the structure of one of them
(de Wolff & Walter-Levy, 1953). Walter-Levy had previously recognized
an oxychlorocarbonate phase in the cement which forms by reaction with
carbon dioxide in the atmosphere (Walter-Levy, 1937, 1938). A very
extensive study of this system was carried out by Cole and coworkers
(Demediuk, Cole & Hueber, 1955; Cole & Demediuk, 1955), who were
able to define the temperature ranges over which the various phases are
stable, though they confined their studies to cements of low MgO content,
e.g. 1-5 g MgO in 75 cm3 MgCl2 solution. Below 100 °C two ternary phases
were found, one with an Mg(OH) 2 : MgCl 2 : H 2 O composition of 5:1:8, the
other with a compositon of 3:1:8. These phases have been referred to as
the 5-form and the 3-form respectively. The 3-form was the reaction
product of MgO with solutions of higher MgCl2 concentration than those
which tended to yield the 5-form (Cole & Demediuk, 1955). Both forms
were found to change gradually with time via slow reaction with
atmospheric CO 2 , so that after long periods a basic magnesium carbonate,
corresponding to a composition of Mg(OH) 2 . 2MgCO 3 . MgCl 2 . 6H 2 O,
had formed (Cole & Demediuk, 1955). Hence the results previously
reported by Walter-Levy (1937, 1938) were confirmed. This basic car-
bonate was found to be much less soluble in water than either of the two
oxychloride phases.
Above 100 °C, a different set of phases was stable in the simple
magnesium oxychloride system (Demediuk, Cole & Hueber, 1955; Cole
& Demediuk, 1955); a 2-form 2Mg(OH) 2 . MgCl 2 . 4H 2 O and a 9-form
9Mg(OH) 2 . 5H2O were found to occur. Apart from identifying these
phases, these workers were not able to give details on the structures.
The phase relationships prevailing in this system have received more
attention at temperatures below 100 °C, and there is greater understanding
of the equilibria which occur in this temperature range. These equilibria
involve a range of solids, namely Mg(OH) 2 , the 5:1:8 and 3:1:8 materials
and MgCl 2 . 6H 2 O, depending on the range of concentrations of MgCl2
used in aqueous solution and the ratio of solution to MgO powder
employed in fabricating the cement.
Of all the studies on this system, that by Sorrell & Armstrong (1976) has
provided the most useful information, both on the phase relationships and
on the kinetics of interconversion. They used three different grades of

292
Magnesium oxychloride cements
MgO, produced by different but well-defined routes and having different
reactivities towards aqueous MgCl2; in this way, it was possible to study
the cementation reactions in some detail and to ensure a reasonably close
approach to equilibrium.
To study reaction kinetics, cement batches of total mass 300 g were
prepared using ingredients measured to the nearest 0-1 g. Mixing was
carried out for 10 minutes using a kitchen blender, after which specimens
were cast in slabs 10 x 10 x 1-2-1-5 cm in polyethylene moulds. When the
setting reaction had proceeded to a sufficient extent and viscosity had risen
to give a reasonably stiff paste, a small portion was removed, placed on a
glass microscope slide and immediately examined by X-ray diffraction.
The remainder of the sample was allowed to set.

7.3.4 Kinetics of cementation


Sorrell & Armstrong formulated cements in proportions corresponding to
the 5:1:8 and 3:1:8 compositions. The initial mixtures were thick slurries
with no observable tendency to separate provided a sufficiently reactive
oxide was used. They tended to set within about 90 minutes, at which time
samples were prepared for X-ray determination. Initially, although the
preliminary hardening process was apparently complete, the only crys-
talline phase that could be found was MgO; moreover, this material was
found in amounts that approximated to the quantity in the initial mixture.
After some two hours, the X-ray diffraction pattern corresponded to
either the 5:1:8 or the 3:1:8 phase; warming of the sample had also
occurred. Growth of the crystalline oxychloride phases continued rapidly
up to about 15 hours, and more slowly thereafter, until after four days
there was no trace of MgO in the diffraction pattern of the cement.
The fact that the initial setting process for magnesium oxychloride
cements takes place without observable formation of either the 5:1:8 or
the 3:1:8 phase is important. It indicates that formation of an amorphous
gel structure occurs as the first step, and that crystallization is a secondary
event which takes place from what is effectively a supersaturated solution
(Urwongse & Sorrell, 1980a). This implies that crystallization is likely to be
extremely dependent upon the precise conditions of cementition, including
temperature, MgO reactivity, heat build-up during reaction and purity of
the components in the original cement mixture.
The reactivity of the magnesium oxide in particular was shown to be
crucial. A relatively unreactive batch gave thin slurries which showed some

293
Oxysalt bonded cements

tendency to settle and took at least 20 hours to set. Reaction was not
complete after 14 days and in the mixture corresponding to the 5:1:8
composition it was the 3:1:8 phase which actually crystallized out.

7.3.5 Phase relationships in the MgO-MgCl2-H2O system


Sorrell & Armstrong (1976) prepared cement slabs in proportions
corresponding to the 5:1:8 phase from each of the three magnesium oxide
samples that were available. The samples were allowed to set for at least
30 days exposed to air, after which time the phase gradient was determined
by X-ray diffraction of surfaces exposed by incremental grinding. These
experiments revealed that the most reactive sample of MgO had yielded a
material consisting essentially of the 5:1:8 phase uniformly distributed
throughout the bulk and containing no more than 2 % residual MgO. The
less reactive samples of oxide, by contrast, gave much less homogeneous
cements in which the surface layer consisted mainly of unreacted MgO
together with MgCl2 solution. The amount of crystalline 5:1:8 phase
increased with depth into the sample, thus demonstrating the development
of phase gradients for these materials.
For studies of the phase equilibria, 20-gram batches of cement were
prepared by mixing appropriate solutions of MgCl2 with the most reactive
of the three samples of MgO that was available (Sorrell & Armstrong,
1976). Mixing was carried out by stirring with a glass rod and cements were
sealed in polyethylene containers for at least four days to allow sufficient
time for equilibration. The use of relatively small samples avoided the
problems of large temperature increases as the cementition reaction
occurred. The studies of phase equilibria resulted in compilation of the
phase diagram Figure 7.2.
The portion of the phase diagram representing the MgCl2-rich compo-
sitions received little attention in the work of Sorrell & Armstrong (1976),
but was the subject of a separate study reported a few years later
(Urwongse & Sorrell, 1980a). Since it had previously been shown that the
MgO-MgCl 2 -H 2 O system is a portion of the MgO-HCl-H 2 O system
(Robinson & Waggaman, 1909; Bury & Davies, 1932), and the equilibrium
assemblages will be the same whichever reagents are used, Urwongse &
Sorrell (1980a) decided to use magnesium oxide and aqueous hydrochloric
acid for their work. This approach had the advantage that magnesium
oxide is much more soluble in hydrochloric acid solutions than in
magnesium chloride solutions, and hence measurements were easier and

294
Magnesium oxychloride cements
more reliable. Having obtained data on solubilities for the MgO-HCl-H2O
system, these were recalculated to give the compositions in the
MgO-MgCl2-H2O phase diagram.
This work was of value in constructing that part of the phase diagram
involving solutions in water. Of greater value in understanding cement
formulations were the results obtained in the earlier study (Sorrell &
Armstrong, 1976) for the MgO-rich portion of the phase diagram.

7.3.6 Consequences for practical magnesium oxychloride cements


The results of Sorrell & Armstrong (1976) show clearly that care is needed
in selecting the magnesium oxide when practical oxychloride cements are
being prepared. In particular, the powder must be of good reactivity with
small uniform crystallites and minimum agglomeration. Such an oxide
sample is capable of being rapidly dissolved by the aqueous magnesium
chloride, thereby forming a thixotropic suspension. This suspension then

MgO
Phase assemblages
90
A Mgo-Mg(OH)2-5:1:8
B MgO-5:l:8-3:l:8
C MgO-3:1:8-MgCl2.6H2O
D 3:1:8-MgCl2.6H2O-gel
E 3:l:8-5:l:8-gel
F 5:1:8-Mg(OH)2-gel
G Mg(OH)2-gel
H 5:l:8-gel
I 3:l:8-gel
J MgCl2.6H2O-gel
K gel
L gel-liquid
Mgcl2.6H2o

Wt% Mgci2

Figure 7.2 Phase relationships in part of the MgO-MgCl2-H2O system at room temperature
(Sorrell & Armtrong, 1976).

295
Oxysalt bonded cements
develops firstly into a homogeneous gel and finally into a crystalline
material consisting of a dense aggregate of the equilibrium ternary oxide
phase.
Relatively unreactive powders lead to a different sequence of events.
First, long periods of time are required for dissolution into the MgCl2
solution, during which water is lost from the liquid surface and a
compositional gradient is established. Second, under such conditions,
MgCl 2 is able to migrate to the surface where it may either precipitate or
remain as a deliquescent layer. The combined effects of the compositional
gradient and the presence of unreacted MgCl2 and MgO are almost certain
dimensional instability, leaching of corrosive salts, and poor weather
resistance.
The amount of variation in reactivity which may be tolerated is small,
since a reasonable balance has to be struck between rapid and uniform
reaction on the one hand and practical working times on the other. Sorrell
& Armstrong (1976) found that the mean crystallite diameter could be
determined adequately by X-ray diffraction, using line-broadening as an
indication of crystallite size, and also by electron microscopy. These
techniques were able to distinguish between suitable and unsuitable oxide
powders.
Both the 5:1:8 and 3:1:8 phases were shown to be unstable in water,
dissociating to give Mg(OH) 2 and MgCl2 solution (Sorrell & Armstrong
1976). This clearly has consequences when magnesium oxychloride
cements are employed as exterior stuccos on buildings. In fact, these
cements do have good weather resistance, but not because of any inherent
stability of the parent oxychloride cement. Rather, the slow conversion of
the oxychloride to the much less soluble basic magnesium carbonate as a
result of reaction with atmospheric carbon dioxide creates a material of
inherently good weather resistance.
Studies of samples of magnesium oxychloride cements used as exterior
structures have been carried out on specimens between one month and 50
years in age (Sorrell & Armstrong, 1976). From these it has been possible
to gain a general understanding of the mechanism of the reaction with
carbon dioxide and hence the way in which weather resistance develops. If
the cement has initially reacted completely to give the 5:1:8 phase, as it will
if the magnesium oxide used is sufficiently reactive, then the weather
resistance develops due to formation of the chlorocarbonate
Mg(OH) 2 .2MgCO 3 .MgCl 2 .6H 2 O. This product arises following inter-
action with atmospheric carbon dioxide, as first reported by Cole &

296
Magnesium oxychloride cements
Demediuk (1955). A surface coating of this carbonate protects the
underlying cement from attack by water. The longer-term stability,
however, depends on slow leaching of the chloride from this new
surface of the cement, resulting in its conversion to hydromagnesite,
5MgO • 4CO2. 5H2O. The precise sequence of these reactions remains to be
elucidated.

73.7 Impregnation with sulphur


One method of overcoming some of the instability and loss in strength of
oxychloride cements when exposed to water has been to modify them by
impregnation with sulphur (Beaudoin, Ramachandran & Feldman, 1977).
The resulting material appears to be a composite in which the respective
components complement each other. The magnesium oxychloride part has
relatively poor resistance to water as initially formed, whereas the sulphur
is difficult to wet and is completely insoluble in water.
Beaudoin, Ramachandran & Feldman (1977) studied this system in
detail, examining both the mechanical properties and water resistance of
sulphur-filled cements. Mechanical properties evaluated included porosity,
modulus of elasticity and microhardness. Thermal characteristics were
determined by DTG, and for certain fractured specimens microstructure
was determined by examining fracture surfaces on the scanning electron
microscope. Specimens were prepared as 5-1-cm cubes and cured at 50%
r.h. for 13 months, after which thin discs were cut from them to enable
mechanical properties to be evaluated. Two series of samples were
prepared, which were respectively (1) immersed in water for 88 days,
impregnated with sulphur, then exposed again to water, and (2) impreg-
nated first, immersed in water for 88 days, followed by further im-
pregnation, then water exposure. Impregnation was carried out in a bath
at 128 °C, at which temperature the sulphur was molten and would diffuse
thoroughly into all the pores of the native cement. Samples were removed
and, after cooling to allow solidification of the sulphur, the excess sulphur
was removed by washing in kerosene.
For the native cement exposed first to water, there was a dramatic and
rapid drop in microhardness, 30^0% in the first hour, and 55-60% at
eight hours. Compressive strength was assumed to have undergone a
similar decrease, since it is linearly related to microhardness for cemen-
titious materials (Beaudoin & Feldman, 1975). Scanning electron micro-
scopy revealed clearly the differences that occurred on soaking in water.

297
Oxysalt bonded cements
These include a change from a non-distinct and amorphous appearance
to a more porous structure with larger platy crystals. The reason for this
was not completely clear, though it was suggested that some degree of
recrystallization might have occurred during immersion in water. It may
also be that in the sample that had not been soaked in water the finely
divided portion obscured the large platy crystals.
The response of magnesium oxychloride cements to water was found to
vary according to whether or not they were impregnated with sulphur. For
microhardness, the impregnated samples gave initially higher values than
the unimpregnated samples. Microhardness decreased after 88 days'
immersion in water, but the decline was greater in the unimpregnated
samples. Unimpregnated samples had extremely low values of micro-
hardness, 10-20 MPa compared with 400-600 MPa for different formu-
lations prepared by impregnation with sulphur (Beaudoin, Ramachandran
& Feldman, 1977).
Modulus of elasticity showed similar behaviour. Impregnated samples
were found to have initially a higher modulus value than unimpregnated
ones; the decrease in modulus was less for the impregnated cements
following the 88 days' immersion in water.
The mechanism by which sulphur has these observed effects is as follows.
Immersion of native magnesium oxychloride cement in water brings about
a slow dissolution which creates pores. When those pores are filled with
sulphur, sites of possible stress concentration at points of contact between
particles are modified. Similar effects occur when sulphur is used to
impregnate hydraulic cements based on Portland cement and silica
(Beaudoin, Ramachandran & Feldman, 1977).
There were, however, important differences between the water-resistance
properties of sulphur-impregnated Portland cement and sulphur-
impregnated magnesium oxychloride cement. The former showed poor
resistance, in one case breaking into small pieces after about two hours'
exposure to water vapour. The latter, by contrast, remained completely
intact after exposure to water for 88 days, and remained in good condition
following reimpregnation with sulphur and a further 28 days of exposure
to water. Beaudoin, Ramachandran & Feldman (1977) attributed this
result in part to the difference in pore size and distribution in the two
cements, which leads to the magnesium oxychloride cement having a
smaller surface area exposed to water than Portland cement.
Overall, these studies showed that sulphur could be used to impregnate
magnesium oxychloride cements thereby yielding materials of superior

298
Magnesium oxysulphate cements

properties in terms of both mechanical characteristics and water resistance.


The preferred method of preparation was to soak the cement in water prior
to impregnation, since of the two impregnation regimes employed this gave
cements of marginally superior properties.

7.4 Magnesium oxysulphate cements


7.4.1 Setting chemistry
The magnesium oxysulphate cements are formed by reactions between
high-reactivity magnesium oxide powder and aqueous solutions of
magnesium sulphate. They have a number of applications in architecture,
including use as binders in lightweight panels and as insulating materials.
There have been problems in fully characterizing these cements, because
often it has not been appreciated that they have not been allowed to
equilibrate, and as a result their composition may be uncertain. To rectify
this situation, Urwongse & Sorrell (1980b) carried out a study of these
materials under conditions which allowed equilibrium to be attained. In
this way they were able to gain an insight into the phase relationships that
occur in these cements.
Previous studies had been carried out on mixtures rich in aqueous
magnesium sulphate and maintained at various temperatures between
30 °C and 120 °C (Demediuk & Cole, 1957). These revealed the existence
of three crystalline phases at lower temperatures, namely magnesium
hydroxide, hydrated magnesium sulphate, MgSO 4 .7H 2 O, and a com-
plex crystalline salt whose composition corresponded to an
Mg(OH) 2 : MgSO 4 : H 2 O ratio of 3:1 :8. At higher temperatures, a number
of other well-defined crystalline phases formed, including 5:1:3,1:1:5 and
1:2:3. Their relationships to each other and to the starting materials were
established. For example, at lower concentrations of magnesium sulphate
solution the 5:1:3 phase was found to be stable at temperatures greater
than 40 °C in equilibrium with Mg(OH) 2 and liquid. At high concentra-
tions the equilibrium altered, so that the 5:1:3 phase occurred in
association with the 1:1:5 phase rather than with Mg(OH) 2 . Above
100 °C, the 1:2:3 phase was stable in equilibrium with the 5:1:3 phase and
liquid at lower concentrations, and with the solid MgSO 4 . 7H 2 O and liquid
at higher concentrations (Demediuk & Cole, 1957).
Other crystalline phases have been reported for the magnesium
oxysulphate system. The 2:3:5 phase has been claimed to be formed from

299
Oxysalt bonded cements

magnesium sulphate solutions with concentrations exceeding 17-5 wt%


(Adomavichiute, Yanitskii & Vektaris, 1962). Unfortunately, no analytical
data accompanied this report and so it is not certain how reliable the claim
is.
Acid salts as well as basic salts have been claimed to occur in this system.
In particular, the complex salt MgSO 4 . H 2 SO 4 . 3H2O was found at 12-6 °C
(Montemartini & Losana, 1929) arising as the product of reaction between
MgSO 4 and aqueous H 2 SO 4 . This acid salt was found to coexist with
MgSO 4 .7H 2 O, MgSO 4 .H 2 O and liquid.

7.4.2 Phase relationships in the MgO-MgSO\-H2O system


A detailed study of the phase relationships in the magnesium oxysulphate
cement was carried out by Urwongse & Sorrell (1980b). They used X-ray
analysis to examine the phases present in the cement, and established the
composition of the invariant liquids after equilibration by measuring
specific gravity with the aid of a pycnometer. Specific gravities were related
to concentration by means of a calibration exercise in which 30 stock
solutions of sulphuric acid at concentrations between 0 and 79-5 wt % were
prepared with distilled water.
A number of species were found at equilibrium in the solid state in this
system, including MgSO 4 .7H 2 O, MgO, MgSO 4 .6H 2 O and MgSO 4 .H 2 O.
Only one complex salt, the 3:1:8 phase, was found under the conditions
studied, with both H 2 SO 4 and H 2 O occurring as discrete entities in certain
of the phases. Non-equilibrium phases were also apparent, including the
1:1:5 phase in certain samples. The non-equilibrium phase MgSO 4 . 7H2O
was also found in numerous samples, particularly those prepared from
sulphuric acid solutions with concentrations above 20wt%. The oc-
currence of these non-equilibrium phases, and also uncertainty about the
possible existence of the phase MgSO 4 . H 2 SO 4 . 3H 2 O, led to problems in
constructing the complete phase diagram for the MgO-H 2 SO 4 -H 2 O
ternary system. In particular there were difficulties in establishing phase
relationships on the H2SO4-rich side of the diagram.
The X-ray diffraction lines corresponding to Mg(OH) 2 were distinctive.
The basal line (001) was very broad, the prism line (110) was very sharp and
the other lines (hkl) were intermediate in breadth. This was interpreted as
implying that Mg(OH) 2 crystallites adopt an exaggerated sheet-like
morphology in this system (Urwongse & Sorrell, 1980b). The (001) line

300
Magnesium oxysulphate cements
became sharper with time for a series of samples of Mg(OH)2 crystals left
in contact with the liquid in sealed containers. By contrast, the prism line
(110) remained sharp and unchanged for 64 days. Thesefindingsindicate
that not only were the initially formed crystallites in the form of very thin
sheets, but that crystal growth continued slowly with time in the c
direction.
The phase diagram constructed by Urwongse & Sorrell (1980b) for the
chemical equilibria in the MgO-H2SO4-H2O system applies also to the
MgO-MgSO4-H2O system, and is thus relevant for gaining an insight into
the phase relationships which occur in the magnesium oxysulphate cements
(Figure 7.3). There is, though, the possibility of non-equilibrium phases
appearing. For example the phases 1:1:5 and MgSO4. 4H2O which have
been found experimentally in this system have been assumed to arise
because of localized temperature increases on mixing. These phases, which

O PREPARED SAMPLES

• REPORTED COMPOUNDS

80 © 1-1-3 NON-EQUILIBRIUM

MgS04"4H20 NON-EQUILIBRIUM
/ \ _
Mg(OH) ;
SOLUBILITY MEASUREMENTS

40 50 90
H 2 S0 4
WEIGHT PERCENT

Figure 7.3 Phase relationships in the MgO-H2SO4-H2O system at room temperature


(Urwongse & Sorrell, 1980b).

301
Oxysalt bonded cements
are known to be stable above room temperature, are then presumably
stranded as the mixture cools. Further work over longer periods of time
would be necessary to confirm this.
The phase diagram of Urwongse & Sorrell (1980b) has important
consequences for the formulation of magnesium oxysulphate cements. In
particular, it indicates the impossibility of preparing cements at 23 °C
which have more than 50 wt % of the 3:1:8 phase in them if the starting
materials are MgO and aqueous MgSO4. Attempting to obtain more of the
phase by the alternative tactic of reacting the oxide with sulphuric acid
solutions is not practical because of the rapid formation of MgSO4. 7H2O
from these starting materials. From these and other observations of the
phase relationships, the bonding phase in commercial oxysulphate cements
is concluded to be 5:1:3. In addition, varying amounts of MgSO4. 7H2O
are likely to occur in thefinishedproduct, though this phase is not desirable
and minimizing its formation must be part of the aim of formulating
practical cements with optimum properties.

7.43 Mechanical properties of magnesium oxysulphate cements


The use of magnesium oxysulphate cement as a binder in building materials
has been recognized since the earliest report of it by Sorel (1867). A study
of the development of strength in these cements cured under pressure was
carried out (Beaudoin & Ramachandran, 1978) in order to determine
whether strength could be improved by this means. Other cementitious
materials that usually produce weak specimens can undergo significant
and technically useful increases in strength when cured under pressure, and
the study aimed to discover whether magnesium oxysulphate also fell into
this category.
For this study a series of magnesium oxysulphate cements was prepared
having different initial compositions from those prepared by Urwongse &
Sorrell (1980b) in their study of the phase equilibria. Beaudoin &
Ramachandran (1978) formulated their cements from solutions of MgO
and MgSO4. 7H2O, the latter being dissolved in distilled water to give a
saturated solution.
Having prepared the cements, a number of physical techniques were
employed to study them. Porosity was determined by measuring the
apparent volume and the solid volume, the former by straightforward
linear measurement, the latter using a helium pycnometer. This technique
was used to avoid the problems of dissolution that arise when water is used

302
Magnesium oxysulphate cements
in a displacement technique to evaluate porosity. However, because of the
water-sensitivity of these cements, care was taken to condition them fully
at 11 % r.h. prior to taking any measurements.
In addition to porosity, modulus of elasticity and microhardness were
determined. Differential thermograms were recorded, at a heating rate of
20 ° min"1.
Compaction pressure and resulting porosity were found to be related in
a logarithmic way, such that plotting log(compaction pressure) against
porosity gave a straight line (Beaudoin & Ramachandran, 1978). This is
typical for cementitious materials, though what was not typical was the
abrupt change of slope at approximately 7-5 % porosity. Below this value
the slope of the graph was found to be much steeper than at higher
porosity. This result was held to arise from the porous nature of the
particles bound together in the cement matrix. Above the critical value of
porosity (7-5 %) compaction occurs when particles slide past each other as
the matrix deforms. Below this value, compaction is possible only when the
particles themselves are heavily deformed or fractured; this difference in
mechanism results in the observed change in relationship between
compaction pressure and porosity.
The microhardness and modulus of elasticity were found to alter on
compaction but in opposite directions. For microhardness, compacted
specimens gave higher values than uncompacted, whereas for modulus of
elasticity compacted specimens gave lower values. The reason for this was
not clear, though a tentative explanation was advanced by Beaudoin &
Ramachandran (1978). They suggested that modulus of elasticity is
determined predominantly by the bonding between particles, while
microhardness is largely dependent on the strength of the solid component.
In the case of interparticle bonding, compaction causes bond fracture and
particle slippage, and once the compacting force is removed only a fraction
of the interparticle bonds will remain unaffected. Hence the property that
depends on the strength of these interparticle forces, i.e. modulus of
elasticity, will decrease. By contrast, microhardness depends on the extent
to which the measurement of hardness reflects the value for the crystalline
component, rather than the matrix phase, and this would favour
compacted specimens. These explanations are somewhat speculative and
qualitative, and it is not clear just how helpful they are in understanding
the effect of compaction on these cements.
One important finding which did emerge from the work of Beaudoin &
Ramachandran (1978) was that, contrary to previous assumptions,

303
Oxysalt bonded cements
magnesium oxysulphate cements are not significantly weaker than mag-
nesium oxychloride cements, provided equal porosities are considered.
Previous strength measurements had been done on samples at varying
porosities, and since the oxysulphate cement more readily develops a
porous structure, such a comparison always showed it up unfavourably.
However, when specimens of equal porosity were examined, the oxy-
chloride cements proved to be only marginally stronger than the
oxysulphate cements.
Overall, the major conclusion from this study was that the magnesium
oxysulphate cement system responds differently from the Portland cement
system to being formed under compaction. The principal difference is that
the oxysulphate cements are weakened by compaction. This appears to be
because compaction, particularly at higher porosities, alters the mor-
phology of the particulate phase, changes the pore size distribution and
reduces the extent of bonding between particles. Nevertheless, from the
practical point of view, pressing oxysulphate cements in order to form
them is capable of producing specimens of adequate strength for use in
architectural applications.

7.5 Other oxysalt bonded cements


The three major oxysalt bonded cements that have already been described
in detail in this chapter are not the only ones that have been prepared,
though they are the ones that have been the most thoroughly studied. For
example, Demediuk, Cole & Hueber (1955) gave some details about the
calcium analogues of the magnesium oxychloride cements. Like the
magnesium cements, they were fabricated by reaction of powdered metal
oxide with aqueous solutions of metal chloride. The resulting calcium
oxychloride cements were similar to the magnesium oxychloride cements.
However, unlike the latter materials, they have found few or no
architectural or similar applications, and as a result there has been hardly
any interest in developing an understanding of their setting chemistry or
structure.
In a similar way there has been a passing reference to a cobalt
oxychloride cement (Prosser et ai, 1986). No explicit details of the
fabrication or chemical behaviour of this material were provided, but the
ingredients were listed among series of acids and bases for forming cements
as agents for the sustained release of trace elements to grazing animals. The
implication of this paper was that cobalt oxide would function as the base

304
References
and aqueous cobalt chloride as the acid, and that these two compounds
would form cements that were chemically and structurally similar to the
zinc and magnesium oxychloride materials.

References
Adomavichiute, O. B., Yanitskii, I. V. & Vektaris, B. I. (1962). On the
hardening of magnesium cement. Zhurnal Priklandnoi Khimii, 35, 2551-4.
Aspelund, H. (1933). Basic salts of bivalent metals: II. Acta Academiae
Aboensis, 7 (6), 1-25.
Beaudoin, J. J. & Feldman, R. F. (1975). Mechanical properties of autoclaved
calcium silicate systems. Cement and Concrete Research, 5 (2), 103-18.
Beaudoin, J. J. & Ramachandran, V. S. (1975). Strength development in
magnesium oxychloride and other cements. Cement and Concrete, 5 (6),
617-30.
Beaudoin, J. J. & Ramachandran, V. S. (1978). Strength development in
magnesium oxysulphate cement. Cement and Concrete, 8 (1), 103-12.
Beaudoin, J. J., Ramachandran, V. S. & Feldman, R. F. (1977). Impregnation
of magnesium oxychloride cement with sulphur. Ceramic Bulletin, 56, 424-7.
Bury, C. R. & Davies, E. R. H. (1932). System magnesium oxide-magnesium
chloride-water. Journal of the Chemical Society, 2008-15.
Cole, W. F. & Demediuk, T. (1955). X-ray, thermal and dehydration studies
on magnesium oxychlorides. Australian Journal of Chemistry, 8, 234-51.
Demediuk, T. & Cole, W. F. (1957). A study of magnesium oxysulphates.
Australian Journal of Chemistry, 10, 287-94.
Demediuk, T., Cole, W. F. & Hueber, H. V. (1955). Studies on magnesium and
calcium oxychlorides. Australian Journal of Chemistry, 8, 215-33.
Droit, A. (1910). Oxychlorides of zinc. Comptes rendus hebdomadaires des
seances de V Academie des Sciences, 150, 1426-8.
Eubank, W. R. (1951). Calcination studies of magnesium oxides. Journal of the
American Ceramic Society, 34, 225-9.
Feitknecht, W. (1930). Reaction of solid substances in liquids: 1. Helvetica
Chimica Acta, 13, 22-43.
Feitknecht, W. (1933). Structure of the basic salts of bivalent metals. Helvetica
Chimica Acta, 16, 427-54.
Feitknecht, W., Ostwald, H. R. & Forsberg, H. E. (1959). Uber die Struktur der
Hydroxidchloride MeOHCl. Chimia, 13 (4), 113.
Forsberg, H. E. & Nowacki, W. (1959). Crystal structure of ZnOHCl. Acta
Chemica Scandinavica, 13 (5), 1049-50.
Greenwood, N. N. & Earnshaw, A. (1984). The Chemistry of the Elements.
Oxford: Pergamon Press.
Harper, F. C. (1967). Effect of calcination temperature on the properties of
magnesium oxides for use in magnesium oxychloride cements. Journal of
Applied Chemistry, 17, 5-10.
Hayek, E. (1932). Basic salts: I. Zeitschrift fur anorganische undallgemeine
Chemie, 207, 41-5.

305
Oxysalt bonded cements

Holland, H. C. (1930). The ternary system zinc oxide-zinc chloride-water.


Journal of the Chemical Society, 643-8.
Mellor, J. W. (1925). A Comprehensive Treatise on Inorganic and Theoretical
Chemistry, vol. IV, pp. 535-46. London: Longmans Green.
Montemartini, C. & Losana, L. (1929). Equilibria between double sulfates and
aqueous solutions of sulfuric acids of various concentrations. Industria
Chimica (Rome), 4, 199-205.
Nowacki, W. & Silverman, J. N. (1961). Crystal structure of zinc
hydroxychloride II, Zn 5 (OH) 8 Cl 2 .1H 2 O. Zeitschriftfur Kristallographie,
Kristallgeometrie, Kristallphysik, Kristallchemie, 115, 21-51.
Nowacki, W. & Silverman, J. N. (1962). Appendix to the paper 'Crystal
structure of zinc hydroxychloride II, Zn 5 (OH) 8 Cl 2 .1H 2 O\ Zeitschrift fur
Kristallographie, Kristallgeometrie, Kristallphysik, Kristallchemie, 117, 238-40.
Prosser, H. J., Wilson, A. D., Groffman, D. M., Brookman, P. J., Allen, W. M.,
Gleed, P. T., Manston, R. & Sansom, B. F. (1986). The development of
acid-base cements as formulations for the controlled release of trace elements.
Biomaterials, 7, 109-12.
Robinson, W. O. & Waggaman, W. H. (1909). Basic magnesium chlorides.
Journal of Physical Chemistry, 13, 673-8.
Sorel, S. (1855). Procedure for the formation of a very solid cement by the
action of a chloride on the oxide of zinc. Comptes rendus hebdomadaires des
seances de P Academie des sciences, 41, 784-5.
Sorel, S. (1867). On a new magnesium cement. Comptes rendus hebdomadaires
des seances de TAcademie des sciences, 65, 102^.
Sorrell, C. A. (1977). Suggested chemistry of zinc oxychloride cements. Journal
of the American Ceramic Society, 60, 217-20.
Sorrell, C. A. & Armstrong, C. R. (1976). Reactions and equilibria in
magnesium oxychloride cements. Journal of the American Ceramic Society, 59,
51-4.
Urwongse, L. & Sorrell, C. A. (1980a). The system MgO-MgCl 2 -H 2 O. Journal
of the American Ceramic Society, 63, 501—4.
Urwongse, L. & Sorrell, C. A. (1980b). Phase relationships in magnesium
oxysulfate cements. Journal of the American Ceramic Society, 63, 523-6.
Walter-Levy, L. (1937). Neutral chlorocarbonate of magnesium. Comptes rendus
hebdomadaires des seances de VAcademie des Sciences, 204, 1943-6.
Walter-Levy, L. (1938). Contribution to the study of halogenocarbonates of
magnesium. Comptes rendus hebdomadaires des seances de VAcademie des
sciences, 205, 1405—7.
de Wolff, P. M. & Walter-Levy, L. (1953). Crystal structure of Mg2(OH)3(Cl,
Br), 4H 2 O. Acta Crystallographica, 6, 40-4.

306
8 Miscellaneous aqueous cements

8.1 General
This chapter is devoted to a miscellaneous group of aqueous acid-base
cements that do not fit into other categories. There are numerous cements
in this group. Although many are of little practical interest, some are of
theoretical interest, while others have considerable potential as sustained-
release devices and biomedical materials. Deserving of special mention as
biomedical materials of the future are the recently invented polyelectrolyte
cements based on poly(vinylphosphonic acids), which are related both to
the orthophosphoric acid and poly(alkenoic acid) cements.

8.2 Miscellaneous aluminosilicate glass cements


In 1968 Wilson published an account of his early search for alternatives to
orthophosphoric acid as a cement-former with aluminosilicate glasses.
Aluminosilicate glasses of the type used in dental silicate cements were used
in the study and were reacted with concentrated solutions of various
organic and inorganic acids. Wilson (1968) made certain general observa-
tions on the nature of cement formation which apply to all cements based
on aluminosilicate glasses.
(1) Silica gel is formed in the reaction but is not associated with
cement formation.
(2) Water is essential to the reaction and cements are not formed
when the acid is present in an organic solvent rather than in
aqueous solution. Water acts as an effective reaction medium and
probably hydrates reaction products.
(3) Cements are formed only with acids that are capable of forming
complexes with calcium and aluminium. Thus, neither hydro-
chloric nor acetic acid forms cements.

307
Miscellaneous aqueous cements
Table 8.1. Properties of aluminosilicate glass cements prepared with
various acids in aqueous solution {Wilson, 1968)

Compressive
Acid Powder: Setting strength Water-leachable
aqueous solution, liquid, time, (24 hour), material,"
% by mass gem"3 minutes MPa % mass
Tannic acid, 50 % 40 8-0 77 11
Tartaric acid, 50 % 40 4-5 75 15
Citric acid, 50 % 40 6-5 42 CD.
Pyruvic acid, 50 % 40 5-5 36 CD.
Malic acid, 50% 40 150 35 CD.
Fluoboric acid, 42 % 3-5 2-2 106 6-3
Glycerol phosphoric 40 3-2 38 3-3
acid, 35%
Conventional silicate 40 3-5 272 01
cement phosphoric
acid liquid
a
On 24-hour-old cements.
CD. Complete disintegration.

(4) Cements set rapidly.


(5) Cement formation with carboxylic acids is associated with
carboxylate (salt) formation.
(6) The ability of a cement to resist aqueous attack depends on the
nature of the acid anion.
The properties of these cements are given in Table 8.1.
All the cement-forming organic acids were multifunctional and capable
of forming strong chelates with aluminium and weaker chelates with
calcium (Perrin, 1964; Martell & Calvin, 1952). The cements of citric,
pyruvic and malic acids were found to disintegrate completely when placed
in water, in accordance with the soluble nature of their metal complexes.
The disintegration of the tartrate cement, though substantial, was
incomplete - the aluminium salt is soluble but that of calcium is not. The
tannic acid cement alone had reasonable hydrolytic stability, but then
tannic acid forms an insoluble salt with aluminium (Welcher, 1947). This
cement was the strongest, with a compressive strength of 77 MPa.
Fluoboric acid was also found to form a cement of some strength
(106 MPa) but of poor hydrolytic stability.

308
Phytic acid cements
From the practical point of view these results were disappointing, but it
was from this unpromising start that the successful glass polyalkenoate
cement was eventually developed.

8.3 Phytic acid cements


Phytic acid, myo-inositol hexakis(dihydrogen phosphate), is an abundant
constituent of plants (Graf, 1986). Its structure is that of a symmetrical six-
membered ring carrying six dihydrogen phosphate groups (Figure 8.1).
This structure suggests that it should form strong complexes, and this is so,
with the strength of the complex increasing with the valency of the cation
(Graf, 1986). Most of the complexes with polyvalent cations are insoluble,
which makes phytic acid a candidate for cement formation. Indeed, it has
been found to form cements with zinc oxide and aluminosilicate glasses
(Lion Corporation, 1980; Prosser et ai, 1983; Prosser & Wilson, 1986) and
these may be compared with phosphate-bonded cements.
Cements based on phytic acid set more quickly than their glass
polyalkenoate or dental silicate cement counterparts, but have similar
mechanical properties (Table 8.2). They are unique among acid-base
cements in being impervious to acid attack at pH = 2-7. Unfortunately,
they share with the dental silicate cement the disadvantage of not adhering
to dentine. They do bond to enamel but this is by micromechanical
attachment - the cement etches enamel - and not by molecular bonding.
Lack of adhesive property is a grave weakness in a modern dental or bone

Figure 8.1 The structure of phytic acid.

309
Miscellaneous aqueous cements

Table 8.2. Properties of cements based on phytic acid (Prosser et al.,


1983)

G-200 G-5

Phytic PAA Phytic Cement


acid, 40% 50% acid, 40 % liquid6

Powder: liquid, g cm 3 40 30 30 3-5


Working time (23 °C), 1-5 50 3-2 —
minutes
Setting time (37 °C), 2-7 4-3 3-8 50
minutes
Compressive strength" 201 166 168 183
(24 hour), MPa
Tensile strength" 12-6 140 5-7 13-0
(24 hour), MPa
Opacity, C 0 7 0-78 0-85 0-88 0-55
Water-leacnable material 0-88 21 0-74 1-25
7 minute cure, % mass
Water-leachable material 0-40 0-45 0-46 0-70
1 hour cure, % mass
Acid erosion pitting depth, 0 11-7 0 12-3
urn/hour

a
Specimens stored for 24 hours in water (37 °C).
b
Cement Liquid: 48-3% H 3 PO 4 ; 2-3% Al; 4-9% Zn.
G-200: 29-0% SiO 2 ; 16-6% A12O3; 34-3% CaF 2 ; 3-0% NaF; 2-0% A1F3;
9-9% A1PO4.
G-5: 43-7% SiO 2 ; 23-0% A12O3; 12-9% CaF 2 ; 10-5% NaF; 7% A1F3;
2-9% A1PO4.

cement. Nor do phytic acid cements have the translucency of the dental
silicate cement. Unless progress is made in remedying these two dis-
advantages it is doubtful whether they will find any use despite their
excellent resistance to erosion. More promising are the poly(vinyl-
phosphonic acid) cements that are described next.

8.4 Poly(vinylphosphonic acid) cements


Ellis and Wilson studied cements formed from concentrated solutions of
poly(vinylphosphonic acid) (PVPA) and oxides and silicate glasses, which
they termed metal oxide and glass polyphosphonate cements (Wilson &

310
Poly{vinylphosphonic acid) cements
Ellis, 1989; Ellis, 1989; Ellis & Wilson, 1990, 1991, 1992; Ellis, Anstice &
Wilson, 1991). They are polyelectrolyte cements related to the poly-
alkenoate cements and represent an attempt to improve on them. PVPA
(Figure 8.2) has a structure similar to that of PAA (Figure 5.2).
PVPA was prepared by the free-radical homopolymerization of vinyl-
phosphonyl dichloride using azobisisobutyronitrile as initiator in a
chlorinated solvent. The poly(vinylphosphonyl chloride) formed was then
hydrolysed to PVPA (Ellis, 1989). No values are available for the apparent
pATas of PVPA, but unpolymerized dibasic phosphonic acids have pKal and
pKa2 values similar to those of orthophosphoric acid, i.e. 2 and 8 (Van
Wazer, 1958). They are thus stronger acids than acrylic acid, which as a
pKa of 4-25, and it is to be expected that PVPA will be a stronger and more
reactive acid than poly (aery lie acid).

8.4.1 Metal oxide polyphosphonate cements


Ellis & Wilson (1991, 1992) examined cement formation between a large
number of metal oxides and PVPA solutions. They concluded that setting
behaviour was to be explained mainly in terms of basicity and reactivity,
noting that cements were formed by reactive basic or amphoteric oxides
and not by inert or acidic ones (Table 8.3). Using infrared spectroscopy
they found that, with one exception, cement formation was associated with
salt formation; the phosphonic acid band at 990 cm"1 diminished as the
phosphonate band at 1060 cm"1 developed. The anomalous result was that
the acidic boric oxide formed a cement which, however, was soluble in
water. This was the result, not of an acid-base reaction, but of complex
formation. Infrared spectroscopy showed a shift in the P = O band from
1160 cm"1 to 1130 cm"1, indicative of an interaction of the type

Setting times and hydrolytic stability of these cements are given in Table
8.3. In some cases the speed of reaction was very high, and practical
cements could not be formed from ZnO or CaO even when these oxides
were deactivated by heating. All the faster-setting cements exhibited good
hydrolytic stability. The stability of the complexes between divalent
cations and PVPA was found by a titrametric procedure to follow the
order Mg ~ Ca < Cu ~ Zn (Ellis & Wilson, 1991). This result was

311
Miscellaneous aqueous cements

Table 8.3. Properties of metal oxide phosphonic acid cements {Ellis, 1989;
Ellis & Wilson, 1991)

Oxide Setting time Hydrolytic stability

ZnO, MgO, CaO, Co(OH) 2 , < 1 minute Stable


HgO, CdO
Bi2O3, PbO 1-5 minutes stable
BA 1-5 minutes complete disintegration
CuO, Pb 3 O 4 , Y 2 O 3 , La 2 O 3 5-20 minutes stable
Cu 2 O 20-60 minutes stable
CoO, SnO, MoO 3 1-24 hours stable
Fe 3 O 4 , MnO 2 1-24 hours softened in water
In 2 O 3 , Cr 2 O 3 , CrO 3 1-24 hours complete disintegration
A12O3, SiO2, ZrO 2 , WO 3 non-setting —

expected, since Cu 2+ and Zn 2+ with filled d orbitals tend to form bonds of


a covalent character. However, Mg 2+ and Ca 2+ do form stable cements
with PVPA, though not with PAA. This is to be attributed to the higher
degree of site binding of cations to PVPA compared with PAA (Begala &
Strauss, 1972; Strauss & Leung, 1965).

I
CH2
OH
I
CH —
OH
CH 2
OH
CH —
OH
Figure 8.2 The structure of poly(vinylphosphonic acid).

I
P-O — M = P-0
I
0 0— M
o'l\>
I o i
I M
I M
M M M
Figure 8.3 Possible structures of metal-polyphosphonate complexes.

312
Poly(vinylphosphonic acid) cements

Table 8.4. Properties of MgO, CuO, Cu2O, Bi2Oz, La2O z phosphoric acid
cements (Ellis, 1989; Ellis & Wilson, 1990)

Oxide MgO CuO Cu 2 O Bi2O3 La 2 O 3

Powder:liquid, gcm" 3 10 50 50 50 50
Setting time (37 °C), minutes 5-7 7-3 — 2-0 —
Compressive strength 7 day (water), 56-6 54-5 31*5 18-2 11-9
MPa
Flexural strength 7 day (water) 4-5 120 — 5-1 —
24 hour, MPa
Water-leachable material 1 hour 0-24 00 — 005 —
cure, % mass
Shrinkage, linear % 6-8 1-7 1-3 3-0 31

Table 8.5. Adhesive bond strength of the magnesium phosphonate


cement to dentine and enamel (Ellis, 1989)

Bond strength, MPa

Cement Dentine Enamel

MgO-PVPA 3-4 4-9


Zinc polycarboxylate [1] 3-2-5-5 4-1-6-9
Glass polyalkenoate [2] 1-9-6-8 3-2-9-9

[1] Walls (1986); [2] Wilson & McLean (1988b).

The exact nature of metal-poly(phosphonic acid) interaction is un-


known (Ellis, 1989) although a number of structures can be drawn (Figure
8.3).
The properties of the most promising cements - those of MgO, CuO,
and Bi 2 O 3 -are given in Table 8.4. All had a short and sharp set.
Compressive and flexural strengths, although moderate, compare fairly
well with those of commercial zinc polycarboxylate cements. All cements
shrank when exposed to an atmosphere of 50 % relative humidity. The
most promising is the cement based on MgO which adheres to both dentine
and enamel with about the same bond strength as the glass polyalkenoate
and zinc polycarboxylate cements (Table 8.5).

313
Miscellaneous aqueous cements

8.4.2 Glass polyphosphonate cements


Ellis and Wilson also examined cement formation from aluminosilicate
glasses and concentrated solutions of PVPA (Wilson & Ellis, 1989; Ellis,
1989; Ellis & Wilson, 1990). These cements, like the glass polyalkenoate
cements, are a type of glass-ionomer cement.
Disadvantages of the glass polyalkenoate cements are the susceptibility
of the young cement to aqueous attack and problems in achieving sufficient
translucency to match that of tooth enamel. PVPA has a higher refractive
index than PAA, and should lessen the refractive index mismatch between
glass particle and cement matrix which is the cause of light-scattering.
PVPA is also a much stronger acid than PAA and the pendant groups are
bifunctional. For this reason it should form stronger associations with
cations and produce cements that set more rapidly.
The reactivity of the PVPA solutions with SiO 2 -Al 2 O 3 -CaO and
SiO 2 -Al 2 O 3 -CaF 2 glasses did, as expected, prove greater than that of PAA
solutions. Its cements had shorter setting times, as is shown by Figure 8.4
which depicts the setting of cements formed from glasses having the generic
composition (mole ratios) zSiO 2 , 1-00 A12O3, 1-30 CaF 2 (Ellis, Anstice &
Wilson, 1991). Note that the glass with Si/Al mole ratio of 2-55 forms
a cement with PVPA but not with PAA. Decreasing the Si/Al mole ratio
accelerates cement formation, and the setting time of cements is corre-

1000 i -

Clear
glasses

0.5 1.0 2.0 2.5


Si / Al mole ratio
Figure 8.4 The effect of Si/Al ratio in the glass on the setting time of glass polyphosphonate
and glass polyalkenoate cements (Ellis, 1989).

314
Copper oxide and cobalt hydroxide cements
spondingly reduced. This result, which is similar to that found for glass
polyalkenoate cements, is to be expected as replacement of Si by Al must
weaken the glass network for reasons discussed in Section 5.9.2. When the
Si/Al mole ratio reaches 1-1, glass PVPA cements set impossibly fast.
Below a mole ratio of 0-57 the effect is reversed as phase separation of
fluorite (CaF2) and corundum (A12O3) occurs (glasses with an Si/Al ratio
at or above 0-80 are clear). Phase separation reduces the reactivity of the
main glass phase as the Si/Al ratio is increased andfluorideis withdrawn.
Heat treatment of the glasses at 600 °C for 6 hours reduced their
reactivity, by promoting phase separation, and prolonged the setting time
of cements. Thus, the setting time of the cement formed from one glass
(Si:Al = 1-7) was increased from 2-0 minutes to 5-3 minutes and that of
another glass (Si: Al = 2-0) from 3*6 minutes to 25 minutes.
These glass polyphosphonate cements are still in an early stage of
development. A recent paper by Ellis, Anstice & Wilson (1991) reports a
maximum compressive strength of 90 MPa and a maximum flexural
strength of 10 MPa. Although more recent data indicate that a compressive
strength of 150 MPa is possible these values are still much lower than those
recorded for the best dental silicate and glass polyalkenoate cements -
systems which are, of course, fully developed. The translucency of the glass
polyphosphonate cement is good, with an opacity value sufficiently low
(C07 = 0-55) to match that of tooth enamel.
Its resistance to early contamination by water is very good and much
superior to that of the glass polyalkenoate cement. The solubility of a
seven-minute-old cement is 0*5 % which compares favourably with values
of 1-0-2-1 % reported for glass polyalkenoate cement (Wilson & McLean,
1988a). The solubility of one-hour-old cements is infinitesimal (< 0-05%)
and very much lower than that of the glass polyalkenoate cement
(0-17-0-33%).

8.5 Miscellaneous copper oxide and cobalt hydroxide cements


Copper(II) oxide and cobalt(II) hydroxide form cements with solutions of
many multifunctional organic acids: propanetricarboxylic acid, tartaric
acid, malic acid, pyruvic acid, mellitic acid, gallic acid, tannic acid and
phytic acid (Allen et al., 1984; Prosser et al., 1986). These have been used
mainly in cement devices for the sustained release of copper and cobalt
(Manston et al., 1985; Manston & Gleed, 1985). Little is known about

315
Miscellaneous aqueous cements

their structure and mechanical properties. Most devices that have been
used in animal husbandry have been based on acid phosphates and have
been dealt with in Section 6.3.

References
Allen, W. M., Sansom, B. F., Wilson, A. D., Prosser, H. J. & Groffman, D. M.
(1984). Release cements. British Patent GB 2,123,693 B.
Begala, A. J. & Strauss, U. P. (1972). Dilatometric studies of counterion binding
by polycarboxylates. Journal of Physical Chemistry, 76, 254-60.
Ellis, J. (1989). Materials based on polyelectrolytes. PhD. Thesis (Council for
National Academic Awards): Thames Polytechnic and Laboratory of the
Government Chemist, London.
Ellis, J., Anstice, M. & Wilson, A. D. (1991). The glass polyphosphonate
cement: a novel glass-ionomer cement based on poly(vinylphosphonic acid).
Clinical Materials, 7, 341-6.
Ellis, J. & Wilson, A. D. (1990). Polyphosphonate cements: a new class of
dental materials. Journal of Materials Science Letters, 9, 1058-60.
Ellis, J. & Wilson, A. D. (1991). A study of cements formed between metal
oxides and polyvinylphosphonic acid. Polymer International, 24, 221-8.
Ellis, J. & Wilson, A. D. (1992). The formation and properties of metal oxide
poly(vinylphosphonic acid) cements. Dental Materials, 8, 79-84.
Graf, E. (1986). Chemistry and applications of phytic acid: an overview. In
Graf, E. (ed.) Phytic Acid, Chapter 1. Minneapolis: Pilatus Press.
Lion Corporation. (1980). Dental cements. Nihon Kokai Tokkyo Koho
80,139,311. Chemical Abstracts, 94, 903,488b.
Manston, R. & Gleed, P. T. (1985). Reaction cements as materials for the
sustained release of trace elements into the digestive tract of cattle and sheep.
II. Release of cobalt and selenium. Journal of Veterinary Pharmacology
Therapeutics, 8, 374-81.
Manston, R., Sansom, B. F., Allen, W. M., Prosser, H. J., Groffman, D. M.,
Brant, P. J. & Wilson, A. D. (1985). Reaction cements as materials for the
sustained release of trace elements into the digestive tract of cattle and sheep.
I. Copper release. Journal of Veterinary Pharmacology Therapeutics, 8,
368-73.
Martell, A. E. & Calvin, M. (1952). Chemistry of the Chelate Compounds, p. 415.
New York: Prentice-Hall.
Perrin, D. D. (1964). Organic Complexing Reagents, p. 269. New York:
Interscience Publishers.
Prosser, H. J., Brant, P. J., Scott, R. P. & Wilson, A. D. (1983). The cement-
forming properties of phytic acid. Journal of Dental Research, 62, 598-600.
Prosser, H. J. & Wilson, A. D. (1986). The cement-forming properties of phytic
acid. In Graf, E. (ed.) Phytic Acid, Chapter 17. Minneapolis, Minnesota:
Pilatus Press.
Prosser, H. J., Wilson, A. D., Groffman, D. M., Brookman, P. J.? Allen, W. M.,
Gleed, P. T., Manston, R. & Sansom, B. F. (1986). The development of

316
References

acid-base reaction cements as formulations for the controlled release of trace


elements. Biomaterials, 1, 109-12.
Strauss, U. P. & Leung, Y. P. (1965). Volume changes as a criterion for site
binding of counterions by polyelectrolytes. Journal of the American Chemical
Society, 87, 1476-80.
Van Wazer, J. R. (1958). Phosphorus and its Compounds, pp. 486-91. New
York: Interscience Publishers Inc.
Walls, A. W. G. (1986). Glass polyalkenoate (glass-ionomer) cements: a review.
Journal of Dentistry, 14, 231-46.
Welcher, F. J. (1947). Organic Analytical Reagents, vol. 2, pp. 142-3. New
York: Van Nostrand.
Wilson, A. D. (1968). Dental Silicate Cements: VII. Alternative liquid cement
formers. Journal of Dental Research, 47, 1133-6.
Wilson, A. D. & Ellis, J. (1989). Poly-vinylphosphonic acid glass ionomer
cement. British Patent Application 2,219,289A.
Wilson, A. D. & McLean, J. W. (1988a). Glass-ionomer Cement, Chapter 4.
Chicago, London and Berlin: Quintessence Publishing Co.
Wilson, A. D. & McLean, J. W. (1988b). Glass-ionomer Cement, Chapter 6.
Chicago, London and Berlin: Quintessence Publishing Co.

317
9 Non-aqueous cements

9.1 General
The non-aqueous or chelate cements are an exceptionally diverse group of
materials (Wilson, 1975a,b, 1978; Smith, 1982b). The term chelate cement
is not strictly speaking correct, as a minority of them do not form chelates,
and some aqueous AB cements do. However, the term is a convenient one.
They are of interest in that the reaction media for the acid-base reaction
are non-aqueous, although sometimes water may play a role in cement
formation. In these cements water is replaced by an organic acid that is
liquid at room temperature and generally has chelating ability.
The low permittivity of these liquids compared with water inhibits
dissociation of the acids so that cement formation demands much more
reactive basic oxides. Oxides and hydroxides that are capable of cement
formation are ZnO, CuO, MgO, CaO, Ca(OH)2, BaO, CdO, HgO, PbO and
Bi2O3 (Brauer, White & Moshonas, 1958; Nielsen, 1963). In practice these
are confined to two: calcium hydroxide and special reactive forms of zinc
oxide.
Examples of liquid organic acids suitable for cement formation are:
(1) Alkoxyphenols, for example the 2-methoxyphenols, which include
eugenol, guaiacol and vanillates. Also, 2,5-dimethoxyphenol.
(2) /?-diketones, e.g. acetylacetone.
(3) /?-keto esters, e.g. jS-keto-ethylate.
(4) Keto acids, e.g. acetyl acetic acid.
(5) Other difunctional aliphatic carboxylic acids, e.g. lactic acid,
pyruvic acid, ethoxyacetic acid.
(6) Aldehydic aromatic acids, e.g. salicylaldehyde.
(7) Alkoxy aromatic acids, e.g. 2-ethoxybenzoic acid.

318
General

These examples are drawn from the work of Brauer, White & Moshonas
(1958), Nielsen (1963), Brauer, Argentar & Durany (1964), Stansbury,
Argentar & Brauer (1981), and Brauer, Argentar & Stansbury (1982).
All of these organic liquids are capable of forming chelates. They all
contain an acidic group (COOH, phenolic OH or enol OH) and a second
functional group, such as an ester or an ether, containing an oxygen
capable of donating an electron pair. The structural requirement is that a
five- or six-membered chelate ring is formed. Typical examples are shown
in Figure 9.1.
That chelate formation is involved in cement formation is demonstrated
by the behaviour of the methoxyphenols. The only methoxyphenols which
are capable of cement formation, the 2-substituted, are those which are
able to form chelates (Brauer, Argentar & Durany, 1964). Further,
Douglas (1978a,b) has shown that magnesium, calcium and zinc form
chelates with potassium guaiacol-4-sulphonate, that the zinc chelate is the
strongest, and that at pH ~ 6 hydroxy complex formation does not
seriously compete with chelate formation. In addition, chelate formation
has been reported between /?-diketones and a number of metals including
zinc and copper (Graddon, 1968). Some of these complexes are multi-
nuclear. In zinc complexes the preferred coordination number is five
although four- and six-coordination are also observed.
Related to these cements are the long chain aliphatic acids and aryl-
substituted butyric acid (Skinner, Molnar & Suarez, 1964). These materials
are on the market as non-eugenol cementing agents but they are unduly

(a) (b)
Figure 9.1 Chelating agents capable of forming cements, (a) 2-methoxyphenol type.
Guaiacol: R, = R2 = H. Eugenol: Rx =-CH 2 -CH=CH 2 , R2 = H. (b) ^-diketone type.

319
Non-aqueous cements
weak and can only be used for the temporary cementation of crowns
(Powers, Farah & Craig, 1976).
Although there are many potential chelate cements only three groups
are of any significance: the zinc oxide eugenol (ZOE) cement, the
2-ethoxybenzoic acid (EBA) cements and the calcium hydroxide alkyl-
salicylate cements. Only cements based on these three groups have been
commercially exploited. Their mechanical properties are poor, but all are
noted for having therapeutic effects and this is perhaps their most
important asset in biological applications. In effect, they are also devices
for the sustained release of biologically active agents.
They are extensively used for temporary cementation of crowns and as
temporaryfillingmaterials because of their therapeutic action. It is claimed
that when strengthened some can be used for permanent cementation or as
an intermediate restoration, that is a restoration that has a life somewhat
greater than a temporary filling material.
However, it must be pointed out that these claims have been made on the
basis of strength measurements made at room temperature rather than
37 °C (body temperature), and so may be misleading. This is especially true
of these materials. When measurements have been made at 37 °C (0ilo &
Espevik, 1978), these cements, in contrast to the water-based ones, show a
marked decline in all mechanical properties. Flow under load increases
sharply and strength decreases by almost an order of magnitude. This
point must be borne in mind when reading the following section of this
chapter where the reported strength measurements have been made at
room temperature.

9.2 Zinc oxide eugenol {ZOE) cements


9.2.1 Introduction and history
The zinc oxide eugenol (ZOE) cement is formed by mixing a reactive form
of zinc oxide with eugenol. It is widely used in dentistry for the temporary
cementation of crowns, for the temporary filling of teeth, as a root canal
sealer, as a sedative cavity liner, as a base in deep cavities and in soft tissue
packs for use in oral surgery (Brauer, 1965; Wilson, 1975b; Smith, 1982a).
Suitably modified with diluents and fillers it can also be used as an
impression paste (Phillips, 1982a,b).
The ZOE cement has a long history. Eugenol is the essential constituent
of oil of cloves, which has been used medically since the fourth century

320
Zinc oxide eugenol (ZOE) cements
(Molnar, 1942). Its use specifically to relieve toothache was recorded by
Vigo in the sixteenth century and reactions with metal oxides were reported
by Bonastre (1827a,b). The earliest zinc oxide chelate cements used
creosote (King, 1872) and later this was mixed with oil of cloves (Chisholm,
1873). Then oil of cloves was used by itself (Flagg, 1875) and finally its
essential constituent, eugenol (Wessler, 1894).
This cement has retained its popularity to this day, despite its poor
mechanical properties, because it is easy to use - it makes no demands on
technique - is tolerant towards living tissues and is a palliative or anodyne.
It is unlikely to be replaced, because, fortuitously, it is formed from two
medicaments and, again fortuitously, is a sustained release agent for
eugenol.
Cement formation is the result of an acid-base reaction between zinc
oxide and eugenol, leading to the formation of a zinc eugenolate chelate.
Water plays a vital role in the reaction.
Important reviews on these materials have appeared from time to time:
Brauer (1965), Wilson (1975a,b, 1978), Smith (1982a).

9.2.2 Eugenol
Eugenol, 4 allyl-2-methoxy phenol, is capable of forming cements with
ZnO, CuO, MgO, CaO, CdO, PbO and HgO (Brauer, White & Moshonas,
1958; Nielsen, 1963). Other 2-methoxy phenols are also capable of forming
cements with metal oxides, provided the allyl group is not in a 3- or
6-position where it sterically hinders the reaction (Brauer, Argentar &
Durany, 1964). These include guaiacol, 2-methoxyphenol, and the allyl
and propylene 2-methoxy phenols.
Eugenol is a very weak acid with a pK of 10-4 (Brauer, Argentar &
Durany, 1964) and occurs as a hydrogen-bonded dimer (Gerner et al.,
1966; Wilson & Mesley, 1972). The dimer contains both intra- and
intermolecular hydrogen bonds (structure I in Figure 9.2a). The presence
of an intramolecular hydrogen bond indicates that eugenol is in the cis
form.

9.2.3 Zinc oxide


Active zinc oxide is capable of forming chelate cements with a number of
liquid organic chelates. These include the /?-diketones, ketoacids and
ketoesters as well as the 2-methoxy phenols (Nielsen, 1963).

321
Non-aqueous cements
The zinc oxide used in ZOE cements differs entirely from that used in
zinc phosphate cements. Whereas the latter has to be ignited to a very high
temperature to deactivate it, the opposite is true of the zinc oxides used in
the ZOE cement, which are of an activated variety. They are normally
prepared by the thermal decomposition of zinc salts at 350 °C to 450 °C;
such oxides are yellow. Zinc oxides prepared by oxidizing zinc in oxygen
may also be used; these are white.
Further discussion of zinc oxide is deferred until the setting reaction is
considered (Section 9.2.5).

9.2.4 Cement formation


The earliest ZOE cements were prepared simply by mixing zinc oxide with
eugenol. These cements set under the moist conditions of the mouth and
then only slowly. Unlike other dental cements, the cement-forming
reaction of the ZOE cement requires acceleration rather than retardation
(Smith, 1958; Wilson & Batchelor, 1970; Crisp, Ambersley & Wilson,
1980). Although it is possible to make cements from zinc oxide and plain
eugenol by using a very reactive zinc oxide (Smith, 1960), the setting
behaviour of these cements is sensitive to variations in humidity and the
source of materials (Smith, 1982a). Commercial cements always contain
accelerators.
The most common accelerators are acetic acid (0-1 % to 2%) dissolved
in the eugenol, and zinc acetate or other zinc alkanoate (0-1 % to 8%)
blended with the zinc oxide powder (Wilson & Batchelor, 1970; Wilson
1975b). Zinc propionate or succinate is also effective (Phillips, 1982b).
Molnar & Skinner (1942) found that many salts acted as accelerators. The
most effective acetates were those of silver, sodium and zinc. Chlorides and
nitrates were even more effective. Rosin (abietic acid) also produced an
accelerating effect. In practice, nearly all commercial materials used either
acetic acid in the liquid, a zinc alkanoate, or even zinc eugenolate (20 %)
itself. Zinc chloride has been found dissolved in massive amounts in
eugenol (10 %); the resulting cement is a cross between a zinc oxychloride
and eugenolate cement. The question of 2-ethoxybenzoic acid (EBA) as an
accelerator is deferred until the EBA-eugenol cements are discussed
(Section 9.4). The setting reaction of all cements is accelerated by increase
in temperature.

322
Zinc oxide eugenol (ZOE) cements

9.2.5 Setting
Zinc oxide eugenol cements set and harden as the result of ionic reactions,
and physical changes are related to these underlying chemical ones. The
setting reaction has been studied systematically first by Copeland et al.
(1955) and by a number of workers since. It is still not fully understood.
The reaction is essentially an acid-base one with eugenol providing
hydrogen ions. Copeland et al. (1955) established that the product of
reaction had the empirical molecular formula Zn(C10H11O2)2 and that its
XRD pattern corresponded closely to that of zinc eugenolate salt. Wilson
& Mesley (1972) confirmed this finding.
The overall reaction can be represented by the following equation, where
HE represents eugenol and E eugenolate.
ZnO + 2HE = ZnE2 + H2O [9.1]
In fact, the reaction is an ionic one involving zinc and eugenolate ions.
A number of infrared spectroscopic studies have been made which have
thrown light on the cement-forming reaction (Copeland et al., 1955;
Gerner et al., 1966; Wilson & Mesley, 1972). Wilson & Mesley (1972) used
ATR spectroscopy to follow the course of the reaction and showed that
major spectral changes were almost entirely associated with loss of the

CH2 = CH CH2 v CH2 = CH CH,

,0 - CH3

CH2CH = CH,
\

II I CH2 CH = CH2

Matrix Region Water eluted

Figure 9.2a Hydrolysis of the zinc eugenolate bis chelate to the hydrogen bonded eugenol
dimer and zinc hydroxide. After Wilson & Mesley (1974).

323
Non-aqueous cements
O-H group. In the cement-forming reaction, phenolic hydrogens (O-H
stretching bands at 3520 and 3460 cm"1) are replaced by zinc ions, and a
weak chelate is formed. In effect, the hydrogen bonds of the eugenol dimer
are replaced by stronger Zn2+ bridges. Wilson & Mesley (1972) found little
or no free eugenol after 30 minutes, when the reaction was deemed to be
complete.
The bisligand chelate structure which is formed differs little from that of
the parent eugenol dimer (Structure II in Figure 92a). The molecule is an
electrically neutral chelate where two eugenolate molecules are attached to
a central zinc atom in square planar or tetrahedral configuration
(Figure 92b).
The CH3O-Zn coordinate bond in the zinc eugenolate chelate is very
weak (Gerner et ai, 1966) and the chelate has poor stability thus, the
cement-forming reaction [9.1] can be reversed. This occurs when the
cement is placed in water, when the matrix is easily hydrolysed to eugenol
and zinc hydroxide (Figure 9.2a) (Wilson, 1978; Wilson & Batchelor,

H2O

CH2=CH.CH

CH2.CH=CHo

CH2=CH.CH

CH2.CH=CH2

CH2=CH.CH

CH CH2.CH=CH2
H2O

Figure 9.2b The bisligand chelate structure of zinc eugenolate, showing bridging water
molecules. After Wilson & Mesley (1974).

324
Zinc oxide eugenol (ZOE) cements
1970). Eugenol can also be extracted from the cement matrix by methanol
(Molnar, 1967); this is further evidence of the weakness of the chelate,
which is decomposed during the extraction.
Water was not found in the set cement by Wilson & Mesley (1972), who
reasoned that it had entered the matrix. They speculated that zinc was in
octahedral coordination with two eugenolate molecules in planar positions
and two shared water molecules occupying diametrically opposed sites.
These water molecules acted as bridges in the individual chelates.
This hypothesis received support from the electrical studies of Braden &
Clarke (1974) and Crisp, Ambersley & Wilson (1980), who attributed
maxima in curves of permittivity and conductivity against time to the
liberation of water and its subsequent reabsorption into the matrix (Figure
93a,b). Crisp, Ambersley & Wilson (1980) also considered that these
maxima were due to generation of both water and ionic zinc species.
Subsequently, as the reaction proceeds the zinc ions arefixedas insoluble
zinc eugenolate.

160

I 120

80

40

20 40 120 140 160 180 200 220 240 260 280~


TIME(minutM)

Figure 9.3a Permittivity/time curves (s) of cements, showing maxima. Prepared from ZnO
ignited at 600 °C with: A - - - A dry eugenol; # . . . • eugenol+1% water; O - . - O
eugenol + 1 % chloracetic acid; • • • • • eugenol + 1 % acetic acid; • . . . • eugenol + 1 %
acetic a c i d + 1 % water. Cement powder/liquid ratio = 2-5 g cm" 3 (Crisp, Ambersley &
Wilson, 1980).

325
Non-aqueous cements
Eugenol is a very weak acid (pK= 10-4) and will not react with zinc
oxide in the absence of promoters. These reaction promoters include
water, acetic acid and zinc acetate.

The role of water in setting


The importance of water as an initiator and catalyst for the reaction
between zinc oxide and plain eugenol has been demonstrated by a number
of studies (Smith, 1958; Crisp, Ambersley & Wilson, 1980; Batchelor &
Wilson, 1969; Prosser & Wilson, 1982). In particular, the reaction is
accelerated by the humidity of the atmosphere during mixing (Batchelor &
Wilson, 1969; Crisp, Ambersley & Wilson, 1980).
As we shall see later, the catalytic effect of water is connected with the

120 140 1M I N
Tim* (minute)

Figure 93b Conductivity/time curves (a) of cements, showing maxima. Prepared from ZnO
ignited at 600 °C with: A . . . A dry eugenol; * . . . • eugenol+1% water; O - . - O
eugenol + 1 % chloracetic acid; • • • • • eugenol + 1 % acetic acid; • . . . • eugenol + 1 %
acetic a c i d + 1 % water. Cement powder/liquid ratio = 2-5 g cm"3 (Crisp, Ambersley &
Wilson, 1980).

326
Zinc oxide eugenol (ZOE) cements

presence of loosely absorbed water on the surface of zinc oxide particles


and not with water contained in eugenol (Prosser & Wilson, 1982). The
addition of water to eugenol has relatively little effect on the setting rate
(Crisp, Ambersley & Wilson, 1980). On the other hand zinc oxide powders
that do not have water absorbed on the surface react very slowly with plain
eugenol (Prosser & Wilson, 1982). Of course, a reaction will take place even
if there is only a trace of water present, for water is generated in the
reaction.
One might suppose that, since water is essential, the first step in the
reaction would be the formation of zinc hydroxide. This is not so, for
Prosser & Wilson (1982) found that zinc hydroxide did not react with
eugenol. They therefore agreed with the earlier suggestion of Crisp,
Ambersley & Wilson (1980) that the reaction was activated by the
formation of ZnOH+ on the surface of zinc oxide powders by water.
Douglas (1978a,b) has also suggested that the reaction takes place in the
water layer or at the water-eugenol interface; this suggestion is based on
measurement of zinc chelate formation constraints and the reflectance
spectrum of the admittedly water-soluble guaiacol-4-sulphonate.
The reaction may be represented as follows. First, the generation of
ions:
ZnO + H2O ^± ZnOH+ + OH" [9.2]
+ 2+
ZnOH — Zn + OH" [9.3]
HE^±H + + E" [9.4]
bination of ions:
Zn2+ + 2 E — Z n E 2 [9.5]
+
H + OH "^±H 2 O [9.6]
ZnE2 + H2O ^ ZnE 2 . H2O [9.7]

The role of additives in setting


This scheme applies only to the reaction in simple zinc oxide and eugenol
systems. The presence of an accelerator such as zinc acetate profoundly
modifies it. The addition of Zn2+ ions or acetic acid (HAc) to the system
eliminates the need for water to initiate the reaction. The reactions can then
be represented by the following series of equations.

ZnO + 2H+^= Zn2+ + H2O [9.8]


HE^± H + + E" [9.9]
Zn2+ + 2E"^± ZnE, [9.10]

327
Non-aqueous cements

Table 9.1. Effect of zinc oxide type on setting time ofZOE cements
{Prosser & Wilson, 1982)

Eugenol

Cement Type Dry Activated"


Indirect process ZnO 15 days 5-5 min
Thermally decomposed ZnO (500 °C) 3 hours 90 min

a
1 % by mass of acetic acid added to eugenol. A powder:liquid 20 g cm 3 was
used for cement preparation.

In this case acetic acid and acetate are reaction promoters. The reaction
is greatly accelerated and the setting time of cements is shortened (Table
9.1).

The role of zinc oxide in setting


The physical and chemical characteristics of zinc oxide powders are known
to affect cement formation (Smith, 1958; Norman et ai, 1964; Crisp,
Ambersley & Wilson, 1980; Prosser & Wilson, 1982). The rate of reaction
depends on the source, preparation, particle size and surface moisture of
the powder. Crystallinity and lattice strain have also been suggested as
factors that may change the reactivity of zinc oxide powders towards
eugenol (Smith, 1958).
Zinc oxide is made either by the oxidation of the metal in oxygen (the
indirect, IP, or French process), by the direct decomposition of zinc ores in
air (the direct or American process) or by the thermal decomposition of
zinc salts (TD zinc oxide). IP zinc oxides differ from TD zinc oxides in that
their surfaces do not contain absorbed water. Also, whereas TD zinc oxide
reacts with plain eugenol, IP zinc oxide hardly reacts unless activated by an
acetic acid or zinc acetate accelerator (Table 9.2).
Particle size is the rate-controlling factor in the case of cements formed
using IP zinc oxide (Smith, 1958; Norman et ai, 1964). Setting time appears
to be proportional to the median particle size (Prosser & Wilson, 1982). By
contrast, the setting times of cements prepared from TD zinc oxide do not
appear to relate to particle size.
The heat treatment of zinc oxide powders reduces their reactivity
towards eugenol, because of an increase in particle size or a decrease in
absorbed water. In the case of zinc oxide powders prepared by the thermal

328
Zinc oxide eugenol (ZOE) cements

Table 9.2. Effect ofZnO ignition temperature on cement setting


time (Prosser & Wilson, 1982)

Ignition Setting Absorbed Particle


temperature time water size
400 °c 18 min 1-26% 0-33 jam
500 °c 3 hours — 0-65 urn
600 °c 18 hours 0-33% —
800 °c 120 hours 0-11% 0-84 jam

Zinc oxide prepared by ignition of zinc carbonate. A powder: liquid


2-0 g cm"3 was used for cement preparation.

decomposition of carbonate or oxalate, increasing the temperature of heat


treatment causes them to react more slowly with eugenol (Smith, 1958;
Crisp, Ambersley & Wilson, 1980; Prosser & Wilson, 1982). This is
paralleled by a decrease in amount of water physically held on the surface
of the powder and an increase in the particle size (Prosser & Wilson, 1982).
These effects are shown in Table 9.2.
The effect of heating zinc oxide powders in various atmospheres has
been studied by several workers (Blackman, 1962, 1963; Lee and
Parravano, 1959; Marshall, Enrigh & Weyl, 1952; Dollimore & Spooner,
1971; Prosser & Wilson, 1982). Heating causes sintering, a process of
coalescence and densification, which finally leads to the formation of a
non-porous body. For example, in air, freshly prepared zinc oxide spheres
sinter at 700 °C (Lee & Parravano, 1959). This sintering is shown clearly in
electronmicrographs of powders before and after heating at 800 °C
(Prosser & Wilson, 1982; Figure 9Aa,b,c,d). Heating results in the loss of
oxygen which creates anion vacancies (F). The structure of such zinc
oxides may be represented as

Thus, there is excess of zinc over that required by stoichiometry. Sintering


was attributed by Lee & Parravano (1959) to the diffusion of Zn2+ ions
consequent on the difference in concentration of excess Zn2+ ions between
the surface and the bulk.
The presence of water on the oxide surface can enhance the sintering of
zinc oxide particles (Dollimore & Spooner, 1971). The amount of water
reversibly absorbed on zinc oxide surfaces is affected by heat treatment

329
Figure 9.4 The effect of sintering temperature on the morphology of zinc oxide particles. Zinc oxide from zinc oxalate: (a)
400 °C, (b) 800 °C. Carmox zinc oxide (c) 400 °C, (d) 800 °C (Prosser & Wilson, 1982).
Zinc oxide eugenol (ZOE) cements
because of the reduction in specific surface area. Nagoe & Morimoto
(1969) found that adsorption of water on zinc oxide particles reached a
maximum after heating the oxide to 450 °C in vacuo and concluded that
both chemisorption and physicosorption occurred. At this temperature,
surface hydroxyls are lost from the surface:
-Zn-OH + HO-Zn- > -Zn-O-Zn- + H2O [9.11]
However, these are rapidly regenerated on exposure to water vapour. This
ability is lost on ignition to higher temperatures and must relate to the
conversion of surface hydroxyls to oxygen bridges that are resistant to
rehydroxylation.

9.2.6 Structure
The set cement consists of zinc oxide particles bonded together by a loose
matrix of zinc eugenolate (Wilson, Clinton & Miller, 1973). Electron-
micrographs show that the zinc oxide particles are covered by zinc
eugenolate (Figure 9.5a).
Different workers have reported different results on the nature of the
zinc eugenolate matrix. Earlier work and some recent work has indicated

Figure 9.5a Electronmicrograph of a ZOE cement matrix, showing zinc oxide particles
covered by zinc eugenolate (Wilson, Clinton & Miller, 1973).

331
Non-aqueous cements
that crystallites are present (Copeland et al., 1955; Wilson, Clinton &
Miller, 1973; Bayne et aL, 1986). By contrast, Steinke et al (1988) found
that the matrix was entirely amorphous.
El-Tahawi & Craig (1971), using thermal analysis, came to the following
conclusions. The setting of ZOE mixes, containing 0 to 1 % of zinc acetate,
does not result in the formation of more than trace amounts of zinc
eugenolate crystallites, so that setting has nothing to do with the formation
of a crystalline phase. Zinc eugenolate crystallites are formed in appreciable
amounts only when the cement formulation contains large amounts of zinc
acetate. It is unlikely, therefore, that the classical explanation of Copeland
et al (1955), that coherence is due to the interlocking of crystals, is correct.
Recently, Bayne & Greener (1985) have cast doubt on El-Tahawi & Craig's
interpretation, so the exact nature of the matrix remains to be resolved.
On exposure to water the matrix decomposes, with release of eugenol
(Figure 9.2a). Wilson, Clinton & Miller (1973) found that the zinc
eugenolate matrix was degraded to a weak zinc hydroxide matrix and the
zinc oxide particles were washed clean of zinc eugenolate (Figure 9.56).

Figure 9.56 Electronmicrograph of a ZOE cement matrix after aqueous attack. The zinc
oxide particles are washed clean of zinc eugenolate and the matrix is degraded to zinc
hydroxide (Wilson, Clinton & Miller, 1973).

332
Zinc oxide eugenol (ZOE) cements

9.2.7 Physical properties


The ZOE cement is easy to mix and a greater amount of powder can be
incorporated into this cement (5:1 by mass) than any other, where even 4:1
by mass is unusual. Because the ZOE cement is sensitive to moisture it can
be formulated to have a long working time under normal room conditions
(23 °C, relative humidity 50 %) and a rapid set once placed in the warm and
moist conditions of the mouth. This is a considerable clinical advantage,
making it convenient to use. The cement can be used in a war pack for use
on the battlefield. Nevertheless, sensitiveness to humidity can give rise to
problems in use under tropical conditions.
The working time of commercial materials varies from 4 to 14 minutes
(Plant, Jones & Wilson, 1972; Wilson, 1975a). The International Standard
(ISO, 1988) requires that these cements when used for temporary
cementation or as a cavity liner should set in 4 to 10 minutes and when used
as a base or for temporary restoration should set in 3 to 10 minutes. The
setting of commercial ZOE bases can be rapid and so sufficient strength
can be developed for an amalgam to be placed over them after 10 minutes
(Plant & Wilson, 1970). Linear shrinkage during setting is high, 0-86 % dry
and 0-32 % wet (Civjan & Brauer, 1964).
Compressive strength is much lower than that of the water-based dental
cements and ranges from 13 MPa to 38 MPa (24 hours) for unreinforced
materials (Brauer, 1972; Wilson, 1975a). ZOE cements are suitable for use
as liners and temporary cements. Gilson & Myers (1970) pointed out that
ZOE cements for temporary cementation should be strong enough to
ensure retention of devices yet weak enough to allow for ease of removal.
They concluded from clinical studies that materials with compressive
strengths ranging from 2 to 24 MPa were required. The International
Standard (ISO, 1988) requires that for temporary cementation compressive
strength should not exceed 35 MPa and for use as liners must be greater
than 5 MPa. Tensile strength is much lower, 1-2 to 2-8 MPa (Civjan &
Brauer, 1964; Hannah & Smith, 1971).
These cements have marked creep characteristics and flow under
pressure even when fully set. In this they contrast markedly with the rigid
phosphate cements (Wilson & Lewis, 1980). This plastic behaviour explains
why such cements provide a good seal despite a high setting shrinkage and
thermal expansion of 35 x 10"6 X" 1 (Civjan & Brauer, 1964).
The hydrolytic instability of ZOE cements, arising from the weakness of
the zinc eugenolate chelate, has been discussed in Section 9.2.5. For this

333
Non-aqueous cements
reason these cements easily decompose under oral conditions. They can
survive, however, when used as a liner where they are not exposed to
aqueous conditions. Otherwise they are strictly temporary materials.
Indeed, the weakening of the cement may prove an advantage when they
are used for the temporary cementation of crowns.
One serious fault of these materials is that the presence of an electron-
rich phenolic hydroxyl group inhibits free-radical polymerization. Thus,
composite resins placed over them do not polymerize completely.

9.2.8 Biological properties


Pulp reaction is minimal: there is a slight reduction in odontoblasts (cells
responsible for dentine formation) but these recover in a few weeks
(Wilson, 1975b). The sealing ability and bactericidal action appear to
facilitate pulpal healing (Beagrie, Main & Smith, 1972). The release of
eugenol by hydrolysis of the zinc eugenolate matrix relieves pain in the
pulp in deep cavities (Smith, 1982a). There are, however, disadvantages
associated with the release of eugenol. Eugenol (and zinc also) is cytotoxic,
and causes toxic cell reactions (Coleman, 1962; Roydhouse & Weiss,
1964). Thus, eugenol is an irritant and can cause inflammatory responses
in soft tissues (Beagrie, Main & Smith, 1972) with haemolysis and delayed
healing (Smith, 1982b). It is also a potential allergen (Smith, 1982a). For
these reasons direct contact with connective tissues must be avoided.
For full accounts of biological responses the reader is referred to reviews
by Brauer (1965), Helgeland (1982) and Smith (1982b).

9.2.9 Modified cements


The ZOE cements are susceptible to modification. Modification by the
addition of accelerators has already been discussed in Sections 9.2.4 and
9.2.5.
One of the oldest additives is rosin (abietic acid) which improves
working, hardening rate and strength (Wallace & Hansen, 1939; Molnar &
Skinner, 1942). El-Tahawi & Craig (1971) report that hydrogenated rosin
inhibits the formation of crystallites.
Some materials intended for temporary cementation and cavity lining
are formulated as two pastes. One paste is formed by blending the zinc
oxide powder with a mineral or vegetable oil and the other by mixing
an inert filler into the liquid. These cements are much weaker, with

334
Zinc oxide eugenol (ZOE) cements
compressive strengths between 1-7 and 7 MPa (Gilson & Myers, 1970).
This strength is adequate for the temporary cementation of some
restorations. Mechanical properties are reduced by water immersion
although the effect is much less than with simple cements. Mechanical
retention is less than for zinc phosphate cements (Grieve, 1969; Richter,
Mitchem & Brown, 1970) but an 83*5% success rate has been reported
(Silvey & Myers, 1976).
Sometimes antimicrobial agents such as thymol or 8-hydroxyquinoline
may be present (Wilson, 1975b). The latter is also capable of forming a
cementitious chelate with zinc.

9.2.10 Impression pastes


The ZOE impression paste is used in taking impressions of the mouth prior
to constructing a denture. It is used as a corrective impression material after
a preliminary impression has been taken (Phillips, 1982a). A preliminary
impression lacks detail so it is necessary to take a secondary or corrective
impression which is placed on a tray that has the contours of the
preliminary impression.
The ZOE impression paste is essentially a two-paste ZOE cement. One
paste is formed by plasticizing the zinc oxide powder with 13 % of mineral
or vegetable oil. The other paste consists of 12 % eugenol or oil of cloves,
50% polymerized rosin, 20% silica filler, 10% resinous balsam (to
improve flow) and 5 % calcium chloride (accelerator).

9.2.11 Conclusions
The ZOE cement has been in use for over 100 years and is still popular
because it is easy to prepare and handle and has palliative and antiseptic
properties. Its plastic nature ensures that it adapts well and even its
weakness is an advantage in its applications as a temporaryfillingmaterial
and cementing agent. In these applications a cement must be sufficiently
strong for the restoration to be held in place for a limited period, but not
so strong as to make removal difficult. Although not by intention, the
cement happens to be prepared from two medicaments, eugenol and zinc
oxide, and these impart anodyne and antibacterial properties. Moreover,
the cement is a sustained-release device for eugenol and zinc. These
accidental benefits make it a difficult material to supplant.
Surprisingly, although the material has been in use for over 100 years,
the cement structure has yet to be fully elucidated.

335
Non-aqueous cements

9.3 Improved ZOE cements


9.3.1 General
The main disadvantage of the ZOE cements is their low strength. The
following approaches have been made to remedy this disadvantage.
(1) Incorporation of reinforcing fillers into the powder
(2) Addition of reinforcing polymers to the liquid
(3) Replacement of eugenol by other chelating agents
(4) Replacement of zinc oxide by other oxides

9.3.2 Reinforced cements


Several attempts have been made to improve the strength of ZOE cements
either by adding fillers to the powder or by dissolving resin in the liquid
(Wilson, 1975b; Smith, 1982a). Examples of fillers used include rosin,
hydrogenated rosins, poly(methyl methacrylate), polystyrene, polycar-
bonate, fused silica and dicalcium hydrogen phosphate.
Messing (1961) found that he could improve ZOE cements by adding
10 % polystyrene to the liquid. Strength developed more rapidly than in an
unmodified but accelerated cement and after 24 hours reached 42 MPa
compared with 36 MPa.
The most important approach was to use poly(methyl methacrylate),
PMMA, in formulations, either as a particulate filler or as a coating on zinc
oxide particles (Jendresen & Phillips, 1969; Jendresen et ai, 1969; Civjan
et al, 1972). It is claimed that such materials can be used for permanent as
well as temporary cementation.
ZOE cements intended for permanent cementation are required by the
International Standard (ISO, 1988) to set in 4 to 10 minutes with a film
thickness of 25 um.
The compressive strength of these reinforced ZOE cements ranges from
36 to 52 MPa (Table 9.3). This compares with a strength of 13 to 38 MPa
for unreinforced cements. ZOE cements are required by the International
Standard (ISO, 1988) to have a minimum compressive strength of 35 MPa
if they are to be used for permanent cementation and 25 MPa if they are
intended for use as a base or for temporary restoration.
The tensile strength of these materials is very much lower and lies
between 2-6 and 4-2 MPa (Table 9.3) and their modulus of elasticity varies
from 2-1 to 3-0 GPa (Powers, Farah & Craig, 1976). All these strength

336
2-ethoxybenzoic acid eugenol (EBA) cements
Table 9.3. Strength (MPa) ofZOE cements

Compressive Tensile
ZOE simple 1-2-2-8 [1,6]
ZOE accelerated 13-38 [1,2] 1-2-2-1 [6]
ZOE reinforced 36-52 [1-4] 2-6-^-2 [4, 6, 7]
ZOE pastes 1-7-7 [5] —
EBA reinforced 40-70 [4, 8] 5-5-7-0 [4, 6, 7]

[1] Brauer (1972); [2] Wilson (1975b); [3] Jendresen et al (1969); [4]
Powers, Farah & Craig (1976); [5] Gilson & Myers (1970); [6] Hannah
& Smith (1971); [7] Williams & Smith (1971); [8] 0ilo & Espevik
(1978).

figures have to be accepted with some reservations, for they represent


measurements made at room temperature and it is certain that these
cements are very much weaker at oral temperatures (0ilo & Espevik,
1978).

9.4 2-ethoxybenzoic acid eugenol (EBA) cements


9.4.1 General
The term EBA cement is not quite exact, but is convenient to use and has
been generally accepted. Originally this cement was a variant of the ZOE
cement - its most important variant - and was based on a liquid which was
a mixture of 2-ethoxybenzoic acid (EBA) and eugenol. More recently,
eugenol has been replaced by other compounds of similar structure. All
these cements contain EBA as a major constituent.
The structure of EBA has been examined using infrared spectroscopy
(Bagby & Greener, 1985) and there are apparently three conformations,
two hydrogen-bonded (Figure 9.6).

9.4.2 Development
In Section 9.1 it was noted that this cement was the only important zinc
oxide cement other than the ZOE cement. Its invention and development
is largely associated with Brauer who, with coworkers, has been carrying
out a programme of research and development from the late 1950s to the
present day.
In 1958, in an attempt to improve on the ZOE cement, Brauer, White &

337
Non-aqueous cements
Table 9.4. Composition of the EBA cement (Brauer, McLaughlin
& Huget, 1968)

Liquid: 62-5 % 2-ethoxybenzoic acid


37-5 % eugenol
Powder: 64% zinc oxide
30 % tabular alumina (particle size from < 1 urn to 20 urn,
with few particles > 20 urn)
6 % hydrogenated rosin

Moshonas (1958) investigated the reactions between zinc oxide and a large
number of chelating agents. Of these, EBA proved to be the most
promising. They then examined cement formation between EBA and
various metal oxides. Cement formation was found with MgO, CaO, BaO,
ZnO, CdO, HgO and PbO.
Finding that pure EBA cements tended to be unduly water soluble, these
workers went on to study cement formation between zinc oxide and
mixtures of EBA and eugenol. Small amounts of EBA added to eugenol
produced a marked acceleration in cement formation and reduced the
setting time from 2 hours to 3 minutes. Setting time remained short in the
range 25 to 75 % EBA. From this basis, Brauer went on to develop what he
termed the EBA cement. Later studies showed that the optimum liquid had
the composition 62-5 % EBA and 37*5 % eugenol, the EBA-eugenol liquid.
This liquid is important for it was used in most of the later experimental
studies and in commercial examples (Table 9.4).
The weakness of this cement was its tendency to dissolve in water. This
was prevented by including rosin (mainly abietic acid) or hydrogenated
rosin in the formulation (Brauer, Simon & Sangermano, 1962). Rosin and
fused quartz or calcium hydrogen phosphate monohydrate were added to

,0 H—0

0 — H-

C2H5 C2H5
Figure 9.6 The structure of 2-ethoxybenzoic acid, showing the two hydrogen-bonded
configurations (Bagby & Greener, 1985).

338
2-ethoxybenzoic acid eugenol (EBA) cements
the zinc oxide powder as strengthening fillers. The highest compressive
strengths were obtained with the EBA-eugenol liquid and two powders
that contained respectively 10% rosin (71 MPa), and 6% rosin and 10 to
30% quartz (74 to 81 MPa). These are stronger than the strongest of the
reinforced ZOE cements (55 MPa).
In a further attempt to improve properties, Brauer, McLaughlin &
Huget (1968) examined the use of alumina as a reinforcing filler. Alumina
is considerably more rigid than fused quartz. They achieved a considerable
increase in strength. The preferred composition was the powder defined in
Table 9.4, which had a compressive strength of 91 MPa. This zinc oxide
based powder was the one most commonly used in subsequent studies by
Brauer and coworkers. We shall refer to it as the EBA powder for it is the
one used in commercial formulations and in a number of experimental
studies.
Although these materials had a high early (10-minute) compressive
strength of 46 MPa, their brittleness limited their use for the temporary
restoration of multiple surfaces subject to heavy masticatory forces (Ciyjan
& Brauer, 1964). Stress bearing was improved by incorporating powdered
polymethacrylate polymers of low elastic moduli into the zinc oxide
powder (Brauer, Huget & Termini, 1970). One such cement used a powder
containing 58-2% zinc oxide, 27*3% alumina, 5-4% rosin and 9*1%
methylmethacrylate copolymer. The compressive strength was 65 MPa
and the tensile strength high at 11 MPa.
Despite all these endeavours, as Brauer himself admits, the rapid
disintegration under oral conditions prevents their use in permanent
restorations (Brauer, Stansbury & Argentar, 1983) and subsequent
development has taken place in a quite different direction.

9.4.3 Setting and structure


Little is known of the setting reaction and structure of EBA cement. The
absence of an infrared band at 1750 cm"1 in the set cement indicates that no
unreacted COOH is present (Brauer, 1972). So far, it is not certain whether
zinc forms a six-membered chelate or merely a simple salt with EBA.
Neither infrared spectroscopy nor solution studies are able to distinguish
between these two forms. Eugenol is much less readily extracted and so
more firmly bound in the complex than is EBA. The suspicion is that the
EBA cement is fundamentally more prone to hydrolysis than the ZOE
cement.

339
Non-aqueous cements
The only other studies are contained in one brief paper by Wilson &
Mesley (1974). The infrared spectrum of the pure EBA cement containing
excess zinc oxide was found to correspond to a salt of EBA. The binding
matrix was amorphous save for a trace of unidentified crystalline material.
Infrared spectroscopy indicated that cements prepared from EBA-eugenol
mixtures were not a simple mixture of the two parent cements. The mixed
cements contained three crystalline phases: zinc eugenolate and two
unidentified phases. Probably, one of these unidentified phases is a zinc 2-
ethoxybenzoate and the other a zinc eugenolate-2-ethoxybenzoate chelate
salt. Different results were obtained when zinc oxide was present only in
stoichiometric amounts. Then the evidence was that only two crystalline
phases were present: a zinc eugenolate and a zinc 2-ethoxybenzoate. The
indications are that there is an equilibrium:
ZnO
eugenolate + 2-ethoxybenzoate • eugenolate-2-ethoxybenzoate
Thus, if zinc 2-ethoxybenzoate is a weak chelate it will be preferentially
extracted and the reaction will move to the left. Eventually the matrix will
contain only eugenol, as, indeed, Brauer (1972) found.

9.4.4 Properties
These cements have unusual rheological properties (Wilson, 1975b). They
can be mixed to higher powder/liquid ratios (6:1 by mass, or more) than
any other dental cements and are very fluid. Whereas pastes of other
cements behave as plastic bodies, the EBA cement has the characteristics
of a very viscous Newtonian liquid and flows under its own weight, even
when mixed very thickly (Wilson & Batchelor, 1971). High powder/liquid
ratios are required for optimum properties: 3-5 g cm"3 for luting and 5 to
6 g cm"3 for linings and bases.
The linear setting shrinkage is 0-24 to 0-52% (dry) and 0-12 to 0-38%
(wet), which compares withfiguresof 0-85 % (wet) and 0-31 % (dry) found
for a ZOE cement by Civjan & Brauer (1964). The working and setting
times of EBA cements range from 7 to 13 minutes (Smith, 1982a) and are
dependent on both humidity and temperature. Film thickness ranges from
40 to 70 |im which is greater than the 25 |im required by the International
Standard (ISO, 1988). Hembree, George & Hembree (1978) found that
under simulated clinical conditions an alumina EBA cement always gave a
greater film thickness (c. 50 Jim) than the 20 jim of the unfilled material.

340
2-ethoxybenzoic acid eugenol (EBA) cements
Table 9.5. Strength (MPa) of reinforced EBA cements

Compressive Tensile
ZOE 36-52 [1-4] 2-6-4-2 [4, 6, 7]
EBA 40-70 [4, 8] 5-5-7-0 [4, 6, 7]
EBA-HV 42-60 [9] 5-0-6-3 [9]
EBA-di-HV 48-70 [10] 6-2-7-5 [10]
EBA-poly-HV 67-70 [10] 5-8-6-8 [10]
Glass-EBA-poly-HV 64-73 [10] 9-9-11-2 [10]
EBA polymer cement 73-112 [11] 101-15-8 [11]

[I] Brauer (1972); [2] Wilson (1975b); [3] Jendresen et al (1969);


[4] Powers, Farah & Craig (1976); [6] Hannah & Smith (1971);
[7] Williams & Smith (1971); [8] 0ilo & Espevik (1978); [9] Brauer,
Stansbury & Argentar (1983); [10] Stansbury & Brauer (1985);
[II] Brauer & Stansbury (1984b).

EBA cements have marked viscoelastic characteristics. They creep under


load to an even greater extent than the ZOE cements (Wilson & Lewis,
1980). When subject to a slowly increasing load they exhibit marked strain
at fracture and low strength (0ilo & Espevik, 1978). These characteristics
may be the reason why retention values for crowns and orthodontic bands,
although better than other ZOE cements, are inferior to those of the zinc
phosphate cements (Williams, Swartz & Phillips, 1965; Grieve, 1969;
Richter, Mitchem & Brown, 1970). Thermal expansion is 60 to
90 x 10~6 °C-\ which is higher than that of the ZOE cement (Civjan &
Brauer, 1964).
Compressive strength depends on powder/liquid ratio and test con-
ditions. At the thin cementing consistency, the compressive strength of
these materials ranges from 40 to 70 MPa (Powers, Farah & Craig, 1976;
0ilo & Espevik, 1978). Tensile strength is much lower: 4-7 to 7-1 MPa
(Williams & Smith, 1971; Powers, Farah & Craig, 1976; Hannah & Smith,
1971). The modulus of elasticity is 5400 MPa (Powers, Farah & Craig,
1976).
In vitro studies by Wilson et al. (1986) using the impinging jet method
show the EBA cement to be the least resistant of all the cement types to
erosion in neutral solution. Clinical studies confirm this result and show
that there is greater dissolution in the mouth than for other dental cements
(Mitchem & Gronas, 1978; Osborne et al, 1978; Andrews and Hembree,
1976; Shilling, 1977). Despite this, a clinical survey by Silvey and Myers

341
Non-aqueous cements
(1976,1977) indicated that the performance of an alumina-reinforced EBA
cement over 3 years was only slightly worse than that of zinc phosphate
and polycarboxylate cements.
Results for cement strengths are summarized in Table 9.5.

9.5 EBA-methoxyhydroxybenzoate cements


9.5.1 EBA-vanillate and EBA-syringate cements
All cements that contain eugenol inhibit the polymerization of acrylates,
and those of EBA-eugenol are no exception. In order to remedy this and
other defects, Brauer and his coworkers examined alternatives to eugenol
(Figure 9.7). These included the esters of vanillic acid (3-methoxy-4-
hydroxybenzoic acid, HV) and syringic acid (3,5-dimethoxy-4-hydroxy-
benzoic acid). Both are 3-methoxy-4-hydroxy compounds and are thus
chemically related to eugenol and guaiacol. Both are solids and have to be
dissolved in EBA where they form satisfactory cements with EBA zinc
oxide powder. The vanillate (EBA-HV) cements are the more important.

EBA-vanillate cements
Using the EBA powder (Table 9.4) and liquids containing 12-5 to 18-3 % of
a vanillate ester (either n-hexyl, n-heptyl or n-decyl) Brauer, Stansbury &
Argentar (1983) obtained cements that set in 4-5 to 5-5 minutes with
compressive strengths from 42 to 60 MPa and tensile strengths from 5-0
to 6-3 MPa (Table 9.5). The preferred liquid was one containing 87-5%
eugenol and 12-5% HV. Brauer & Stansbury (1984a) found also that the
EBA-HV cements bonded much more strongly to composite resins
(5-5 MPa), stainless steel (4 MPa), nickel-chrome (5 MPa) and porcelain
(4-1 to 5-5 MPa) than the ZOE cement (Table 9-6).

COOR COOR

OCHq H 3 CO OCH3

(a)
Figure 9.7 Structure of (a) vanillates and (b) syringates.

342
EBA-methoxyhydroxybenzoate cements
Table 9.6. Bond strength (MPa) to substrates

Composite resin Stainless steel

ZOE 0-3 [12] 0-6 [12]


EBA-HV 5-5 [12] 4-1 [12]
EBA-poly HV 3-8-6-9 [10] 5-9-7-9 [10]
Silica EBA-poly HV 60-7-9 [10] 7-1-9-4 [10]
EBA polymer cement 4-1-10-3 [11] 9-8-15-6 [11]

[10] Stansbury & Brauer (1985); [11] Brauer & Stansbury (1984b);
[12] Brauer & Stansbury (1984a).

EBA-syringate cements
With a liquid containing 7 to 14% n-hexyl syringate, Brauer & Stansbury
(1984a) obtained cements that set in 4 minutes with a compressive strength
from 54 to 62 MPa and a tensile strength of 5-5 MPa, but they were brittle
(Table 9.5). Replacement of n-hexyl syringate by 2-ethylhexyl syringate
yielded cements that, depending on powder/liquid ratio, set in 6 to 9-5
minutes with compressive strengths of 40 to 50 MPa and tensile strengths
of 5-2 to 5-7 MPa. These cements were less brittle than those of n-hexyl
syringate.
The best formulation proved to be one based on a liquid containing 88 %
EBA, 5% n-ethylhexyl syringate and 7% n-hexyl vanillate. Cements
prepared from these liquids set in 5-5 to 6-5 minutes with a compressive
strength of 66 MPa and tensile strength of 6 to 7 MPa.
All these vanillate and syringate cements are about as strong as those of
EBA-eugenol.

Setting
There is little information available on their setting and structure. Bagby &
Greener (1985) used Fourier transform infrared spectroscopy (FTIR) to
examine the cement-forming reaction between zinc oxide and a mixture of
EBA and n-hexyl vanillate. Although they found evidence for reaction
between zinc oxide and EBA, they were unable to find any for reaction
between zinc oxide and n-hexyl vanillate because of peak overlaps, the
minor concentration of n-hexyl vanillate and the subtle nature of the
spectral changes.

343
Non-aqueous cements

Advantages
Brauer (1988) considers that EBA-HV cements possess a number of
advantages over eugenol cements. They bond much more strongly to
composite resins and stainless steel than does the ZOE cement. They are
compatible with composite resins for, unlike eugenol, vanillates do not
inhibit the polymerization of acrylate polymers, because the phenolic
hydroxyl is electron-poor. Brauer, Stansbury & Argentar (1983) speculate
that they are probably less toxic, as n-hexyl vanillate has been considered
as a food preservative. Vanillates should yield cements with bactericidal
properties. But all these supposed biological advantages need to be
substantiated. One recent evaluation by Keller et al. (1988) has shown that
this cement has an acceptable performance when compared with the
clinically acceptable zinc oxide eugenol and zinc phosphate cements.
Syringate cements possess similar advantages to the vanillate cements.
In addition, syringic acid possesses cariostatic properties, so syringates
may inhibit the development of caries (dental decay). Again these
advantages need to be confirmed.
Modifications
Brauer, Stansbury & Flowers (1986) modified these cements in several
ways. The addition of various acids - acetic, propionic, benzoic etc. -
accelerated the set. The use of zinc oxide powders coated with propionic
acid improved mixing, accelerated set, reduced brittleness and increased
compressive strength from 63 to a maximum of 72 MPa. The addition of
plasticizing agents such as zinc undecenylate yielded flexible materials.
Incorporation of metal powders had a deleterious effect and greatly
increased the brittleness of these cements. The addition offluorideswas not
very successful, for fluoride release was not sustained.

9.5.2 EBA-divanillate and polymerized vanillate cements


Stansbury & Brauer (1985) investigated the use of divanillates in EBA
cements (Figure 9.7). Divanillates consist of two vanillate groups linked by
esterifying the COOH groups with diols. Cements based on EBA
containing 10 to 11% of divanillates set in 4-5 to 5 minutes with
compressive strengths of 48 to 70 MPa and tensile strengths of 6-2 to
7-5 MPa using the EBA zinc oxide powder (Table 9.5).
In addition, Stansbury & Brauer (1985), taking advantage of the fact
that vanillates do not inhibit free-radical polymerization, incorporated

344
EBA-methoxyhydroxybenzoate cements
polymerizable vanillates in cement formulations.They used the methyl-
acryloylethyl, -CH 2 . OOC. C(CH3)=CH2, ester of vanillic acid (Figure
9.7), and added 3 to 10 % to the EBA liquid. In this system the zinc oxide
powder contained 1 % benzoyl peroxide as a polymerization initiator and
the accelerator, 7V,7V-dihydroxyethyl-/?-toluidine, was added to the liquid.
These cements set in 6-5 to 8-5 minutes with compressive strengths of 67 to
70 MPa and tensile strengths of 5-8 to 6-8 MPa (Table 9.6). Thus, these
cements are no stronger than other vanillate cements and they are brittle.
They possess adhesive properties: bond strengths of 5*9 to 7-9 MPa to
stainless steel and 3-8 to 6-9 MPa to composite resins were recorded (Table
9.6).
Some improvement in physical properties was obtained by adding a filler
of silanized glass (325 mesh) to the EBA zinc oxide powder (1:1). Although
the compressive strengths of these reinforced cements were no greater than
those of the unreinforced cement (64 to 73 MPa) there were significant
improvements in tensile strengths with values of 9-9 to 11-2 MPa being
obtained (Table 9.5). There was improved adhesion, with bond strengths
of 7-1 to 9-4 MPa to stainless steel and 6-0 to 7-9 MPa to composite resins
(Table 9.6).

9.5.3 EBA-HVpolymer cements


The last stage in the development of the EBA cement is represented by the
polymer cements. Brauer & Stansbury (1984b), taking advantage of the
fact that the EBA-HV liquid does not inhibit vinyl polymerization,
included methacrylates into the cement composition. The object was to
produce a material that set after mixing, both by polymerization and by
salt or chelate formation.
The liquids used were 1:1 mixtures of EBA-HV and liquid methacrylate
which also contained dihydroxyethyl-/?-toluidine as the accelerator. Both
mono- and di-methacrylates were used. The benzoyl peroxide initiator was
included in the EBA zinc oxide/silanized (1:1) glass powder. These
polymer cements set 5 to 10 minutes after mixing. Since there is a
substantial amount of monomer in the liquid (50%) the contribution of
the polymer to the strength of the cement must be considerable. Brauer &
Stansbury (1984b) suggested that the two matrices, the polymer matrix and
the salt matrix, may be interpenetrating; but separation of the two phases
is likely.
Very high strengths were obtained compared with other ZOE and EBA

345
Non-aqueous cements
cements. The most important monomethacrylates studied were methyl-,
cyclohexyl- and dicyclophenyl- and mixtures of them. Cements containing
these monomers had compressive strengths ranging from 73 to 112 MPa
and tensile strengths from 10-1 to 15-8 MPa (Table 9.5).
Good bonding was obtained to several substrates under aqueous
conditions. Values obtained were 4-1 to 10-3 MPa to composite resins, and
9-8 to 15-6 MPa to stainless steel (Table 9.6). They were also reported as
adhering to porcelain. No adhesion was obtained to untreated dentine or
enamel. The cements could be bonded to enamel etched with acid (3-5 MPa)
and to dentine conditioned with poly(acrylic acid) (1-0 MPa).
The mechanism of adhesion to various substrates has not been fully
explained. Brauer & Stansbury (1984b) consider that bonding to composite
resins occurs by the diffusion of methacrylate polymer chains into the
resin. Bonding to base metals is, perhaps, by salt or chelate bridges. Here
it is significant that ZOE cements do not bond, so perhaps bonding is due
to the action of free EBA on the substrate. The adhesion to porcelain is
surprising. Porcelain is inert so that the attachment can hardly be chemical.
Also, it would be expected that if a cement adheres to porcelain then it
should adhere to untreated enamel and dentine, but this is not so.
The use of dimethacrylates led to even greater cement strengths.
Compressive strengths ranged from 50 to 199 MPa, although the higher
strengths were obtained with pastes that were hardly workable and
132 MPa represents the practical limit of compressive strength. Tensile
strength ranged from 8-0 to 15-9 MPa. Unfortunately, the EBA-
dimethacrylate liquids were unstable and partial polymerization occurred
within hours.
The biological properties of these materials are unknown, but the
presence of methacrylate monomers must adversely affect their bio-
compatibility.
Brauer & Stansbury (1984b) claim that these materials can be used in
'intermediate' restorations which are used in holding-type situations
where an extensive cleaning-up regime is required over many weeks prior
to the placement of a permanent restorative.

9.5.4 Conclusions
We have seen how Brauer and coworkers have developed EBA cements
since 1958, during which time they have become increasingly more
complex. They are somewhat stronger than reinforced ZOE cements. EBA

346
Calcium hydroxide chelate cements
cannot be used alone but always requires the addition of a 2-methoxy-
phenol. Chemical differences between the different 2-methoxyphenols have
little effect on cement strength. Non-eugenol EBA cements have the
distinct advantage, however, in not inhibiting vinyl polymerization and are
thus compatible with composite resins. This is an important attribute
because these cements are used as bases under composite resins. It also
means that methacrylate monomers can be mixed with EBA to give
polymer cements. These materials are considerably stronger than plain
cements.
Unfortunately, although EBA cements have been subjected to a
considerable amount of development, this work has not been matched by
fundamental studies. Thus, the setting reactions, microstructures and
molecular structures of these EBA cements are still largely unknown. In
addition, the mechanism of adhesion to various substrates has yet to be
explained. Such knowledge is a necessary basis for future developments.

9.5.5 Other zinc oxide cements


Skinner, Molnar & Suarez (1964) studied the cement-forming potential of
28 liquid aromatic carboxylic acids with zinc oxide. Twelve yielded
cohesive products of some merit. Of particular interest were cements
formed with hydrocinnamic, cyclohexane carboxylic, ^-tertiary butyl-
benzoic, thiobenzoic and cyclohexane butyric acids. One of these cements
is on the market as a non-eugenol cement. It is very weak with a
compressive strength of 40 MPa, a tensile strength of 11 MPa and a
modulus of 177 MPa, and is only suitable as a temporary material (Powers,
Farah & Craig, 1976).

9.6 Calcium hydroxide chelate cements


9.6.1 Introduction
Pastes of calcium hydroxide with water have been used as pulp-capping
materials for many years and it is the material of choice for this application
(Granath, 1982). Its favourable tissue responses have been known for
many years (Zander, 1939). It has a healing effect, for it induces the
formation of hard tissues of reparative dentine when pulp has been
exposed (Eidelman, Finn & Koulourides, 1965). This action seems to be
associated with its high alkalinity (pH ~ 12-5) and consequent bactericidal
and proteinlysing effect (Fisher, 1977).

347
Non-aqueous cements

Table 9.7. Composition of a calcium hydroxide chelate cement {American


Dental Association, 1977)

Basic paste: 51 % calcium hydroxide


9-23 % zinc hydroxide
0-29 % zinc stearate
in JV-ethyl toluene sulphonamide
A cid paste: 13-8% titanium dioxide
31*4% calcium sulphate
15-2% calcium tungstate
in 1-methyl trimethylene disalicylate butane-1-3-diol ester

The manipulation of calcium hydroxide paste is not easy, however, and


Dougherty (1962) introduced the calcium hydroxide salicylate cements.
These are based on the reaction between calcium hydroxide and salicylate
esters and come in two-paste packs which are easy to mix in the dental
surgery.

9.6.2 Composition
All commercial materials are based on calcium hydroxide and liquid alkyl
salicylates (Prosser, Groffman & Wilson, 1982) and are supplied as a two-
paste pack. Zinc oxide is sometimes added to the calcium hydroxide, as are
neutral fillers. A paste is formed from this powder by the addition of a
plasticizer; examples include Af-ethyl toluenesulphonamide (p- or/?-) and
paraffin oil, with sometimes minor additions of polypropylene glycol. The
other paste is based on an alkyl salicylate as the active constituent
containing an inorganic filler such as titanium dioxide, calcium
sulphate, calcium tungstate or barium sulphate. Alkyl salicylates used
include methyl salicylate, isobutyl salicylate, and 1-methyl trimethylene
disalicylate. An example of one commercial material, Dycal, is given in
Table 9.7, but its composition has been subjected to change over the years.

9.6.3 Setting
Prosser, Stuart & Wilson (1979) and Prosser, Groffman & Wilson (1982)
examined the setting of a number of these cements using infrared
spectroscopy. The infrared spectrum of the alkyl salicylates showed an
O-H stretch band at 3190 cm"1 and a C-O stretch band at 1675-95 cm"1,

348
Calcium hydroxide chelate cements
which were displaced from the normal frequencies of these groups because
of very strong intramolecular hydrogen-bonding conjugate chelation. This
conjugate chelation arises from resonance between the ester and its
enolized form, and is shown for an alkylsalicylate in Figure 9.8.
The essential chemical reaction is an acid-base one between calcium
hydroxide and the phenolic group of an alkylsalicylate. Cement formation
occurs as calcium replaces phenolic hydrogen. During setting, the O-H
stretch band diminished as a carboxylate band appeared at 1540-60 cm"1
(asymmetric stretch). The ester band at 1675-95 cm"1 diminished, showing
a conversion of C=O to C~O". The molecular structure can be represented
as a chelate containing two bidendate ligands (Figure 9.9). In this case,
unlike that of the zinc oxide eugenol cement, infrared spectroscopy can
provide proof of a chelate structure, although the setting reaction appears
to be very similar. The cement structure consists of an amorphous
disalicylate complex filled by the unreacted calcium hydroxide and other
inorganic fillers.
The case of the disalicylate, 1-methyl trimethylene disalicylate, is
interesting. Because of steric hindrance it is unlikely that the two salicylate
ligands can chelate to one calcium atom. In theory the disalicylate

Figure 9.8 Alkylsalicylate structure.

Figure 9.9 Chelate of calcium and an alkylsalicylate.

349
Non-aqueous cements
structure could bridge calcium ions and so form an ionically linked chain,
but this too is unlikely. All the calcium hydroxide cements are weak and
friable, indicating that the chelates are bound together by only weak
secondary forces. The coordination number of calcium is usually six, and
since two water molecules are generated for every calcium ion during the
course of the reaction, it is possible that these two water molecules are
attached to the central calcium ion to form an octahedral complex. This
structure would be similar to that proposed for the zinc eugenolate cement,
and water bridges may play a similar structural role.

9.6.4 Physical properties


These materials are prepared by mixing two strips of paste of equal length,
and when mixed formfluidpastes. Some set rapidly, but setting depends on
the availability of water (Plant & Wilson, 1970; Bryant & Wing, 1976a,b).
In contact with water one example set in 2 minutes, and in the absence of
water another did not set at all (Bryant & Wing, 1976b). Normally they set
in 3-5 to 8 minutes (Plant, Jones & Wilson, 1972).
They show marked plastic deformation under compressive load and
although this decreases as they harden they still deform rather than
fracture when 25 minutes old (Plant & Wilson, 1970). This probably
accounts for the observation that there is little incidence of fracture or
displacement when amalgams are placed on these cements (Fisher, 1977).
The cements are very weak and after 24 hours their compressive strengths
only range from 11 to 14 MPa (Bryant & Wing, 1976b).
These materials are hydrolytically unstable and weaken when stored in
water for a week (Bryant & Wing, 1976b). Prosser, Groffman & Wilson
(1982) found that calcium and hydroxide ions and salicylates were released
and that the rate of release was controlled by the plasticizer used in the
cement formulation. Hydrophilic sulphonamide plasticizers allowed ready
ingress of water and promoted decomposition, whereas the hydrophobic
hydrocarbon plasticizer repelled water and retarded hydrolytic decompo-
sition.

9.6.5 Biological properties


Hydrolytic decomposition of these cements is clinically advantageous.
Free calcium hydroxide is present in excess so that large amounts of
calcium are released which, together with high alkalinity, promotes

350
Calcium hydroxide chelate cements
sterilization and calcification of carious dentine (McWalter, El-Kafrawy &
Mitchell, 1976). There is formation of dentine bridges when they are used
for pulp capping.
Fisher (1977) found that the bactericidal effects varied from brand to
brand and Fisher & McCabe (1978) related this to chemical composition.
Only cements which give rise to high alkalinity (pH = 11) are effective.
These are the cements which are readily decomposed by water, and this
relates to the plasticizer used. Hydrophilic plasticizers are required if these
cements are to be clinically effective.
Hydrolytic decomposition brings disadvantages. There is continued
leakage at the margins where complete dissolution can occur (Gourley &
Rose, 1972) and, indeed, these bases have been observed to disappear
entirely (Akester, 1979; Barnes & Kidd, 1979).
These cements are the materials of choice for pulp capping (a wound
dressing for covering an exposed or surgically treated pulp). They are
superior to zinc oxide eugenol cements for this purpose (Mjor, 1963;
Paterson, 1976).
These materials also protect the pulp against invasion by acids from
overlying dental cements of the phosphate or polyacrylate type and act as
a barrier to the penetration of harmful chemicals such as the unpolymerized
methacrylates (Smith, 1982a).

9.6.6 The calcium hydroxide dimer cement


Cowan & Teeter (1944) reported a new class of resinous substances based
on the zinc salts of dimerized unsaturated fatty acids such as linoleic and
oleic acid. The latter is referred to as dimer acid. Later, Pellico (1974)
described a dental composition based on the reaction between zinc oxide
and either dimer or trimer acid. In an attempt to formulate calcium
hydroxide cements which would be hydrolytically stable, Wilson et al.
(1981) examined cement formation between calcium hydroxide and dimer
acid. They found it necessary to incorporate an accelerator, aluminium
acetate hydrate, A12(OH)2(CH3COO)4. 3H2O, into the cement powder.
These cements set in 3-5 to 56 minutes (at 37 °C). Infrared spectroscopy
showed that as the cement set there was loss of acid carbonyl groups and
OH groups associated with calcium hydroxide, and simultaneously
formation of ionic carboxylate groups and hydrogen-bonded OH groups.
Although these cements were capable of setting under water and were
impervious to aqueous attack they were not a success. They failed to

351
Non-aqueous cements

release calcium hydroxide into solution; consequently, there was no


alkaline reaction and hence no favourable biological responses.

References
Andrews, J. T. & Hembree, J. H. (1976). In vivo evaluation of the marginal
leakage of four inlay cements. Journal of Prosthetic Dentistry, 35, 532-7.
American Dental Association. (1977). Accepted Dental Therapeutics, 36th edn.,
p. 235.
Akester, J. (1979). Disappearing Dycal. British Dental Journal, 146, 369.
Bagby, M. & Greener, E. H. (1985). Infrared spectral analysis of EBA-hexyl
vanillate-ZnO cement. Dental Materials, 1, 86-8.
Barnes, I. E. & Kidd, E. A. M. (1979). Disappearing Dycal. British Dental
Journal, 141, 111.
Batchelor, R. F. & Wilson, A. D. (1969). Zinc oxide eugenol cements. I. The
effect of atmospheric conditions on rheological properties. Journal of Dental
Research, 48, 883-7.
Bayne, S. C. & Greener, E. H. (1985). ZnO cements: phase identification by
thermal analysis. Dental Materials, 1, 165-9.
Bayne, S. C , Greener, E. H., Lautenschlager, E. P., Marshall, S. J. & Marshall,
jr, G. W. (1986). Zinc eugenolate crystals: SEM detection and
characterization. Dental Materials, 2, 1-5.
Beagrie, G. S., Main, J. H. P. & Smith, D. C. (1972). Inflammatory reaction
evoked by zinc polyacrylate and zinc eugenate cements: a comparison. British
Dental Journal, 132, 351-7.
Blackman, L. C. F. (1962). Lattice defects and the sintering of oxides. Industrial
Chemist, 38, 620-6.
Blackman, L. C. F. (1963). Lattice defects and the sintering of oxides. Industrial
Chemist, 39, 23-6.
Bonastre, J. F. (1827a). De la combinaison des huiles volatiles de girofle et de
pimet de la Jamai'que, avec des alcalis et autres bases salifiables. Journal de
Pharmacie, 13, 464-76.
Bonastre, J. F. (1827b). De la combinaison des huiles volatiles de girofle et de
piment de la Jamai'que avec des alcalis. 2. Combinaison d'huiles volatiles de
girofle avec les oxides metalliques. Journal de Pharmacie, 13, 513-21.
Braden, M. & Clarke, R. L. (1974). Dielectric properties of zinc oxide-eugenol
type cements. Journal of Dental Research, 53, 1263-7.
Brauer, G. M. (1965). A review of zinc oxide-eugenol type filling materials and
cements. Revue Beige de Medecine Dentaire, 20, 323—64.
Brauer, G. M. (1972). Cements containing 2-ethoxybenzoic acid (EBA).
National Bureau of Standards Special Publication 354, pp. 101—11.
Brauer, G. M. (1988). Vanillate or syringate cements. Trends & Techniques, 5,
No. 1, 6.
Brauer, G. M., Argentar, H. & Durany, G. (1964). Ionization constants and
reactivity of isomers of eugenol. Journal of Research of the National Bureau of
Standards, 68A, 619-24.

352
References

Brauer, G. M., Argentar, H. & Stansbury, J. W. (1982). Cementitious dental


compositions which do not inhibit polymerization. US Patent 4,362,510,
December 7, 1982.
Brauer, G. M., Huget, E. F. & Termini, D. J. (1970). Plastic modified
o-ethoxybenzoic acid cements as temporary restorative materials. Journal
of Dental Research, 49, Supplement, 1487-94.
Brauer, G. M., McLaughlin, R. & Huget, E. F. (1968). Aluminum oxide as a
reinforcing agent for zinc oxide eugenol-o-ethoxybenzoic acid cements.
Journal of Dental Research, 47, 622-8.
Brauer, G. M., Simon, L. & Sangermano, L. (1962). Improved zinc
oxide-eugenol type cements. Journal of Dental Research, 41, 1096-102.
Brauer, G. M. & Stansbury, J. W. (1984a). Cements containing syringic acid
ester-o-ethoxybenzoic acid and zinc oxide. Journal of Dental Research, 63,
137^0.
Brauer, G. M. & Stansbury, J. W. (1984b). Intermediate restorative from
n-hexyl vanillate-EBA-ZnO-glass composites. Journal of Dental Research,
63, 1315-20.
Brauer, G. M., Stansbury, J. W. & Argentar, H. (1983). Development of high-
strength, acrylic resin-compatible adhesive cements. Journal of Dental
Research, 62, 366-70.
Brauer, G. M., Stansbury, J. W. & Flowers, D. (1986). Modification of cements
containing vanillate or syringate esters. Dental Materials, 2, 21-7.
Brauer, G. M., White, jr, E. E. & Moshonas, M. G. (1958). The reaction of
metal oxides with o-ethoxybenzoic acid and other chelating agents. Journal of
Dental Research, 37, 547-60.
Bryant, R. W. & Wing, G. (1976a). The rate of development of strength in base
forming materials for amalgam restorations. Australian Dental Journal, 21,
153-9.
Bryant, R. W. & Wing, G. (1976b). The effects of manipulative variables on
base forming materials for amalgam restorations. Australian Dental Journal,
21,211-16.
Chisholm, E. S. (1873). Discussion. Dental Register, 27, 517.
Civjan, S. & Brauer, G. M. (1964). Physical properties of cements based on zinc
oxide, hydrogenated resin, o-ethoxybenzoic acid and eugenol. Journal of
Dental Research, 43, 281-99.
Civjan, S., Huget, E. F., Wolfhard, G. & Waddell, L. S. (1972). Characteristics
of zinc oxide eugenol cements reinforced with acrylic resin. Journal of Dental
Research, 51, 107-14.
Coleman, G. (1962). A study of some antimicrobial agents used in oral surgery.
British Dental Journal, 113, 22-8.
Copeland, H. I., Brauer, G. M., Sweeney, W. T. & Forziati, A. F. (1955).
Setting reaction of zinc oxide and eugenol. Journal of Research of the National
Bureau of Standards, 55, 133-8.
Cowan, J. H. & Teeter, H. M. (1944). Salts of residual dimerized fat acids: a
new class of resinous substance. Industrial and Engineering Chemistry, 36,
148-52.

353
Non-aqueous cements

Crisp, S., Ambersley, M. & Wilson, A. D. (1980). Zinc oxide eugenol cements.
V. Instrumental studies of the catalysis and acceleration of the setting
reaction. Journal of Dental Research, 59, 44-54.
Dollimore, D. & Spooner, P. (1971). Sintering studies on zinc oxide.
Transactions of the Faraday Society, 67, 2750-9.
Dougherty, E. W. (1962). Dental cement material. US Patent 3,047,408.
Douglas, W. H. (1978a). The metal oxide/eugenol cements. I. The chelating
power of eugenol type molecules. Journal of Dental Research, 57, 800-4.
Douglas, W. H. (1978b). The metal oxide/eugenol cements. II. A diffuse
reflectance spectrophotometric study of the setting of zinc oxide and
magnesium oxide cements. Journal of Dental Research, 57, 805-9.
Eidelman, E., Finn, S. B. & Koulourides, T. (1965). Remineralization of carious
dentin treated with calcium hydroxide. Journal of Dentistry for Children, 32,
218-25.
El-Tahawi, H. M. & Craig, R. G. (1971). Thermal analysis of zinc
oxide-eugenol cement. Journal of Dental Research, 50, 430-5.
Fisher, F. J. (1977). The effect of three proprietary lining materials on micro-
organisms in carious dentine. An in vivo investigation. British Dental Journal,
143, 231-5.
Fisher, F. J. & McCabe, J. F. (1978). Calcium hydroxide base materials: an
investigation into the relationship between chemical structures and
antibacterial properties. British Dental Journal, 144, 341-4.
Flagg, J. F. (1875). Dental pathology and therapeutics. Dental Cosmos, 27,
465-9.
Gerner, M. M., Zadorozhnyi, B. A., Ryabina, L. V. & Batovskii, V. N. (1966).
Infrared spectra of eugenol and zinc eugenolate. Russian Journal of Physical
Chemistry, 40, 122-3 (translation).
Gilson, T. D. & Myers, G. E. (1970). Clinical studies of dental cements. III.
Seven zinc oxide eugenol cements used for temporarily cementing completed
restorations. Journal of Dental Research, 49, 14-20.
Gourley, J. M. & Rose, D. E. (1972). Cavity bases under liners. Journal of the
Canadian Dental Association, 38, 246.
Graddon, D. P. (1968). Divalent transition metal /Mtetone-enolate complexes as
Lewis acids. Coordination Chemistry Reviews, 4, 1-28.
Granath, L. E. (1982). Pulp capping materials. In Smith, D. C. & Williams,
D. F. (eds.) Biocompatibility of Dental Materials. Volume II. Biocompatibility
of Preventive Dental Materials and Bonding Agents, Chapter 11. Boca Raton:
CRC Press Inc.
Grieve, A. R. (1969). A study of dental cements. British Dental Journal, 127,
405-10.
Hannah, C. M. & Smith, D. C. (1971). Tensile strengths of selected dental
restorative materials. Journal of Prosthetic Dentistry, 26, 314-23.
Helgeland, K. (1982). In vitro testing of dental cements. In Smith, D. C. &
Williams, D. F. (eds.) Biocompatibility of Dental Materials. Volume II.
Biocompatibility of Preventive Dental Materials and Bonding Agents, Chapter
9. Boca Raton: CRC Press Inc.

354
References

Hembree, J. H., George, T. A. & Hembree, M. E. (1978). Film thickness of


cements beneath complete crowns. Journal of Prosthetic Dentistry, 39, 533-5.
ISO. (1988). International Standard, ISO 3107. Dental zinc oxide/eugenol
cements and zinc oxide non-eugenol cements.
Jendresen, M. D. & Phillips, R. W. (1969). A comparative study of four zinc
oxide eugenol formulations as restorative materials. Part II. Journal of
Prosthetic Dentistry, 21, 300-9.
Jendresen, M. D., Phillips, R. W., Swartz, M. L. & Norman, R. D. (1969). A
comparative study of four zinc oxide eugenol formulations as restorative
materials. Part I. Journal of Prosthetic Dentistry, 21, 176-83.
Keller, J. C , Hammond, B. D., Kowlay, K. K. & Brauer, G. M. (1988).
Biological evaluation of zinc hexyl vanillate cement using two in vivo test
methods. Dental Materials, 4, 341-50.
King, J. (1872). Treatment of exposed pulps. Dental Cosmos, 14, 193-4.
Lee, V. J. & Parravano, G. (1959). Sintering reactions on zinc oxide. Journal of
Applied Physics, 30, 1735-^0.
McWalter, G. K., El-Kafrawy, A. H. & Mitchell, D. F. (1976). Long term study
of pulp capping in monkeys with three agents. Journal of the American Dental
Association, 93, 105-111.
Marshall, P. A., Enrigh, D. P. & Weyl, W. A. (1952). On the mechanism of
sintering and recrystallization of oxides. In The Proceedings of the
International Symposium on the Reactivity of Solids, pp. 273-84. Gothenburg.
Messing, J. J. (1961). A polystyrene-fortified zinc oxide/eugenol cement.
Investigation into its properties. British Dental Journal, 110, 95-100.
Mitcham, J. C. & Gronas, D. G. (1978). Clinical evaluation of cement solubility.
Journal of Prosthetic Dentistry, 40, 453-6.
Mjor, I. A. (1963). The effects of calcium hydroxide zinc oxide/eugenol and
amalgam on pulp. Odontologisk Tidsskrift, 71, 94-105.
Molnar, E. J. (1942). Cloves, oil of cloves and eugenol. Their medico-dental
history. Dental Items of Interest, 64, 521-8.
Molnar, E. J. (1967). Residual eugenol from zinc oxide-eugenol compounds.
Journal of Dental Research, 46, 645-9.
Molnar, E. J. & Skinner, E. W. (1942). A study of zinc oxide-rosin cements. I.
Some variables which affect hardening time. Journal of the American Dental
Association, 29, 744-51.
Nagoe, M. & Morimoto, T. (1969). Differential heat of adsorption and entropy
of water absorbed on zinc oxide surface. Journal of Physical Chemistry, 73,
3809-14.
Nielsen, T. H. (1963). The ability of 39 chelating agents to form cements with
metal oxides, respecting their usability as root-filling materials. Ada
Odontologica Scandinavica, 21, 159-74.
Norman, R. D., Phillips, R. W., Swartz, M. L. & Frankiewicz, T. (1964). The
effect of particle size on the physical properties of zinc oxide-eugenol
mixtures. Journal of Dental Research, 43, 252-62.
0ilo, G. & Espevik, S. (1978). Stress/strain behaviour of some dental luting
cements. Acta Odontologica Scandinavica, 36, 45-9.

355
Non-aqueous cements

Osborne, J. W., Swartz, M. L., Goodacre, C. J., Phillips, R. W. & Gale, E. M.


(1978). A method for assessing the clinical solubility and disintegration of
luting cements. Journal of Prosthetic Dentistry, 40, 413-17.
Paterson, R. C. (1976). The reaction of the rat molar pulp to various materials.
British Dental Journal, 140, 93-6.
Pellico, H. A. (1974). Settable dental compositions. US Patent 3,837,865.
Phillips, R. W. (1982a). Skinner's Science of Dental Materials, Chapter 7.
Philadelphia: W. B. Saunders.
Phillips, R. W. (1982b). Skinner's Science of Dental Materials, Chapter 29.
Philadelphia: W. B. Saunders.
Plant, C. G., Jones, I. H. & Wilson, H. J. (1972). Setting characteristics of lining
and cementing materials. British Dental Journal, 133, 21-74.
Plant, C. G. & Wilson, H. J. (1970). Early strengths of lining materials. British
Dental Journal, 129, 269-74.
Powers, J. M., Farah, J. W. & Craig, R. G. (1976). Modulus of elasticity and
strength properties of dental cements. Journal of the American Dental
Association, 92, 588-91.
Prosser, H. J., Groffman, D. M. & Wilson, A. D. (1982). The effect of
composition on the erosion properties of calcium hydroxide cements. Journal
of Dental Research, 61, 1431-5.
Prosser, H. J., Stuart, B. & Wilson, A. D. (1979). An infra-red spectroscopic
study of the setting reaction of a calcium hydroxide dental cement. Journal of
Materials Science, 14, 2894-900.
Prosser, H. J. & Wilson, A. D. (1982). Zinc oxide eugenol cements. VI. Effect of
zinc oxide type on the setting reactions. Journal of Biomedical Materials
Research, 16, 585-98.
Richter, W. A., Mitchem, J. C. & Brown, D. (1970). The predictability of
retentive value of dental cements. Journal of Prosthetic Dentistry, 24, 298-303.
Roydhouse, R. H. & Weiss, M. E. (1964). Tissue reactions in restorative
materials. Journal of Dental Research, 43, 807.
Shilling, G. (1977). The permanency of EBA cements. Journal of the American
Dental Association, 95, 187-9.
Silvey, R. G. & Myers, G. E. (1977). Clinical studies of dental cements. V.
Recall evaluation of restorations cemented with a zinc oxide-eugenol cement
and a zinc phosphate. Journal of Dental Research, 55, 289-91.
Silvey, R. G. & Myers, G. E. (1976). Clinical studies of dental cements. VI. A
study of zinc phosphate EBA-reinforced zinc oxide eugenol and polyacrylic
acid cements as luting agents in fixed prostheses. Journal of Dental Research,
56, 1215-18.
Skinner, E. W., Molnar, E. J. & Suarez, G. (1964). Reactions of zinc oxide with
carboxylic acids - physical properties. Journal of Dental Research, 43, 915.
Smith, D. C. (1958). The setting of zinc oxide/eugenol mixtures. British Dental
Journal, 105, 313-21.
Smith, D. C. (1960). A quick-setting zinc oxide/eugenol mixture. British Dental
Journal, 108, 232.
Smith, D. C. (1982a). Composition and characteristics of dental cements. In

356
References

Smith, D. C. & Williams, D. F. (eds.) Biocompatibility of Dental Materials.


Volume II. Biocompatibility of Preventive Dental Materials and Bonding
Agents, Chapter 8. Boca Raton: CRC Press Inc.
Smith, D. C. (1982b). Tissue reactions to cements. In Smith, D. C. & Williams,
D. F. (eds.) Biocompatibility of Dental Materials. Volume II. Biocompatibility
of Preventive Dental Materials and Bonding Agents, Chapter 10. Boca Raton:
CRC Press Inc.
Stansbury, J. W., Argentar, H. & Brauer, G. M. (1981). Cements from 2,5-
dimethyloxyphenol and zinc oxide. Journal of Dental Research, 60, 373.
Stansbury, J. W. & Brauer, G. M. (1985). Divanillates and polymerizable
vanillates as ingredients of dental cements. Journal of Biomedical Materials
Research, 19, 715-25.
Steinke, R., Newcomer, P., Komarneni, S. & Roy, R. (1988). Dental cements:
investigation of chemical bonding. Materials Research Bulletin, 23, 13-22.
Wallace, D. A. & Hansen, H. L. (1939). Zinc oxide eugenol cements. Journal of
the American Dental Association, 26, 1536-40.
Wessler, J. (1894). Pulpol, ein neues medicamentoses Cement. Deutsche
Monatsschrift fur Zahnheilkunde, 12, 478-84.
Williams, J. D., Swartz, M. L. & Phillips, R. W. (1965). Retention of
orthodontic bands as influenced by the cementing media. Angle Orthodontics,
4, 276-85.
Williams, P. D. & Smith, D. C. (1971). Measurement of the tensile strength of
dental restorative materials by use of a diametral compressive strength test.
Journal of Dental Research, 50, 436—42.
Wilson, A. D. (1975a). Dental cements - general. In von Fraunhofer, J. A. (ed.)
Scientific Aspects of Dental Materials, Chapter 4. London: Butterworths.
Wilson, A. D. (1975b). Zinc oxide dental cements. In von Fraunhofer, J. A. (ed.)
Scientific Aspects of Dental Materials, Chapter 5. London: Butterworths.
Wilson, A. D. (1976). Examination of the test for compressive strength applied
to zinc oxide eugenol cements. Journal of Dental Research, 55, 142-7.
Wilson, A. D. (1978). The chemistry of dental cements. Chemical Society
Reviews, 7, 265-96.
Wilson, A. D. & Batchelor, R. F. (1970). Zinc oxide eugenol cements. II. Study
of erosion & disintegration. Journal of Dental Research, 49, 593-8.
Wilson, A. D. & Batchelor, R. F. (1971). The consistency of dental cements. The
specification test for filling materials. British Dental Journal, 130, 437-41.
Wilson, A. D., Clinton, D. J. & Miller, R. P. (1973). Zinc oxide eugenol
cements. IV. Microstructure and hydrolysis. Journal of Dental Research, 52,
253-60.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986). An
evaluation of the significance of the impinging jet method for measuring the
acid erosion of dental cements. Biomaterials, 7, 55-60.
Wilson, A. D. & Lewis, B. G. (1980). The flow properties of dental cements.
Journal of Biomedical Materials Research, 14, 383-91.
Wilson, A. D., Prosser, H. J., Paddon, J. M. & Gilhooley, R. A. (1981). Calcium
hydroxide dimer (Cal-mer) cements. British Dental Journal, 150, 351-3.

357
Non-aqueous cements

Wilson, A. D. & Mesley, R. J. (1972). Zinc oxide eugenol cements. III. Infra-red
spectroscopic studies. Journal of Dental Research, 51, 1581-8.
Wilson, A. D. & Mesley, R. J. (1974). Chemical nature of cementing matrices of
cements formed from zinc oxide and 2-ethoxybenzoic acid-eugenol liquids.
Journal of Dental Research, 53, 146.
Zander, H. A. (1939). Reaction of the pulp to calcium hydroxide. Journal of
Dental Research, 18, 373-9.

358
10 Experimental techniques for the
study of acid-base cements

10.1 Introduction
The chief problem in studying the chemical nature of AB cements is that
many are essentially amorphous, so that the powerful tool of X-ray
diffraction (XRD) analysis cannot be used. Some AB cements do exhibit a
degree of crystallinity, but rarely in significant amounts; indeed, complete
crystallinity is usually a sign that the reaction product is not cementitious.
The literature contains numerous examples of workers being misled by
the results of XRD analysis into neglecting the presence and significance of
the amorphous phase.
A number of techniques have been employed that are capable of giving
information about amorphous phases. These include infrared spectro-
scopy, especially the use of the attenuated total reflection (ATR) or Fourier
transform (FT) techniques. They also include electron probe micro-
analysis, scanning electron microscopy, and nuclear magnetic resonance
(NMR) spectroscopy. Nor are wet chemical methods to be neglected for
they, too, form part of the armoury of methods that have been used to
elucidate the chemistry and microstructure of these materials.
In addition to spectrosopic studies of the setting chemistry of AB
cements, numerous mechanical tests have been used to measure properties
of the set materials. This latter group has included determination of
compressive and flexural strengths, translucency, electrical conductivity
and permittivity. The present chapter describes each of these techniques in
outline, and shows how they have been applied. Results obtained using
these techniques are described in earlier chapters which deal more
thoroughly with each individual type of AB cement.

359
Experimental techniques

10.2 Chemical methods


Broadly speaking, two groups of purely chemical methods have been
employed in the study of AB cements. These have been (1) studies of
cement formation and (2) degradative studies on set cements. The first
group has generally used the simple approach of attempting to react the
candidate acid, usually as an aqueous solution, with the candidate base,
which is generally a powder that is only sparingly soluble in water. The
success or otherwise of the attempted cementition has then been assessed
by a simple criterion, such as stability of the product in water.

10.2.1 Studies of cement formation


A number of studies have been carried out using the criterion of water
stability to assess the success of cementation. For example, Hodd &
Reader (1976) studied a range of metal oxide-poly acid cements by this
technique in order to determine the influence of the metal cation and the
polymer structure on stability. A number of polymeric carboxylic acids
were used in this study, namely poly(acrylic acid), poly(ethylene-maleic
acid), poly(methacrylic acid) and poly(vinyl methyl ether-maleic acid).
Poly(ethylene sulphonic acid) was also used, but it proved to be a poor
cement former and did not form stable cements with many of the cations
examined. A large number of metal oxides of both main group and
transition metals were examined, and a number of them were found to
form water-stable cements with most of the poly carboxylic acids. In rare
cases, metal oxides formed stable cements with a few of the acids, but
yielded only unstable mixtures with others. For example, bismuth oxide,
Bi2O3, gave a stable cement with poly(acrylic acid) that showed no
disintegration after 16 hours immersion in water at room temperature, but
with ethylene-maleic acid gave a mixture which disintegrated completely.
More common, however, was the behaviour of zinc oxide, which gave
stable cements with all of the polycarboxylic acids.
An equally simple chemical study was carried out on phytic acid-
aluminosilicate cements (Prosser et al, 1983). Phytic acid, myo-inositol
hexakis(dihydrogen phosphate), is a naturally occurring substance found
in seeds, and it is a stronger acid than phosphoric acid. Cements were
prepared using aqueous solutions of phytic acid, concentrated to 50 wt%,
and with 5 wt% zinc dissolved in the acid to moderate the rate of reaction
with the glass powder. Discs of cement were prepared and these were

360
Infrared spectroscopic analysis
placed in distilled water exactly seven minutes after the start of mixing of
the aqueous acid with powder. Soluble material was determined gravi-
metrically by evaporating the eluates to dryness. This approach demon-
strated that these cements had excellent resistance to early attack by water.
Other properties of these cements were promising for dental application
and a patent protecting this use has been sought (Lion Corporation, 1980).

10.2.2 Degradative studies


Degradative methods have been employed on cements that have been
allowed to set. In a typical degradative study, Cook (1983) treated
glass-ionomer cements with 3-3-molar potassium hydroxide solution. The
resulting solution was analysed for release of ions using atomic absorption
spectroscopy. This overall degradation technique allowed the measure-
ment of time-dependent concentrations of Al3+, Ca2+ and Na+ ions which
had entered the matrix from the glass. Cook concluded that both
aluminium and calcium ions were involved in the initial setting reaction of
these cements, although this has been disputed by other workers
(Nicholson et ai, 1988b; Wilson & McLean, 1988). Cook also concluded
that aluminium ions were the least easily extracted of all the cations
removed from the glass particles, which is not surprising given that the
aluminium is present in the glass initially as part of the aluminosilicate
network structure, and is anionic in character (Hill & Wilson, 1988).
Extraction studies have also been carried out by grinding the ageing
cements and extracting the soluble ions with water (Wilson & Kent, 1970;
Crisp & Wilson, 1974). Ion content was determined using atomic
absorption spectroscopy. The experiments give different, but complemen-
tary, results to those of Cook (1983), since what is extracted are those ions
that have been released from the glass powder but not yet insolubilized by
reaction with the polyacid.

10.3 Infrared spectroscopic analysis


10.3.1 Basic principles
The infrared region of the electromagnetic spectrum lies between the
wavelengths 1000 and 15000 nm (Kemp & Vellaccio, 1980). Absorption of
radiation in this region by organic compounds has been known since 1866,
when Tyndall first conducted experiments on the interaction of radiation
with compounds such as chloroform, methyl and ethyl iodides, benzene,

361
Experimental techniques
ethyl acetate and so on (Tyndall, 1866). More systematic work in the 30 or
so years from 1881 by a number of workers, most notably W. Coblentz,
showed that the interaction was attributable to individual groups within
the molecule (Cesaro & Torracca, 1988). From this deduction, and the
subsequent painstaking acquisition of empirical data, the modern ap-
proach to infrared spectroscopy has emerged, in which individual
functional groups can readily be identified and structural conclusions can
be drawn on the basis of a rapidly prepared infrared absorption spectrum.
In the most usual modern form of infrared spectroscopy, absorption is
plotted against reciprocal wavelength or wavenumber in cm"1 (Williams &
Fleming, 1973). The usual range of such a spectrum is 4000 cm"1 at the
high-frequency end to 625 cm"1 at the low-frequency end. Functional
groups in organic molecules absorb infrared radiation at well-defined parts
of this spectrum. Although this actually occurs due to absorption by the
whole of a complex organic molecule, such an absorption can be considered
to a reasonable approximation, for a number of spectral bands, to be
localized at individual functional groups. These localized absorptions give
rise to vibrations of the particular chemical bonds, which may include
stretching, bending, rocking, twisting or wagging (Williams & Fleming,
1973).
The specific requirement for a vibration to give rise to an absorption in
the infrared spectrum is that there should be a change in the dipole moment
as that vibration occurs. In practice, this means that vibrations which are
not centrosymmetric are the ones of interest, and since the symmetry
properties of a molecule in the solid state may be different from those of the
same molecule in solution, the presence of bands may depend on the
physical state of the specimen. This may be an important phenomenon in
applying infrared spectroscopy to the study of AB cements.

10.3.2 Applications to AB cements


It is apparent from the preceding discussion that the kinds of molecules
generally studied by infrared spectroscopy are organic. This means that the
AB cements which have been particularly studied by this technique are
those containing organic functional groups, most typically carboxylic acid
and carboxylate groups. In particular, extensive studies have been carried
out on the setting reactions and final structures of glass-ionomer cements
(Wilson & McLean, 1988), zinc polycarboxylates (Wilson, 1982) and other
metal oxide-poly(acrylic acid) cements (Crisp, Prosser & Wilson, 1976).

362
Infrared spectroscopic analysis

The range of acids for which infrared spectroscopy can be used is


limited. Despite this, the amount of detailed information which can be
obtained using the technique is very high. This is because of the extent and
comprehensiveness of the data available concerning the effect of subtle
changes in the bonding of carboxylate groups on the position of the
corresponding infrared absorption bands. Many detailed structures of
simple metal carboxylates have been established by complementary
techniques, such as X-ray crystallography, and these findings have been
used to establish correlations between band position in the infrared
spectrum and structure (Mehrotra & Bohra, 1983). There are limits to the
extent to which band position and structure can be correlated, however,
since there are exceptions to many of the apparently established empirical
rules. Moreover, some literature data have been shown to be plainly wrong
since they have ignored, for example, possible spectral changes caused by
cation and halide exchange due to mounting the sample between alkali-
metal halide discs (Deacon & Phillips, 1982). Nonetheless, some valuable
conclusions have been drawn using general correlations of carboxylate
band position with the nature of the bonding in set or setting cements.
Broadly speaking, in the study of AB cements derived from poly-
carboxylic acids, the band of interest falls in the region 1550-1620 cm"1
(Mehrota & Bohra, 1983; Bellamy, 1975). This band is the asymmetric
stretch of the carboxylate group and its exact position depends on both the
nature of the bonding involved (i.e. whether purely ionic or partially
covalent), and the nature of any chelation by the carboxylate group
(Bellamy, 1975).
The four possible modes of carboxylate bonding which have been
identified are purely ionic, unidentate, bridging bidentate and chelating
bidentate, as illustrated in Figure 5.3.
The ionic band falls at 1570-1575 cm"1, as it does in sodium and lithium
poly(acrylates) (Nicholson & Wilson, 1987) as well as in the simple
monomeric carboxylates (Mehrotra & Bohra, 1983). The unidentate
covalent binding gives a band close to 1550 cm"1, as does the chelating
bidentate, while the bridging bidentate generally gives a band in the region
1600-1620 cm"1. The unidentate mode of binding has been found to be
relatively rare in monomeric carboxylate compounds (Mehrotra & Bohra,
1983), and on these grounds was rejected as a probable structure in
polycarboxylate materials by Nicholson, Wasson & Wilson (1988).

363
Experimental techniques

10.3.3 Fourier transform infrared spectroscopy


In recent years, infrared spectroscopy has been enhanced by the possibility
of applying Fourier transform techniques to it. This improved spectro-
scopic technique, known as Fourier transform infrared spectroscopy
(FTIR), is of much greater sensitivity than conventional dispersive IR
spectroscopy (Skoog & West, 1980). Moreover, use of the Fourier
transform technique enables spectra to be recorded extremely rapidly, with
scan times of only 02 s. Thus it is possible to record spectra of AB cements
as they set. By comparison, conventional dispersive IR spectroscopy
requires long scan times for each spectrum, and hence is essentially
restricted to examining fully-set cements.
FTIR has been applied to both zinc polycarboxylate cements (Nicholson
et al., 1988a) and glass-ionomer cements (Nicholson et al, 1988b), in both
cases yielding significantfindings.The zinc polycarboxylate was shown for
the first time to become partially covalent with time after setting, while the
role of (+ )-tartaric acid in glass-ionomer cements was shown to be to
suppress early formation of calcium polyacrylate acid and to enhance later
formation of aluminium polyacrylate. These results are discussed in more
detail in Chapter 5.

10.4 Nuclear magnetic resonance spectroscopy


10.4.1 Basic principles
The NMR spectrum can be recorded for compounds containing those
elements whose nuclei have spin values of \ (Williams & Fleming, 1973).
A large number of such nuclei exist, including 1H, 13C, 19F, 27A1 and 31P.
Unfortunately, many other nuclei of importance in chemistry, such as 12C
and 16O, have nuclear spin values of 0 and hence do not give nuclear
resonance signals in a magnetic field.
To study NMR spectra of compounds, apparatus is required that
consists of three sets of components. These are a radio-frequency
transmitter, a homogeneous magneticfieldand a radio-frequency receiver.
In addition to these, the apparatus includes a unit to sweep the magnetic
field over a small range, a mere few parts per million.
The earliest NMR technique to gain importance in chemistry was that of
proton NMR. Spectra could be obtained for compounds containing the *H
nucleus by continuously sweeping the field at constant frequency. This

364
Nuclear magnetic resonance spectroscopy

approach results in so-called continuous wave spectra, which suffer from


the general disadvantage that only a small portion of the spectrum is
excited at any one time. Such spectra have unduly low signal-to-noise
ratios, which is particularly undesirable when studying nuclei of low
abundance and/or low sensitivity, such as 13C. As a result continuous wave
is not used for 13C NMR spectra but instead pulsing techniques are used
together with Fourier transformation of the data thus obtained.
For 13C NMR, the radio frequency is applied as a short, powerful pulse
which acts like a spread of frequencies. All the NMR-active nuclei are
excited by this pulse and then decay back to their equilibrium states. These
decays result in a series of complex sine waves which diminish exponentially
with time. By the application of Fourier transform techniques, such decay
curves can be converted into spectra. By making use of the capability of
storing many such pulsed spectra on a computer and adding the signals
together prior to applying the Fourier transform, a significant improve-
ment in the signal-to-noise ratio can be obtained.
The nuclei studied by NMR spectroscopy are affected by the precise
nature of their electronic environments. This means, for example, that
protons will resonate at different frequencies according to their position in
a molecule. In this way, it is possible to distinguish between protons in
methyl groups and methylene groups, in hydroxyl groups or as part of the
benzene ring. In addition, protons interact with each other, leading to what
is known as spin-spin coupling and giving rise to well-characterized
splitting patterns in their NMR spectrum. All of these features may be used
to give structural information about molecules containing XH nuclei.
Carbon-13 nuclei, due to their low natural abundance, do not interact
with each other in a molecule, though they are affected by adjacent
protons. In practice, such couplings are removed by irradiation of the
whole spectrum as it is recorded, in a technique known as proton noise
decoupling. This means that practical 13C NMR spectra exhibit one unsplit
signal for each type of carbon atom present in the sample.

10.4.2 Applications to AB cements


NMR spectroscopy of various nuclei has been used in the study of AB
cements derived from various acids, including phosphoric acid and
poly(acrylic acid). For example, 31P NMR has been used in studies of
dental silicate cement, i.e. the AB cement made from aqueous phosphoric
acid and powdered aluminosilicate glass (Wilson, 1978). In this cement, the

365
Experimental techniques
setting reaction is controlled by dissolving small amounts of aluminium
and/or zinc metal in the phosphoric acid. Using 31P NMR, O'Neill et al.
(1982) were able to distinguish between the various complexes formed by
ions of these metals in the presence of phosphoric acid.
The chemistry of polyelectrolyte cement liquids has been studied using
13
C NMR. Watts (1979) used this technique to distinguish between the
homopolymer of acrylic acid and its copolymer with itaconic acid in
various commercial polyelectrolyte dental cements. This was readily
achieved because of the ability of 13C NMR to differentiate between
carbon atoms in chemical environments that are only slightly different.
Prosser, Richards & Wilson (1982) used 13C NMR spectroscopy to study
the role of ( + )-tartaric acid in modifying the setting behaviour of
glass-ionomer dental cements. For this, a model system was used, with a
lower ratio of glass powder to poly (acrylic acid) liquid than in conventional
cements; this slowed the reaction, enabling the spectra to be recorded on
acceptable time scales. This study showed for the first time that ( + )-
tartaric acid was an effective additive for controlling the setting character-
istics of these cements because it reacts preferentially with Ca2+ ions
released from the glass. Hence, (+ )-tartaric acid acts to extend the working
time of these cements. The work also showed that there was no difference
between the reactivity of the acrylic acid and the itaconic acid segments
when the copolymer was used as the acidic component.

10.5 Electrical methods


Changes in electrical conductivity have occasionally been used to study the
setting chemistry of AB cements. Conductivity has been particularly used
in the study of dental cements, notably the dental silicate (Wilson & Kent,
1968), the zinc polycarboxylate (Cook, 1982), the glass-ionomer cement
(Cook, 1982) and the ZOE cement (Crisp, Ambersley & Wilson, 1980).
In a typical study of conductivity, Cook (1982) used a cell consisting of
two platinum disc electrodes, 12 mm in diameter and 1-5 mm apart. The
setting AB cement was examined in this cell which had been calibrated
using a standard solution of 0-02 M potassium chloride. Plots were recorded
of specific conductance against time for each of the setting cements. For
zinc polycarboxylate there was found to be a rapid drop in specific
conductance about 10 minutes after the start of mixing. This behaviour
was consistent with the replacement of relatively mobile protons by
significantly less mobile zinc ions in the polycarboxylate chain. Con-

366
X-ray diffraction

ductivity was found to drop rapidly by a factor of 200 which was held to
be consistent with quantitative neutralization of the poly(acrylic acid) and
with a very low diffusion coefficient for zinc ions in the set cement. By
contrast, the specific conductance of glass-ionomer cements was found to
decrease much more gradually, indicating that the setting reaction in these
cements is much slower than for the zinc polycarboxylates. Moreover,
setting was still not complete even 1000 minutes after the start of mixing.
An alternative electrical method that has been used in the study of
glass-ionomer cements has been the measurement of dielectric properties.
Tay & Braden (1981, 1984) measured the resistance and capacitance of
setting cements at various times from mixing. From the results obtained,
relative permittivity and resistivity were calculated. In general, as these
cements set, their resistivity was found to fall rapidly, then to rise again.
Both these results and the results of relative permittivity measurements
were consistent with the cements comprising highly ionic and polar
structures.

10.6 X-ray diffraction


10.6.1 Basic principles
X-ray diffraction is the most accurate and powerful method of both
identifying solids and determining their structure. This is essentially
because the regular array of atoms or ions in a crystalline solid is spaced at
dimensions corresponding to the wavelength of X-rays, and such arrays
are consequently able to act as an X-ray diffraction grating. The resulting
diffraction pattern is recorded either on photographic film or by an
electronic detector.
The underlying principle of X-ray diffraction is as follows. When a beam
of X-rays passes through a crystalline solid it meet various sets of parallel
planes of atoms. The diffracted beams cancel out unless they happen to be
in phase, the condition for which is described in the Bragg relationship:

nk = 2dsin9

where X is the wavelength of the X-rays, d is the distance between the


planes, and 6 is the angle of incidence of the X-rays on the planes.
In practice, values of 6 can be measured and, since the wavelength of the
X-rays is known in any given experiment, values of d can be calculated.
These lvalues are related to the unit cell symmetry and dimensions. If the

367
Experimental techniques
intensities of the diffracted beam in each direction are also measured, the
complete structure of the solid can be determined (Mackay & Mackay,
1972).
Each atom in the lattice acts as a scattering centre, which means that the
total intensity of the diffracted beam in a given direction depends on the
extent to which contributions from individual atoms are in phase. Relating
the underlying structure to the observed diffraction pattern is not
straightforward, but is essentially a trial-and-error search involving
extensive computer-based calculations.
For materials which are available not in the form of substantial
individual crystals but as powders, the technique pioneered by Debye and
Scherrer is employed (Moore, 1972). The powder is placed into a thin-
walled glass capillary or deposited as a thin film, and the sample is placed
in the X-ray beam. Within the powder there are a very large number of
small crystals of the substance under examination, and therefore all
possible crystal orientations occur at random. Hence for each value of d
some of the crystallites are correctly oriented to fulfil the Bragg condition.
The reflections are recorded as lines by means of a film or detector from
their positions, the lvalues are obtained (Mackay & Mackay, 1972).

10.6.2 Applications to AB cements


X-ray diffraction has been applied to certain AB cements. For example,
Crisp et al. (1979), in a study of silicate mineral-poly(acrylic acid) cements,
used the technique both to assess the purity of the powdered minerals
employed and to monitor mineral decomposition in mixtures with
poly(acrylic acid), in order to indicate whether or not cement formation
had taken place. They employed Cu Ka radiation passed through a nickel
filter; for most of the samples, a seven-hour exposure time was found to be
adequate for the development of a discernible diffraction pattern. Samples
were identified by reference to published powder diffraction data.
A number of other studies of AB cements have used X-ray diffraction.
For example, Sorrell (1977) and Sorrell & Armstrong (1976) employed the
technique in the study of oxychloride cements formed in aqueous solution
by interaction of oxides and chlorides of either zinc or magnesium.
Individual phases were identified, again using Cu Ka radiation, this time
comparing results with those previously obtained for pure compounds.
Results from these two studies are described in detail in Sections 7.2 and
7.3 respectively.

368
Electron probe microanalysis

10.7 Electron probe microanalysis


10.7.1 Basic principles
This technique can be applied to samples prepared for study by scanning
electron microscopy (SEM). When subject to impact by electrons, atoms
emit characteristic X-ray line spectra, which are almost completely
independent of the physical or chemical state of the specimen (Reed, 1973).
To analyse samples, they are prepared as required for SEM, that is they are
mounted on an appropriate holder, sputter coated to provide an electrically
conductive surface, generally using gold, and then examined under high
vacuum. The electron beam is focussed to impinge upon a selected spot on
the surface of the specimen and the resulting X-ray spectrum is analysed.
Analysis is most frequently done qualitatively since there are problems
in quantification (Reed, 1973). Although intensity is approximately
proportional to mass concentration of a given element there are significant
deviations, depending on which other elements are present.
There is also a minimum atomic number which can usually be detected,
since the practical maximum X-ray wavelength that is used is fixed at
0-12 nm. This in turn fixes sodium (atomic number 11) as the lightest
element for which this technique is valid. Special techniques are available
to overcome this limitation, but they are not in general use, and have not
been applied to AB cements.

10.7.2 Applications to dental silicate cements


In a study of dental silicate cements, Kent, Fletcher & Wilson (1970) used
electron probe analysis to study the fully set material. Their method of
sample preparation varied slightly from the general one described above,
in that they embedded their set cement in epoxy resin, polished the surface
toflatness,and then coated it with a 2-nm carbon layer to provide electrical
conductivity. They analysed the various areas of the cement for calcium,
silicon, aluminium and phosphorus, and found that the cement comprised
a matrix containing phosphorus, aluminium and calcium, but not silicon.
The aluminosilicate glass was assumed to develop into a gel which was
relatively depleted in calcium.

10.7.3 Applications to glass-ionomer cements


A similar study was carried out on glass-ionomer cements (Barry, Clinton
& Wilson, 1979), which showed some interesting similarities to the dental

369
Experimental techniques
silicate cements, as well as some differences. Firstly, aluminium and
calcium were found to be removed from the glass particles and to reside in
the continuous phase. However, by contrast with the dental silicate
cement, some silicon was also found in the continuous phase. Attack by the
poly(acrylic acid) had apparently occurred preferentially at the calcium-
rich sites of the glass, a finding that was significant in formulating the
theory of the setting chemistry of these materials.

10.8 Measurement of mechanical properties


Strength has been widely measured for AB cements. It is formally defined
as the force experienced by a material at the point where fracture occurs
(Gillam, 1969). The study of strength is complicated in that fracture is a
point of discontinuity, so cannot be readily interpreted in terms of events
leading up to it.
Strength can be measured in compression, in tension, in shear and
transversely (flexural strength). However, if we exclude plastic flow as a
means of failure, then materials can only fracture in one of two ways: (1)
by the pulling apart of planes of atoms, i.e. tensile failure, or (2) by the
slippage of planes of atoms, i.e. shear failure. Strength is essentially a
measure of fracture stress, which is the point of catastrophic and
uncontrolled failure because the initiation of a crack takes place at
excessive stress values.
AB cements tend to be essentially brittle materials. This means that
when subjected to mechanical loading, they tend to rupture suddenly with
minimal deformation. There are a number of different types of strength
which have been identified and have been determined for AB cements.
These include compressive, tensile and flexural strengths. Which one is
determined depends on the direction in which the fracturing force is
applied. For full characterization, it is necessary to evaluate all of these
parameters for a given material; no one of them can be regarded as the sole
criterion of strength.
Generally, strength is determined by applying forces uniaxially using an
apparatus consisting of a pair of jaws which move either together or apart
in a controlled manner. A chart recorder is employed to give a permanent
record of results obtained, so that the force at fracture can be determined.
Whether such an apparatus measures tensile, compressive or flexural
strength depends on how the sample is oriented between the jaws and on
the direction that the jaws are set to travel relative to each other. Such tests

370
Measurement of mechanical properties
need to be repeated on several samples in order to provide sufficient data
for statistical analysis and to allow calculation of both the mean and the
scatter of the results.

10.8.1 Compressive strength


The most common mechanical property of cements that has been measured
routinely is compressive strength (Polakowski & Ripling, 1966). Measure-
ment is easy to carry out but there are several reasons to consider that the
results from the technique are unsatisfactory. Interpretation of results is
uncertain because of the complexities in the mode of failure. Minor
imperfections in the material lead to localized stress concentrations which
affect the magnitude of the result.
Failure is complex because both the mode and plane of failure are
variable. Failure under compressive load can occur by plastic yielding,
cone failure (secondary shear forces) or axial splitting (secondary tensile
forces) (Kendall, 1978). The mode of failure depends on the size and
geometry of the specimen, the nature of the material tested and the rate of
loading. The studies of Selenrath & Gramberg (1958) are of interest in
showing the effect of the nature of the material on fracture. They found
that when cylindrical specimens were placed unrestrained in the testing
machine, glass and lithographic limestone exhibited simple vertical fracture
or axial splitting, evidence of tensile failure. By contrast, coarse-grained
marble and bothfine-grainedand coarse-grained sandstone yielded cones
at the ends of the specimen with typical shearing fracture planes. However,
when the ends of the lithographic limestone were clamped they exhibited
both types of fracture.
The variation in the mode of failure makes comparison of different types
of cement quite impossible. As Darvell (1990) has pointed out, compressive
strength is not a material property under any condition, but can only be
used to compare materials of a very similar nature.
The compressive strength of AB cements used in dentistry has been
widely studied (Wilson & McLean, 1988). It is the method, for example,
specified in the British Standard on dental cements. However, there is
concern that the result is less clinically relevant than the evaluation of
flexural strength. Moreover, the latter is more discriminating (Prosser et
al., 1984). Despite this, compressive strength has been used to indicate
clinical acceptability; phosphate-bonded cements with low compressive
strength tend to be unsatisfactory in other respects such as durability, and

371
Experimental techniques
hence there is value in using compressive strength as a criterion of general
material quality (Wilson & McLean, 1988).

10.8.2 Diametral compressive strength


The diametral compressive strength has been used to estimate the tensile
strength of certain AB cements (Smith, 1968). In this test, the load is
applied diametrically across a cylinder of cement. Theoretical consider-
ation of the test geometry shows that for a perfectly brittle material the
failure that occurs is tensile in character. The difficulty in applying this test
to AB cements is that they are not sufficiently brittle for this to hold true.
In particular, the zinc polycarboxylate and glass-ionomer cements show
sufficient plastic character to make the relationship between diametral
compressive and tensile strength vary between AB cements of different
types; like the compressive strength test, this test is valid only as a means
of comparison between similar materials (Darvell, 1990).
For glass-ionomer cements, there have been several studies on the
factors affecting strength. For example Crisp, Lewis & Wilson (1977)
showed that both compressive and tensile strengths increased linearly with
concentration of polyacid in the liquid component, though neither
extrapolated to zero at zero acid concentration. Ageing was also shown to
influence compressive strength of these materials, older cements being
stronger than younger ones. However, the exact development of strength
was found to depend on storage conditions (Crisp, Lewis & Wilson, 1976).

10.8.3 Flexural strength


Flexural strength is determined using beam-shaped specimens that are
supported longways between two rollers. The load is then applied by either
one or two rollers. These variants are called the three-point bend test and
the four-point bend test, respectively. The stresses set up in the beam are
complex and include compressive, shear and tensile forces. However, at the
convex surface of the beam, where maximum tension exists, the material is
in a state of pure tension (Berenbaum & Brodie, 1959). The disadvantage
of the method appears to be one of sensitivity to the condition of the
surface, which is not surprising since the maximum tensile forces occur in
the convex surface layer.
Of all the methods of determining strength, theflexuraltest appears to
be the most satisfactory. While not ideal, it does have the advantage of

372
Measurement of mechanical properties
measuring a clearly defined parameter. However, very few studies of AB
cements involving this test have been carried out. Prosser, Powis & Wilson
(1986) studied the influence of various formulation changes on the flexural
strength of glass-ionomer cements. They found that the flexural strength
of glass-ionomer cements was dependent on the glass and the polyacid
used to prepare them. Opaque and opal glasses containing crystallites were
found to yield cements of higherflexuralstrength than those prepared from
clear glasses; increasing the relative molar mass of the polyacid was also
found to improve flexural strength.

10.8.4 Fracture toughness


The use of fracture stress as a measure of resistance to fracture is suspect.
In all these tests, whether compressive, tensile or flexural, failure is
catastrophic because there is no suitable flaw for crack propagation. A
high force is needed to start a crack and as a result the subsequent
propagation takes place under too great a stress. Then there is the curious
finding that compressive strength values are 10 times those for tensile
strength although, in principle, both are measures of cohesion.
For these reasons, attention has been paid to energy criteria as a measure
of toughness. The following points need to be noted. Materials do not
reach their theoretical strength (that is of their primary chemical bonds),
because of the presence of minute flaws. Stress is concentrated at these
flaws and so is enhanced. In effect this amounts to a weakening of the
material. Under load, cracks propagate from these flaws and lead to
failure.
Propagation of cracks requires energy to create new surfaces but also
releases stored energy. Unstable propagation of cracks occurs when the
strain energy released exceeds that required to create new surfaces, and
occurs when the crack reaches a certain length. This is because strain
energy released is proportional to (crack length)2, and the energy to create
a new surface is proportional to the crack length.
Fracture toughness is the resistance to propagation of cracks through a
material and is usually quantified by the stress intensity factor, Kx, defined
as
Kx = GF(na)*

where a is the flaw size and aF the fracture stress.


There are a number of methods of determining fracture toughness, but
the one used so far for AB cements is the double torsion method introduced

373
Experimen tal techn iques
by Outwater et al. (1974), which appears to have some advantages over
other methods.
Double torsion test specimens take the form of rectangular plates with
a sharp groove cut down the centre to eliminate crack shape corrections.
An initiating notch is cut into one end of each specimen (Hill & Wilson,
1988) and the specimen is then tested on two parallel rollers. A load is
applied at a constant rate across the slot by two small balls. In essence the
test piece is subjected to a four-point bend test and the crack is propagated
along the groove. The crack front is found to be curved.
The double torsion test specimen has many advantages over other
fracture toughness specimen geometries. Since it is a linear compliance test
piece, the crack length is not required in the calculation. The crack
propagates at constant velocity which is determined by the crosshead
displacement rate. Several readings of the critical load required for crack
propagation can be made on each specimen.
When the load has reached a critical plateau value, the crack continues
to propagate at constant load. Crack propagation can be stopped by
removing the load, with the implication that several readings can be made
on one test specimen. Crack velocity is determined by the crosshead speed,
modulus of the material and specimen dimensions.
To calculate fracture toughness using the double torsion test piece, the
following equation is used:

where Kx = fracture toughness, Pc = critical load for crack propagation,


Wm = specimen width, t = specimen thickness, tc = crack depth, Wc =
distance between the supports and v = Poisson's ratio.

10.9 Setting and rheological properties


The rheological properties of AB cements are important where those
cements have been used in dentistry. It is not sufficient simply to have a
cement which eventually sets to give a resistant, strong material. The
cement must also remainfluidfor a sufficient time to allow placement, and
ideally must develop a good degree of hardness very rapidly following
placement. Thus the setting chemistry must be such that the reaction which
occurs immediately the acid and the base are mixed does not lead too
rapidly to the development of a viscous paste that is difficult to manipulate.
Ideally, viscosity should be low enough to allow manipulation but high
enough so that the fluid cement does not flow appreciably once in place.

374
Setting and rheological properties
Thus, the rheological requirements of materials for this application are
demanding, and their evaluation has been important in the study of this
group of AB cements.
The rheological characteristics of AB cements are complex. Mostly, the
unset cement paste behaves as a plastic or plastoelastic body, rather than
as a Newtonian or viscoelastic substance. In other words, it does not flow
unless the applied stress exceeds a certain value known as the yield point.
Below the yield point a plastoelastic body behaves as an elastic solid and
above the yield point it behaves as a viscoelastic one (Andrade, 1947). This
makes a mathematical treatment complicated, and although the theories of
viscoelasticity are well developed, as are those of an ideal plastic (Bingham
body), plastoelasticity has received much less attention. In many AB
cements, yield stress appears to be more important than viscosity in
determining the stiffness of a paste.

10.9.1 Problems of measurement


Consistency, working time, setting time and hardening of an AB cement
can be assessed only imperfectly in the laboratory. These properties are
important to the clinician but are very difficult to define in terms of
laboratory tests. The consistency or workability of a cement paste relates
to internal forces of cohesion, represented by the yield stress, rather than
to viscosity, since cements behave as plastic bodies and not as Newtonian
liquids. The optimum stiffness or consistency required of a cement paste
depends upon its application.
Three useful tests have been used to evaluate the working and setting
properties of experimental cements. These are the parallel plate plas-
tometer, the penetrometer and the oscillating rheometer. They are
described in the following sections of this chapter.

10.9.2 Methods of measurement


Initially, the test that was used to determine setting time was one based on
response of the newly mixed AB cements to the application of a weighted
needle of known dimensions. This test was originally devised by Gillmore
(1864) for his studies on the setting of hydraulic cements and mortars, and
the weighted device, known as a Gillmore needle, had a mass of 454 g (1 lb)
and a tip diameter of 105 mm (Crisp, Merson & Wilson, 1980).
The drawback with the use of the Gillmore needle is that it is not a test
of any well-defined rheological property, but of resistance to indentation.

375
Experimental techniques
This property does not necessarily correlate with the changes in rheological
characteristics undergone by a cement as it sets. A more satisfactory test,
developed in the early 1970s, is oscillating rheometry, first described by
Bovis, Harrington & Wilson (1971) and subsequently refined slightly by
Wilson, Crisp & Ferner (1976). A theoretical treatment of the results of
oscillating rheometry was provided by Cook & Brockhurst (1980).
The apparatus used for oscillating rheometry consists of a pair of
grooved metal plates clamped together, between which the sample of
cement is placed. The bottom plate is connected via a spring to a motor,
which causes reciprocating motion at the end of the spring furthest from
the plate. Initially, motion of the motor causes corresponding motion in
the bottom plate. As the cement sets, and the force required to move the
plate increases, so the motion in the lower plate diminishes. A trace is
produced on a chart recorder which shows the changes in oscillation with
time as the cement sets, beginning from the large initial amplitude and
declining to negligible, or in some cases, zero oscillation when the cement
is fully set.
The original definition of working time using this apparatus was the time
taken from the start of mixing to reach an oscillation 95 % of the original
value (Bovis, Harrington & Wilson, 1971). An alternative method of
estimating working time was suggested by Wilson, Crisp & Ferner (1976),
which consisted of drawing lines as extensions to the initial straight portion
of the rheogram and extending the tangent at the maximum setting rate
back to cross these lines. The point of intersection is then taken as the
working time. In practice very little difference is observed in the working
times from oscillating rheometry obtained by either of these construction
methods.
Since its development as a technique, oscillating rheometry has been
widely applied to the study of AB cements for use in dentistry, most
notably to the glass-ionomer cement. For example, it was used to study the
effect of chelating comonomers on the setting of glass-ionomers (Wilson,
Crisp & Ferner, 1976). This study examined a range of compounds
including hydroxybenzoic acids, diketones, and most significantly,
hydroxyacids. The technique was useful in identifying the particular
advantages which are associated with the use of 5% ( + )-tartaric acid in
glass-ionomer cements, namely delayed onset of gelation and eventual
increased rate of set. These improvements in handling properties have been
crucial in making glass-ionomer cements fully acceptable for use in clinical
dentistry (Wilson & McLean, 1988).

376
Setting and Theological properties

In another study, oscillating rheometry was used to examine the effect of


adding various simple metal salts to glass-ionomer cements (Crisp, Merson
& Wilson, 1980). It was found that cement formation for certain glasses
which react only slowly with poly(acrylic acid) could be accelerated
significantly by certain metal salts, mainly fluorides such as stannous
fluoride and zinc fluoride. Some non-reactive glasses could be induced to
set by the addition of such compounds.
As a further example of the use of the technique of oscillating rheometry,
the work of Crisp et al. (1979) can be cited. This study was of the formation

Figure 10.1 Parallel-plate plastometer for determination of consistency.

377
Experimental techniques
of AB cements from poly(acrylic acid) and basic minerals. The minerals
examined were ortho- and pyro-silicates, which had been ground into fine
powders of sufficiently small particle size to pass through a 38-//m test
sieve. With a number of such minerals, including willemite, gehlenite and
hardystonite, oscillating rheometry demonstrated that there was a reason-
ably rapid setting reaction at 23 °C. Infrared spectroscopy was used to
confirm reaction in fully hardened cements and mechanical property
measurements were carried out, though it was concluded that the cements
were too weak and porous to be practically useful.
Oscillating rheometry continues to be useful in the study of AB cements,
and has recently been used to give further insight into the role of ( + )-
tartaric acid in glass-ionomer cements (Hill & Wilson, 1988). Further
examples of its use are described in earlier chapters of this book.
Consistency is tested on a measured volume of freshly mixed cement in
the form of a cylinder. This specimen is placed between two horizontal
plates using the apparatus illustrated in Figure 10.1 and subjected to a
vertically applied load. The cement then flows out rapidly to form a disc.
This radial flow ceases almost instantaneously because the applied stress
decreases as the disc expands and rapidly reaches the yield stress, at which
point outward flow ceases. This is the behaviour expected for a plastic
body.
The diameter of the disc is measured and this gives an indication of the
shear strength of the paste. It is not a measure of viscosity becauseflowhas
ceased at this point. The shear strength of the paste can be calculated from
the following formula, which was derived by Wilson & Batchelor (1971).
Shear strength = 4SPV/TZ2D5
where P = applied load, V = volume of the cement paste, and D =
diameter of the disc.
The consistency depends on the powder/liquid ratio used to mix the
cement, and the parallel plate plastometer can be used to determine the
optimum ratio for a particular cement system.

10.10 Erosion and leaching


10.10.1 Importance in dentistry
For the various AB cements used in clinical dentistry, erosion and/or
leaching of components have been considered important in assessing
durability (Wilson & McLean, 1988). In fact, the two aspects are not

378
Optical properties
necessarily related to durability. Loss of soluble species from the set cement
affects durability only if the species concerned are matrix-formers. If not,
such loss has no effect and indeed may be beneficial. Thefluoriderelease by
glass-ionomer cements is regarded as clinically advantageous, and
apparently takes place by ion exchange, there being no discernible loss of
material as the process occurs (Wilson & McLean, 1988).

10.10.2 Studies of erosion


Erosion is the result of both chemical attack and mechanical wear. In
dentistry the chemical attack comes from acids either generated in the
mouth by dental plaque or present in foods and beverages (Pluim &
Arends, 1987; Wilson & Batchelor, 1968). To mimic this attack in
laboratory testing, a static solubility test was originally carried out, which
employed appropriate solutions of eroding acids (Kent, Lewis & Wilson,
1973). More recently, the mechanical aspect has been introduced by using
a test in which jets of aqueous acid impinge on a sample of cement (Wilson
et al.,1986); this test gives results for erosion that agree with clinical studies
of durability (Setchell, Teo & Kuhn, 1985). In particular, using dilute lactic
acid as the eroding agent in an impinging jet test, it has been shown (Wilson
et al., 1986) that extent of erosion increased in the order
glass-ionomer < silicate < zinc phosphate < zinc polycarboxylate
Setchell, Teo & Kuhn (1985) observed that glass-ionomer cements
prepared from poly(acrylic acid) were more resistant to erosion than such
cements prepared from maleic acid copolymers. This has been confirmed
by Wilson et al. (1986) and by Billington (1986), even when, as in the latter
case, the same glass was used in both cements. The method has been
reviewed recently by Billington, Williams & Pearson (1992).

10.11 Optical properties


10.11.1 Importance in dentistry
For certain AB cements, used in dentistry, optical properties are important
for their overall acceptability as materials. The two particular properties of
interest have been colour and translucency, both of which need to match
natural tooth material as closely as possible if good aesthetics are to be
developed (Wilson & McLean, 1988). Of the AB cements currently used in
dentistry, the glass-ionomer cement has the best aesthetics, since it has a

379
Experimen tal techniques
degree of translucency. This translucency arises because thefilleris a glass,
and unlike zinc oxide used in the zinc polycarboxylate cements, is not
opaque.
Evaluation of these optical properties may be done by simple ob-
servation; this approach is useful clinically (Knibbs, Plant & Pearson,
1986), since acceptability of the colour match to the surrounding tooth
material can be readily seen without the need for instrumental measure-
ment. On the other hand, for quantitative evaluation of optical properties,
some kind of instrumental measurement is necessary, and the property
usually evaluated is opacity.

10.11.2 Measurement of opacity


The main technique that has been used for the measurement of opacity has
been to prepare a standard disc of AB cement 1-0 mm thick and aged for
24 hours at 37 °C. This disc, contained in a small trough of water to prevent
desiccation, is placed in a reflectometer on a black background. It is then
illuminated with diffuse light and the amount of light reflected from it, Ro,
is measured. The disc is then placed on a white background of 70%
reflectivity, and the new amount of reflected light, R07, measured. The
contrast ratio R0/R07 is defined as the C0.7 opacity (Crisp et al., 1979).
Using this technique, it has been shown that the opacity of glass-ionomer
cements decreases as they age; in other words, their translucency increases
over this time. This change has been found to be rapid in thefirsthour after
mixing, but becomes much slower after this time (Wilson & McLean,
1988).
Colour and opacity have been found to be connected for glass-ionomer
cements (Crisp et al., 1979; Asmussen, 1983), with darker shades giving
increased opacity. However, this is merely a consequence of the underlying
physical relationships, and is not thought to be a clinical problem (Wilson
& McLean, 1988), mainly because the stained tooth material for which the
darker shades are necessary for colour match is itself of reduced
translucency.

10.12 Temperature measurement


When AB cements harden there may be a considerably exothermic
reaction. For those cements used in dentistry, this may have clinical
significance, since excessive temperature rise can cause damage to the

380
Other test methods
dental pulp. In a survey of a wide range of AB dental cements, Crisp,
Jennings & Wilson (1978) made use of a fairly simple miniature
calorimeter. This consisted of a block of expanded polystyrene, into which
a hole 6 mm in diameter and 6 mm deep was drilled. This hole could be
covered with a lid also made of polystyrene. Cements were mixed and
placed in this calorimeter and the temperature rise on setting was
monitored using a sealed NiCr-NiAl thermocouple.
Of the cements examined, the zinc phosphate cement showed the
greatest exotherm on setting, a maximum temperature rise of 22-1 °C being
observed. By contrast, glass-ionomer cement gave the smallest exotherm,
only 3-9 °C. Dental silicate and silicophosphate cements also showed
significant exotherms, at about 8 °C; should such an amount of heat be
generated in clinical use, this would be considered potentially harmful to
dental pulp (Paffenbarger et al, 1949). Of the remaining cements, the zinc
oxide-eugenol and the zinc polycarboxylate gave smaller exotherms of
about 5 °C.
Difficulties have been found in relating laboratory measurements of
exotherm behaviour in these AB cements to what happens when they are
used in clinical practice. As a consequence there have been few other
studies of temperature rise in the setting of such materials.

10.13 Other test methods


This chapter has covered the more important test methods used in the
study of AB cements and their underlying principles, but it is not
exhaustive. A number of experimental methods have been used to assess
particular properties of certain cements for particular applications. For
example, for AB dental cements, and especially glass-ionomer cements,
adhesion to tooth material has been studied (Wilson & McLean, 1988).
However, this is not a property of general interest for AB cements, nor
have satisfactory standard methods been developed. Consequently,
adhesion to tooth material has not been covered in the present chapter, but
has been left to chapters covering cements for which it is appropriate.
Other properties of glass-ionomer cements, such as fluoride-ion release
and effect of relative molar mass of the poly(acrylic acid) on strength have
been studied, but using techniques that are of interest only in the context
of these particular cements. Once again, these tests are discussed in the
chapter devoted to the AB cements in question.

381
Experimental techniques

References
Andrade, E. N. da C. (1947). Introduction to Rheology. London: Oil & Colour
Chemists' Association.
Asmussen, E. (1983). Opacity of glass-ionomer cements. Acta Odontologica
Scandinavica, 41, 151-7.
Barry, T. I., Clinton, D. J. & Wilson, A. D. (1979). The structure of a
glass-ionomer cement and its relationship to the setting process. Journal of
Dental Research, 58, 1072-9.
Bellamy, L. J. (1975). The Infrared Spectra of Complex Molecules. London:
Chapman and Hall.
Berenbaum, R. & Brodie, I. (1959). The tensile strength of coal. Journal of the
Institute of Fuel, 32, 320-7.
Billington, R. W. (1986). Personal communication, cited in Wilson & McLean
(1988).
Billington, R. W., Williams, J. A. & Pearson, G. J. (1992). In vitro erosion of
20 commercial glass ionomer cements measured using the lactic acid jet test.
Biomaterials, 13, 343-7.
Bovis, S. C, Harrington, E. & Wilson, H. J. (1971). Setting characteristics of
composite filling materials. British Dental Journal, 131, 352-6.
Cesaro, S. N. & Torracca, E. (1988). Early applications of infrared spectroscopy
to chemistry. Ambix, 35, 39—46.
Cook, W. D. (1982). Dental polyelectrolyte cements. I. Chemistry of the early
stages of the setting reaction. Biomaterials, 3, 232-6.
Cook, W. D. (1983). Degradative analysis of glass-ionomer polyelectrolyte
cements. Journal of Biomedical Materials Research, 17, 1015-27.
Cook, W. D. & Brockhurst, P. (1980). The oscillating rheometer-what does it
measure? Journal of Dental Research, 59, 795-9.
Crisp, S., Abel, G. & Wilson, A. D. (1979). The quantitative measurement of
the opacity of aesthetic dental filling materials. Journal of Dental Research, 58,
1585-96.
Crisp, S., Ambersley, M. & Wilson, A. D. (1980). Zinc oxide eugenol cements.
V. Instrumental studies of the catalysis and acceleration of the setting
reaction. Journal of Dental Research, 59, 44-54.
Crisp, S., Jennings, M. A. & Wilson, A. D. (1978). A study of temperature
changes occurring in setting dental cements. Journal of Oral Rehabilitation, 5,
139^4.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1976). Characterisation of
glass-ionomer cements. 1. Long term hardness and compressive strength.
Journal of Dentistry, 4, 162-6.
Crisp, S., Lewis, B. G. & Wilson, A. D. (1977). Characterisation of
glass-ionomer cements. 3. Effect of polyacid concentration on the physical
properties. Journal of Dentistry, 5, 51-6.
Crisp, S., Merson, S. A. & Wilson, A. D. (1980). Modification of ionomer
cements by the addition of simple metal salts. Industrial and Engineering
Chemistry, Product Research and Development, 19, 403-8.

382
References

Crisp, S., Merson, S., Wilson, A. D., Elliott, J. H. & Hornsby, P. R. (1979). The
formation and properties of mineral-poly acid cements. Part 1. Ortho- and
pyro-silicates. Journal of Materials Science, 14, 2941-58.
Crisp, S., Prosser, H. J. & Wilson, A. D. (1976). An infrared spectroscopic study
of cement formation between metal oxides and aqueous solutions of
poly(acrylic acid). Journal of Materials Science, 11, 36-48.
Crisp, S. & Wilson, A. D. (1974). Reactions in glass-ionomer cements: I.
Decomposition of the powder. Journal of Dental Research, 53, 1408-13.
Darvell, B. W. (1990). Uniaxial compression tests and the validity of indirect
tensile strength. Journal of Materials Science, 25, 757-80.
Deacon, G. B. & Phillips, R. (1982). Relationship between the carbon-oxygen
stretching frequency of carboxylate complexes and the type of carboxylate
coordination. Coordination Chemistry Reviews, 33, 227-50.
Gillam, E. (1969). Materials under Stress. London: Newnes-Butterworth.
Gillmore, Q. A. (1864). Practical Treatise on Limes, Hydraulic Cements and
Mortars. New York.
Hill, R. G. & Wilson, A. D. (1988). Some structural aspects of glasses used in
ionomer cements. Glass Technology, 29, 150-8.
Hill, R. G., Wilson, A. D. & Warrens, C. P. (1989). The influence of poly(acrylic
acid) molecular weight on the fracture toughness of glass-ionomer cements.
Journal of Materials Science, 24, 363-71.
Hodd, K. A. & Reader, A. L. (1976). The formation and hydrolytic stability of
metal ion-polyacid gels. British Polymer Journal, 8, 131-9.
Hondras, G. (1959). The evaluation of Poisson's ratio and the modulus of
materials of a low tensile resistance by the Brazilian (indirect tensile) test with
particular reference to concrete. Australian Journal of Applied Science, 10,
245-68.
Kemp, D. S. & Vellaccio, F. (1980). Organic Chemistry. New York: Worth.
Kendall, K. (1978). Complexities of compression failure. Proceedings of the
Royal Society of London, A 361, 245-63.
Kent, B. E., Fletcher, K. E. & Wilson, A. D. (1970). Dental silicate cements. XL
Electron probe studies. Journal of Dental Research, 49, 86-92.
Kent, B. E., Lewis, B. G. & Wilson, A. D. (1973). The properties of a
glass-ionomer cement. British Dental Journal, 135, 322-6.
Knibbs, P. J., Plant, C. G. & Pearson, G. J. (1986). A clinical assessment of an
anhydrous glass-ionomer cement. British Dental Journal, 161, 99-103.
Lion Corporation. (1980). Dental cements. Nihon Kokai Tokkyo Koho
80,139,311. Chemical Abstracts, 94: 903488, 1981.
Mackay, K. M. & Mackay, R. A. (1972). Introduction to Modern Inorganic
Chemistry. London: Intertext Books.
Mehrotra, R. C. & Bohra, R. (1983). Metal Carboxylates. London and New
York: Academic Press.
Moore, W. J. (1972). Physical Chemistry, 5th edn. London: Longman Group
Ltd.
Nicholson, J. W., Brookman, P. J., Lacy, O. M., Sayers, G. S. & Wilson, A. D.
(1988a). A study of the nature and formation of zinc polycarboxylate cement

383
Experimental techniques

using Fourier transform infrared spectroscopy. Journal of Biomedical


Materials Research, 22, 623-31.
Nicholson, J. W., Brookman, P. J., Lacy, O. M. & Wilson, A. D. (1988b).
Fourier transform infrared spectroscopic study of the role of tartaric acid in
glass-ionomer dental cements. Journal of Dental Research, 67, 145-4.
Nicholson, J. W., Wasson, E. A. & Wilson, A. D. (1988). Thermal behaviour of
films of partially neutralised poly(acrylic acid). 3. Effect of calcium and
magnesium ions. British Polymer Journal, 20, 97-101.
Nicholson, J. W. & Wilson, A. D. (1987). Thermal behaviour of films of
partially neutralised poly (aery lie acid). 1. Effect of different neutralising ions.
British Polymer Journal, 19, 67-72.
O'Neill, I. K., Prosser, H. J., Richards, C. P. & Wilson, A. D. (1982). Nuclear
magnetic resonance spectroscopy of dental materials. 1.31P studies on
phosphate-bonded cement liquids. Journal of Biomedical Materials Research,
16, 39-49.
Outwater, J. O., Murphy, M. C, Kumble, R. G. & Berry, J. T. (1974). Double
torsion techniques as a universal fracture toughness test method. Fracture
Toughness and Slow-Stable Cracking, ASTM Special Technical Publication
559, pp. 127-37, American Society for Testing and Materials.
Paffenberger, G. C, Swaney, A. C, Schoonover, I. C, Dickson, G. & Glasson,
G. F. (1949). An investigation of Diafil, a dental silicate cement. Journal of
the American Dental Association, 39, 283.
Pluim, L. J. & Arends, J. (1987). The relationship between salivary properties
and in vivo solubility of dental cements. Dental Materials, 3, 13-18.
Polakowski, N. H. & Ripling, E. J. (1966). The Strength and Structure of
Engineering Materials, Chapter 10. Englewood Cliffs, New Jersey: Prentice-
Hall Inc.
Prosser, H. J., Richards, C. P. & Wilson, A. D. (1982). NMR spectroscopy of
dental materials. II. The role of tartaric acid in glass-ionomer cements.
Journal of Biomedical Materials Research, 16, 431-45.
Prosser, H. J., Brant, P. J., Scott, R. P. & Wilson, A. D. (1983). The cement-
forming properties of phytic acid. Journal of Dental Research, 62, 598-600.
Prosser, H. J., Powis, D. R., Brant, P. & Wilson, A. D. (1984). Characterization
of glass-ionomer cements. 7. The physical properties of current materials.
Journal of Dentistry, 12, 231-40.
Prosser, H. J., Powis, D. R. & Wilson, A. D. (1986). Glass-ionomer cements of
improved flexural strength. Journal of Dental Research, 65, 146-8.
Reed, S. J. B. (1973). In Anderson, C. A. (ed.) Microprobe Analysis. New York:
Wiley-Inter science.
Selenrath, Th. R. & Gramberg, J. (1958). Stress-strain relations and breakages
of rocks. In Walton, W. H. (ed.) Mechanical Properties of Non-metallic Brittle
Materials, Chapter 6, pp. 79-105. London: Butterworths.
Setchell, D. J., Teo, C. K. & Kuhn, A. T. (1985). The relative solubilities of four
modern glass-ionomer cements. British Dental Journal, 158, 220-2.
Skoog, D. A. & West, D. M. (1980). Principles of Instrumental Analysis, 2nd
edn, Chapter 8. Tokyo: Holt-Saunders Japan Ltd.

384
References

Smith, D. C. (1968). A new dental cement. British Dental Journal, 125, 3 8 1 ^ .


Sorrell, C. A. (1977). Suggested chemistry of zinc oxychloride cements. Journal
of the American Ceramic Society, 60, 217-20.
Sorrell, C. A. & Armstrong, C. R. (1976). Reactions and equilibria in
magnesium oxychloride cements. Journal of the American Ceramic Society, 59,
51-4.
Tay, W. M. & Braden, M. (1981). Dielectric properties of glass-ionomer
cements. Journal of Dental Research, 60, 1311-14.
Tay, W. M. & Braden M. (1984). Dielectric properties of glass-ionomer
cements-further studies. Journal of Dental Research, 63, 74-5.
Tyndall, J. (1866). On calorescence. Philosophical Magazine, Series 4, 31,
386-96; 435-50.
Watts, D. C. (1979). C-13 NMR spectroscopic analysis of polyelectrolyte cement
liquids. Journal of Biomedical Materials Research, 13, 423-35.
Williams, D. H. & Fleming, I. (1973). Spectroscopic Methods in Organic
Chemistry, 2nd edn. London: McGraw-Hill.
Wilson, A. D. (1978). The chemistry of dental cements. Chemical Society
Reviews, 1, 265-96.
Wilson, A. D. (1982). The nature of the zinc polycarboxylate cement matrix.
Journal of Biomedical Materials Research, 16, 549—57.
Wilson, A. D. & Batchelor, R. F. (1968). Dental silicate cements: III.
Environment and durability. Journal of Dental Research, 41, 115-20.
Wilson, A. D. & Batchelor, R. F. (1971). The consistency of dental cements. The
specification test for filling materials. British Dental Journal, 130, 437-41.
Wilson, A. D., Crisp, S. & Ferner, A. J. (1976). Reactions in glass-ionomer
cements: IV. Effect of chelating comonomers on setting behavior. Journal of
Dental Research, 55, 489-95.
Wilson, A. D., Groffman, D. M., Powis, D. R. & Scott, R. P. (1986). A study of
variables affecting the impinging jet method for measuring the erosion of
dental cements. Biomaterials, 7, 217-20.
Wilson, A. D., & Kent, B. E. (1968). Dental silicate cements. V. Electrical
conductivity. Journal of Dental Research, 47, 463-70.
Wilson, A. D. & Kent, B. E. (1970). Dental silicate cements. IX. Decomposition
of the powder. Journal of Dental Research, 49, 7-13.
Wilson, A. D. & McLean, J. W. (1988). Glass-ionomer Cement. Chicago,
London, etc.: Quintessence Publishers.

385
Index

abietic acid (rosin) 322, 334, 338-9 obstacles to 93-4


acid-base balance of polyalkenoates 94-6
in glasses 123-5 of zinc polycarboxylates 107
in silicate rocks 17 adsorption of carboxylates (alkenoates)
acid-base (AB) cements 96-7
crystallinity in 8-10, 205-13 agriculture 4
formation 7-11, 307-8 A12O3 cements 102
theory 1-5, 5-26, 307-8 A1 2 O 3 -P 2 O 5 -H 2 O 199-201
acid-base concepts 12-26 aldehydic aromatic acid cements 318,
aprotic acids 6, 17-20 321
Arrhenius theory 14-5, 19 alkanoate adsorption on hydroxyapatite
Bronsted-Lowry theory 15-6, 19-20, 96-7
48, 284 alkanoate bonding modes 363
Cartledge theory 20-1 alkanoic acids 5-6, 308, 315, 318,
classification 22-3 320-1, 337, 348, 351
hard acids and bases see HSAB theory alkoxy aromatic acids 318, 320, 337
history 1 2 ^ citric acid 308
HSAB theory 24-6 dimer (dimerized fatty) acids 351
ionic potential 20-1 2-ethoxybenzoic acid 6, 318, 320-1,
ionization potential 21-2 337, see also EBA cement
Lewis theory 17-20 malic acid 6, 308, 315
Lux-Flood theory 17, 19-20 mellitic acid (benzenehexacarboxylic
relevance 19-20 acid) 6, 315
soft acids and bases see HSAB theory 1,2,3-propanetricarboxylic acid 6,
solvent system theory 16-7, 19 315
strength 20-2 pyruvic acid 6, 308, 315
Usanovich theory 18-20 salicylates 318, 348
acid-decomposable glasses 6, see also tartaricacid 6, 308, 315
aluminosilicate glasses, tricarballylic acid 6
aluminoborate glasses alkenoic acid polymers see poly(akenoic
acid-etching of enamel 93 acid)s
acids for cement-formation see cement- alkoxy aromatic acid cements 318, 320,
forming acids 337
acrylic acid allyl-2-methoxyphenol cements 318, 321
copolymers see poly(alkenoic acid)s aluminium coordination 101-2, 120-1,
homopolymer see poly(acrylic acid), 123, 125, 129, 131,137-8
poly(alkenoic acid)s alumino complexes 244
adhesion 1, 4, 56, 92-7, 107, 152-4, 381 fluoride 135-8,244
to bone 94-6, 111 phosphate 57, 85, 200-1, 210, 244-5
to dentine and enamel (tooth) 92-6 aluminoborate glasses 165-6
of glass polyalkenoate cements 152-4 aluminophosphoric acids 57, 200-1
to hydroxyapatite 95-6 aluminosilicate gels 91

386
Index
aluminosilicate glass cements 307-9, see 147, 166-8, 204, 214, 235-6, 320-1,
also glass polyalkenoate cement, 333
dental silicate cements fire resistant materials 283
aluminosilicate glasses 2, 6, 9, 90, floor fabrication 2, 290
117-32,236-40,310,314-5 floor repair 222
acid-base balance in 123 foundry sands 2
acid-decomposition of 119-23, 127-8 handyman materials 3
acid-washing of 163 horticulture 4
Al 2 O 3 :SiO 2 123-7 human health care 4
aluminium coordination 120-1, 123-5, impression material 335
128-9, 130, 137 insulating materials 283
anorthite in 122, 130 investment materials 222
apatite in 125 nuclear 283
beryllium-containing 236 plasters 290
Ca:Al 123-5 pulp capping 347
calcium-containing 6, 118-20, road repairs 222
123-33,236-7,310,314-5 runway repairs 222
composition and properties 118-9, slip casting 2-3,91, 169
122-9, 238^1 soil consolidation 90
coordination polyhedra 119-21 splint bandage 2, 91, 117, 168
corundum in 125-6 surfacing 4
electron micrographs 128, 238-9 sustained release devices 3, 4, 157-8,
fluoride role 118-9, 129^30, 236 222, 304
fluoride type 117-9, 125-132, 135, underwater cements 2, 91
140, 236-40 architectural applications 283
fluorite in 125-6, 129-30 aromatic carboxylic acid 96-7, 347
indium-containing 237 attenuated total reflectance (infrared)
lanthanum-containing 117-9 spectroscopy (ATR) 359; see also
NMR studies 121, 125, 128, 131 infrared spectroscopy
oxide type 117, 122-6, 135, 236-40
phase-separation 126-8, 130-1, 238-9 B 2 O 3 cements 102, 312
SiO 2 -Al 2 O 3 -CaF 2 119, 126-9, 238 BaO cements 204, 318, 338
SiO 2 -Al 2 O 3 -CaF 2 -AlPO 4 119,131 bactericides 335
SiO 2 -Al 2 O 3 -CaF 2 -AlPO 4 -Na 3 AlF 6 - bases for cement formation see cement-
A1F3 119,131-2 forming liquids
SiO 2 -Al 2 O 3 -CaO 118, 123-6, 237^0 battlefield dental material 333
SiO 2 -Al 2 O 3 -CaO-CaF 2 119, 130 BeO and Be(OH)2 cements 201-2
SiO 2 -Al 2 O 3 ^CaO-Na 2 O 118 Bi2O3 cements 102, 201-2, 312-3,
SiO a -Al a O 8 -CaO-P a O 6 118,239-40 318
spinodal decomposition 130 bioactive materials 3
strontium-containing 117-9 bioadhesion see adhesion
structure 118-131,238-9 bis-GMA 170-1
transition temperatures 130 Bjerrum ion-pairs 67
types 118-9,235-7 Boltzmann distribution 61
animal husbandry 4 bone 2
applications adhesion to 94-6, 111
agriculture 4 bone cements 2, 90-1, 117, 147,
animal husbandry 4 161-2, 168
architecture 283 bone substitute 4, 161-2, 169
battlefield dental material 333 Born-Oppenheimer approximation 32
bone cements 2, 90-1, 117, 147, Bragg equation 367
161-2, 168 Bronsted-Lowry theory 15-6, 19-20,
bone substitute 4, 161-2, 169 48, 284
controlled release devices see sustained 3-butene 1,2,3-tricarboxylic acid/acrylic
release devices acid copolymer 91, 1 0 3 ^ , 131-2,
dental cements 2, 90-1, 103, 116-7, see also poly(alkenoic acid)s

387
Index
CaO cements 204, 318, 321, 338 oil of cloves 320-1
Ca(OH) 2 cements see calcium hydroxide orthophosphoric acid see phosphoric
chelate cements, calcium hydroxide acid
dimer (dimerized) acid cements phosphoric acid 6, 22, 56, 85,
calcium aluminosilicate glasses see 197-201
aluminosilicate glasses phytic acid 3, 5, 309-10
calcium hydroxide chelate cements 318, poly(acrylic acid) 6, 22, 56-8, 69,
347-50 70-1, 74^5, 78-9, 90-4, 97-8, 1 0 3 ^
applications 347 poly(alkenoic acid)s 56-8, 69-71,
composition 348 74-5, 78-9, 90-1, 97-8, 103-5, 360
properties 350-1 poly(phosphonic acid)s 310-1
setting 348-9 poly(vinyl phosphonic acid) see
calcium hydroxide dimer (dimerized) acid poly(phosphonic acid)s
cements 351 propylene-2-methoxyphenol 321
carboxylate adsorption on pyruvic acid 6
hydroxyapatite 95-7 salicylates 348
carboxylate bonding modes 363 syringic acid see
carboxylic acids see alkanoic acids, methoxyhydroxybenzoic acids
poly(alkenoic acid)s tannic acid 6,308,315
caries, effect of fluoride on 258 tartaric acid 6, 308, 315
CdO cements 201-2, 204, 312, 318, 321, tricarballylic acid 6
338 vanillic acid see
cement classification 7 methoxyhydroxybenzoic acids
cement-forming acids 3, 5-6, 308 zinc chloride 6, 283-9
aldehydic aromatic acids 318, 321 zinc selenate 6
alkoxy aromatic acids 318, 321, 337 zinc sulphate 6
allyl-2-methoxyphenol 318-9, 321 cement-forming cations 9, 19-22,
aromatic carboxylic acids 347 198-9, 201-4, 244
citric acid 308 cement-forming liquids see cement-
cobalt chloride, selenate, sulphate 6 forming acids
copper chloride, selenate, sulphate 6 cement-forming metal oxides 5-6, 102,
creosote 321 201-2, 312-6, 318, 321
yS-diketones 318-9,321 A12O3 102
dimer (dimerized fatty) acids 351-2 B 2 O 3 102, 312
2,5-dimethoxyphenol 318 BaO 204,318,338
2-ethoxybenzoic acid 318, 320-1, 337, BeO and Be(OH)2 201-2
see also EBA cements Bi2O3 102, 201-2, 312-3, 318
eugenol 318, 320, see also zinc oxide CaO 312,318
eugenol cements Ca(OH) 2 204, 311-2, 318, 321, 338
fluoboric acid 308 CdO 201-2, 312, 318, 321, 338
gallic acid 6, 315 CoO 312
glycerol phosphoric acid 308 Co(OH) 2 6, 202, 222, 312, 315-6
guaiacol (2-methoxyphenol) 318-9, Cr 2 O 3 , CrO 3 312
321 CuO 6, 102, 201-2, 204, 311-2,
ketoacids 318, 321 315-6, 318, 321, see also CuO
ketoesters 318, 321 cements
magnesium chloride 6, 2 8 3 ^ , 290, Cu 2 O 201-2, 311-2, see also Cu 2 O
292-6 cements
magnesium selenate 6 Fe 2 O 3 312
magnesium sulphate 6, 283-4, HgO 102,312,318,321,338
299-302 In 2 O 3 312
malic acid 6 La 2 O 3 102, 312
mellitic acid (benzene hexacarboxylic MgO 6, 102, 201-2, 204, 311-3, 321,
acid) 6, 315 338, see also magnesium phosphate
methoxyhydroxybenzoic acids 342-6 cements
2-methoxyphenols 318-9, 321 MoOo 312

388
Index

MnO 2 312 polyalkenoate 101-2


PbO 102, 312, 318, 321, 338 polyphosphonate 311—2
Pb 3 O 4 201-2, 312
SnO 201-2,312 Debye-Hiickel theory 43-5, 67
WO 3 312 degradative studies (chemical) 105,
Y 2 O 3 102, 312 136-9, 244-7, 339, 360-1
ZnO 6, 102, 201-2, 204, 311-3, 318, dehydration
see also zinc polyalkenoate cement, and gelation 72, 84
zinc phosphate cement, zinc oxide and precipitation 77-9
eugenol cements, EBA cements dental cements 90-1, 103, 106-13,
cement gels 8-10 116-7, 146-7, 166-8, 204, 214,
cementitious bonding 7-11, 307-8 235-6, 320-1, 333
cementitious substances dental impression material 335
mortars 1 dental materials 2
plaster of Paris 1, 7 dental plaque 379
Portland cement 1, 2, 5, 7 dental silicate cement 235-63, 366, 369,
refractory cements 197 381
silicate/silica gel cements 140 applications 236-7, 249
chelate agents 6 composition: glasses 238-9, see also
chelate cements 318-52 aluminosilicate glasses; liquids 218,
chelate formation 71 241-3, see also phosphoric acid
citric acid cement 308 history 235-7
cloves, oil of 321 modified materials 237-8: acid-
CoO cements 312 resistant 237; indium-containing
Co(OH) 2 cements 202, 222, 312, 315-6 237
composite resin 154—6, 235 properties 253-65: acid erosion
compressive strength 359, 370-2, see 259-60; bacterial contamination
also experimental techniques 261; biological 260-1; erosion
concrete 1 255-8; fluoride release 255-6,
condensation see counterion 257-8; physical 253-7;
condensation powder: liquid effect 256;
condensation cements 7 translucency 255; water absorption
configuration see polymer conformation 256-7
conformation see polymer conformation setting 243-9: electrical conductance
consistency 375, 378 247, 366-7; exotherm 381;
controlled release devices see sustained hydration 247, 249; ion release
release devices 25-7; infrared spectra 243; NMR
copper phosphate cements 201-2, 221-2 spectra 245,252,365-6;
coulombic forces 80-2 permittivity 367; pH changes
counterion 249; precipitation of ions 243-8;
binding 7-8, 59-83, 106, see also ion salt formation 244-8; silica
binding formation 247, 250-1; water
distribution 59-63, 82 deficiency effect 249
condensation 63-7, 78 structure 249-53: electron
creosote cement 321 microscopy 250-1; electron probe
Cr 2 O 3 , CrO 3 cements 312 analysis 250-2, 369; element
crystallinity in cements 8-10, 205-13 distribution 250-3; optical
CuO cements 102, 201-2, 221-2, 231, microscopy 250
311-2, 315-6, 318,321 dentine 91
chelate 231 bonding to 92-5, 111, 152-4
miscellaneous 315-6 composition 94
phosphate 201-2, 221-2 treatment 152-^
polyalkenoate 101-2 desolvation
polyphosphonate 311-2 and gelation 72, 84
Cu 2 O cements 6, 201-2, 220-1,311-2 and precipitation 77-9
phosphate 201-2, 220-1 diffusion in water 37

389
Index
/?-diketone cements 318-9, 321 experimental techniques 359-381
dilatometry 59 adsorption 95-7
dimer acid (dimerized fatty acid) cements compressive strength 359, 370-2
351-2 degradative studies (chemical) 105,
2,5-dimethoxyphenol cements 318 136-9, 244-7, 339, 360-1
dipole interactions 82 diametral compressive strength 372
dissolution dilatometry 59
of polymers in water 45-7 dipole interactions 82
of salts in water 41 double torsion test of fracture
distribution functions for ion-pairs toughness 374
67-8, 72-3 electrical conductance 247, 325-6,
double-torsion test of fracture toughness 359, 366-7
374 electron diffraction 33, 35
dual-cure resin cement see resin glass electron probe analysis 105, 144—5,
polyalkenoate cement 233, 247, 250
erosion 378-9
EBA 6,318,320,337 exotherm measurements 147, 380-1
EBA cements 320, 337-47 flexural strength 359, 370, 372-3
EBA divanillate cements 344-5 four-point bend test 374
EBA eugenol cement 337-42 Fourier transform infrared
composition 338-9 spectroscopy (FTIR) 105, 359, 364,
physical properties 340-2 see also infrared spectroscopy
setting 339-40 fracture toughness 373-4
structure 339-40 impinging jet 158-9, 216-8, 341, 379
EBA-methoxyhydroxybenzoate cements infrared spectroscopy 99-101, 105-6,
342-4 137, 142, 198, 210, 243-4, 247,
composition 343-3 250-2, 311, 323-5, 337, 339-40, 343,
properties 3 4 2 ^ 348-9, 351, 359, 361-4
setting 343-4 leaching studies 106, 378-80
EBA polymer cement 344-6 mass spectrometry 42
EBA syringate cement see EBA- mechanical properties 370-4
methoxyhydroxybenzoate cements neutron diffraction 33, 35-6, 42
EBA vanillate cement see EBA- NMR spectroscopy 59, 141, 198,
methoxyhydroxybenzoate cements 200-1, 245, 359, 364-66
'egg-box' model for gelation 85 NMR spectroscopy solid state 121,
electrical conductance 247, 325-6, 125, 131, 145-6, 252
359, 366-7 opacity 127, 148, 151-2, 379-80, see
electron diffraction 33, 35 also translucency
electron probe microanalysis 105, optical microscopy 143, 249
144-5, 233, 247, 250, 369-70 optical properties 127, 146-8, 151-2,
electrons, hydrated 44 166, 379-80
enamel (tooth) parallel plate plastometer 377
acid-etching 93, 153—4 permittivity 325-6, 359, 367
aluminium uptake 258 Raman spectroscopy 198
bonding to 92-6, 111-2, 1 5 2 ^ rheometry 141,374-8
composition 94 scanning electron microscopy 106,
conditioning 15 3—4 128, 226-30, 232-3, 329, 331
fluoride uptake 158, 258 setting measurements 374-8
opacity 152 tensile strength 370
translucency 152 titrimetric methods 311
erosion measurement 378-9 transition temperatures 130
2-ethoxybenzoic acid see EBA translucency 147, 151-2, 166, 359,
ethylene glycol dimethacrylate 170 379, see also opacity
eugenol 2, 6, 318, 321, see also zinc transmission electron microscopy 145
oxide eugenol cements viscosity measurements 141
exotherm measurements 147, 308-1 water analysis 105-6

390
Index
X-ray diffraction (XRD) spectroscopy 123-31; glasses see aluminosilicate
9-10, 33, 35, 47, 51, 105, 125-6, 130, glasses; liquids see poly(alkenoic
198, 202-3, 208-9, 224^31, 250, acid)s; metal fluoride additives
283-6, 293, 323, 359, 367-8 134, 163; oxide glasses 135;
poly(alkenoic acid), molecular mass
Fe 2 O 3 cement 312 effect 163; poly(alkenoic acid)
fire-resistant materials 283 type, effect on properties 132;
flexural strength 359, 370, 372-3 tartaric acid effect 133-4, 376
floor fabrication 290 history 116—7
floor repair 222 light-cured see resin glass
fluoboric acid cement 308 polyalkenoate cement
fluoride properties 94-7, 117, 146-165, 174:
additive to glass polyalkenoate cement acid erosion 158-9; acid-etching of
133 155-6; adhesion 94-7, 117, 147,
alumino complexes 135-8, 244 152-6, 164, 174-5; biological
bone remineralization 161-2 159-61; bonding to composite resins
and caries 258 154-6; bonding to tooth material
cement reaction 106-8 152-4; bone remineralization
glass polyalkenoate cement reactions 161-2; consistency of pastes 148;
133-41 creep 148; erosion 148, 156-9,
release from cements 117, 147, 157-8, 165;exotherm 147; fluoride release
255-8, 379 157-8, 379; glass composition, effect
fluoride glasses 136-46 123-31; modulus 149, 164;
foundry sand 2 opacity, see translucency; plasticity
four-point bend test 374 147-8; poly(alkenoic acid),
Fourier transform infrared spectroscopy molecular mass effect 163;
(FTIR) 105, 359, 364, see also poly(alkenoic acid) type, effect 132;
infrared spectroscopy setting behaviour 122-8, 132-3,
fracture toughness 373-4 165; strength 122, 125, 127, 132-4,
freezing point of water, D-structure 38 138-9, 147-50, 163-6, 372-3; stress
Fuoss distribution 68-9 relaxation 148-9; tartaric acid,
effect 133-4; translucency 127,
gallic acid cement 6 147-8, 151-2, 166, 380;
gelatinising minerals 6, 114-6 viscoelasticity 148-9; water
gels and gelation 8-11, 49, 56, 64, 83-5, absorption 156-7; wear 159
138 resin hybrids see resin glass
cations, gel-forming 9 polyalkenoate cement
* egg-box' model 8 5 setting reaction 98-9, 134-43:
hydration 72, 83-4 aluminofluoride complexes in 137-8;
ion binding 84-5 coordination changes 145-6;
models 10-1,85 desolvation 135; exotherm 147;
neutralization 84 fluoride, effect of 13^41, 134,
polyion interaction 84 163; Fourier transform infrared
polymer conformation 77, 84 spectra 364; gelation 134-5,137-8;
structures 10-1,85 glass ions release 361; hydration
Gibbs equation 40 139; infrared spectra 137, 362,
Gillmore needle 375 364; ion binding during setting
glass-ionomer cement see glass 137-9; NMR spectra 145-6, 366;
polyalkenoate cement, glass pH changes 134, 136, 138; polymer
polyphosphonate cement configuration 135; precipitation of
glass polyalkenoate cement 2-4, 56, ions 137-8; release of ions from
90-1, 116-175, 235 134, 137-8; rheometry 376; salt
applications 117, 147, 160-2, 166-9 formation 135; silicic acid and
composition 123-46, 162-3: additives silica gel formation 134, 139-40,
133-4, 376; fluoride glasses 145-6; solvation 139; strength
136-46; glass effect on properties increases 139, 148-9; tartaric acid,

391
Index
setting reaction (cont.) hydrologic cycle 32
effect 131-5, 141-3, 162-3; hydrophobic interactions 40-1
viscosity increases 135, 141 hydrosphere 32
structure 138, 142-6: electron hydroxyapatite 95-7
microscopy 143-145; electron adsorption on 95-7
probe analysis 145-6, 369-70; aluminium uptake 258
element distribution 144-5; carboxylate uptake 95-7
molecular structure 99-101; optical fluoride uptake 258
microscopy 142-3; reptation hydroxydimethyl acrylates 170
model 139; water states 31,49, hydroxyethyl methacrylate (HEMA) 3,
146 169-173
glass polyphosphonate cement 117,
314-5 ice structures 35-6
glass transition temperatures 52 impinging jet test 158-9, 216-8, 341,
glasses see aluminoborate glasses, 379
aluminosilicate glasses In 2 O 3 cement 312
glycerol phosphoric acid cement 308 infrared spectroscopy 359, 361-4
guaiacol cement 318, 321 attenuated total reflectance (ATR)
Gurney potential 45 359
calcium hydroxide chelate cements
handyman materials 3 348-9
hard acids see HSAB theory calcium hydroxide dimer cements 351
hard and soft acids and bases see HSAB dental silicate cement 243, 247, 250-1
theory EBA (2-ethoxybenzoic acid) cements
hard bases see HSAB theory 339^0
HEMA see hydroxyethyl methacrylate EBA-methoxyhydroxybenzoate
hexaquo cations 16, 47, 284 cements 343
HgO cements 102, 312, 318, 321 2-ethoxybenzoic acid (EBA) 337
horticulture 4 eugenol 323-^
HSAB theory 2 ^ 6 , 4 7 - 8 Fourier transform (FTIR) 105, 359,
human health care 4 364
hybrid light-cured cements see resin glass glass polyalkenoate cement 136-7,
polyalkenoate cements 142
hydrated electrons 44 metal oxide polyphosphonate cements
hydration 74-9, 139, 247, 249, 307 311
and gelation 49-50, 72, 77-9, 84 molecular structure of polyalkenoate
and ion binding 76-7 (polyelectrolyte) cements 99-101
and ionization 74—7 phosphate bonded cements 198, 210,
of ions 31,41-4,47-8 244, 247, 250-2
of polyions 31, 73-5 polyalkenoate cements 104-5, 136-7,
and precipitation 77-9 142
in solid state 47 zinc phosphate cement 210
hydration number 42 zinc polycarboxylate cement 104-5
hydration regions see hydration shells ZOE cements 323-6
hydration shells 42-3, 49-50, 72-7 insulating materials 283
hydration states 31, 49-50, 59 investment materials 222
hydration zones see hydration shells ion binding 7-8, 59-83, 106
hydraulic cements 7 cation effect 65-67
hydrides 33 cements 106
hydrogel 1 complex formation effect 69-70
hydrogen bonding density changes 73-4
in cements 9, 203 dipole changes 7
in 2-ethoxybenzoic acid 338 electric conductance changes 59
in phenols 321-2 hydration effects 72-9
in phosphoric acid 198 molar volume changes 74
of water 38 polymer type effect 70-2

392
Index
refractive index changes 63, 73-5 magnesium oxysulphate cement
turbidity 79 299-304
ultrasonic changes 74 phases 300-2: MgO-H 2 SO 4 -H 2 O
viscosity changes 78 301; MgO-MgSO 4 -H 2 O 300-2
ion coordination 47-8, 69, 99-101, porosity 303
117 properties 302-304
ion-ion interactions 44-5, see also ion setting chemistry 299-300
pairs magnesium phosphate cements 2, 102,
ion-pairs 49, 72-3, 79 204, 222-35
contact 72-3, 79 aluminum acid phosphate type 233-5
distribution functions 67-8, 72-3 ammonium dihydrogen phosphate type
hydration (solvation) of 72-3 224-31
solvent separated 72-3, 79 ammonium polyphosphate type 232
types 72-3 composition 222-3
Irving-Williams series 69-70 diammonium phosphate type 231-2
itaconic acid/acrylic acid copolymer 56, phosphoric acid type 224
91, 97-8, 103-4, 132-3, see also types 223-4
poly(alkenoic acid)s magnesium polyalkenoate cement 235
magnesium polyphosphonate cement
jet test see impinging jet test 312-3
magnesium titanate phosphate cement
ketoacid cements 318, 321 235
ketoester cements 318, 321 maleic acid/acrylic acid copolymer 56,
91, 97-8, 103-4, 132-3, see also
La 2 O 3 cements 102, 312 poly(alkenoic acid)s
leaching studies 106 malic acid cements 6
Lewis acids 6, 18, 22-4, 47, 284 mass spectrometry 42
Lewis theory 17-20 mellitic acid cement 6, 315
light-cured cements 3-4, see also resin metal oxide cements
glass polyalkenoate cement eugenol cements 321; see also zinc
liquids for cement formation see cement- oxide eugenol cements
forming acids oxysalt bonded cements 2-3, 5-6,
Lux-Flood theory 17-20 304-5, see also magnesium
oxychloride cement, magnesium
magnesia cements 283, see also oxysulphate cement, zinc
magnesium oxychloride cement, oxychloride cements
magnesium oxysulphate cement, phosphate cements 201-2, 221-22,
magnesium phosphate cements see also CuO cements, Cu 2 O
magnesia phosphate cements see cements, dental silicate cement,
magnesium phosphate cements magnesium phosphate cements, zinc
magnesium chloride solution 6, 283-4, phosphate cement
290, 292-6 polyalkenoate cements 90, 102-3, see
magnesium oxide 103-4, 290-1 also glass polyalkenoate cement, zinc
cements 283, see also magnesium polycarboxylate cement
oxychloride cement, magnesium polyelectrolyte cements see
oxysulphate cement, magnesium polyalkenoate cements,
phosphate cements polyphosphonate cements
deactivation 290-1 polyphosphonate cements 311-3
magnesium oxychloride cement 2,31, metal oxides for cement formation see
51, 283, 290-9 cement-forming metal oxides
applications 290 metal poly(acrylic acid) complexes
components 290 69-70
phases 294: MgO-MgCl 2 -H 2 O methods see experimental techniques
294-5 methoxyhydroxybenzoate cements see
setting chemistry 291-3 EBA-methoxyhydroxybenzoate
setting kinetics 293—4 cements

393
Index
methoxyhydroxybenzoic acids 342-6 parallel plate plastometer 377
2-methoxyphenol cements 318-9, 321, PbO cements 102, 312, 318, 321, 338
342 Pb 3 O 4 cements 201-2,312
MgO cements 6, 102, 201-2, 204, 311-3, permittivity 325-6, 359, 367
318, 321, see also magnesium phenolic cement-formers 5-6, 308,
phosphate cements 318-21
micromechanical attachment 93, 154-5 allyl 2-methoxyphenol see eugenol
mineral cements 2,5-dimethoxyphenol 318
ionomer 90, 114-6 eugenol 318, 321, see also zinc oxide
phosphate 265 eugenol cements
polyalkenoate 90, 114-6 gallic acid 6, 315
MnO 2 polyphosphonate cement 312 guaiacol (2-methoxyphenol) 318, 321
MoO 3 polyphosphonate cement 312 2-methoxyphenols 318-9, 321
model materials 91 propylene-2-methoxyphenol 321
molecular dynamics 42 tannicacid 6, 308, 315
mortars 1-2 phlogiston theory 31
mouth fluids 216 phosphate-bonded cements 3, 7,
197-265
neutron diffraction 34-6, 42 liquids for 218, 241-3
NMR spectroscopy 59, 141, 198, 200-1, see also copper phosphate cements,
245, 359, 364-66 dental silicate cement, magnesium
NMR spectroscopy solid state 121, 125, phosphate cements, metal oxide
131, 145-6,252 phosphate cements, zinc phosphate
non-aqueous cements 318-52 cement
nuclear applications 283 phosphoric acid 2, 5-6, 22, 56, 85,
197-204, 241-3
oil of cloves 320-1 aluminium complexes 198-200
opacity see optical properties cations in 198-201, 203-4
optical microscopy 143, 249 cement forming liquids 218, 241-3
optical properties 127, 146-8, 150-2, concentration effect on cement
165, 379-80 properties 218, 241-3
opacity 127, 148, 151-2, 379-80, see dimers H 6 P 2 O 8 198
also translucency hydrogen bonding 198
translucency 1, 3, 147, 151-2, 166, infrared spectra 198
see also opacity NMR spectra 198, 200
organic polymers 1 phase diagrams 199-200:
orthophosphoric acid see phosphoric A1 2 O 3 -H 2 O-P 2 O 5 199-200;
acid ZnO-H
orthosilicic acid 7, 121, 134, 139-40, pK 197
243-4, 247 properties 197-8
osmotic forces 80-1 structure 198
osteogenesis 162 triple ion H5P2Og 198
oxychloride cements 3, 5-6 see also dental silicate cement,
calcium 304 magnesium phosphate cements,
cobalt 305 mineral phosphate cements, zinc
magnesium 2, 31, 51, 283, 290-9 phosphate cement
zinc 2, 31, 51, 285-90 phytic acid [myo-inositol
oxygen polyhedra 9, 120, 123, 125, hexakis(dihydrogen phosphate)]
128-9 cement 3, 5, 309-10
oxysalt bonded cements 2-3, 5-6, 31, 51, cements 309-10, 360
283-305 pipelines 2, 91
components 6, 284 plaster of Paris 1,91
setting 284 plasters 290
oxyselenate cements 3, 6 Poisson distribution 61
oxysulphate cements 3, 5-6 poly(acrylic acid) 6, 22, 31, 46, 49, 56-9,
magnesium 283-4 69-71, 90-1, 97-8, 103-4, 132-3,

394
Index
237, 360, 366, see also poly(alkenoic poly(ethylene maleic acid) 71, 75, 98,
acid)s 360
poly(acrylic acid), modified vinyl polyHEMA see poly(hydroxyethyl
containing 3, 170-2 methacrylate)
polyalkenoate cements 2, 3, 90-175 poly(hydroxyethyl methacrylate) 169-73
molecular structure 99-101 polyion-polyion interactions 82-3
setting 98-9 polyions
see also glass polyalkenoate cement, attraction 82-3
mineral polyalkenoate cements, zinc conformation 58-9
polycarboxylate cement extension 82-3
polyalkenoates hydration 73-5
adhesion 94-6 interaction 82-3
adsorption 96-7 ion binding to 56-7, 59-64
complexes 69-70 repulsion 82-3
poly(alkenoic acid)s 56-8, 69-71, 74-5, poly(itaconic acid) 71, 75
90-1, 97-8, 103-5, 132-3, 360 poly(maleic acid) 71, 75
acrylic acid homo and copolymers polymer configuration see polymer
56-7, 70-1, 74-5, 91, 97-8, 103—4, conformation
132-3, 360 polymer conformation 58-9, 79-81
binding properties 90 coulombic attraction effects 80-1, 84
3-butene 1,2,3-tricarboxylic acid factors affecting 79-81
copolymers 91-2, 103-4, 132 gelation and 81, 83^4
and cement properties 132,162-3 neutralization effects 84
concentration effect on cement osmotic pressure, effect on 80-1, 84
properties 132, 162 polymer morphology see polymer
conformation 46, 58-9 conformation
dissolution in water 45-7 polymer shape see polymer conformation
ethylene-maleic acid homopolymer polymerization 1, 3
71, 75, 98, 360 poly(methacrylic acid) 70-1, 75, 360
gelation 83-5 polyphosphonate cements 117,310-5
in glass polyalkenoate cements 132-3 glass (aluminosilicate) 314-5
hydration 74-7 metal oxide 311-4
ion binding 77-9 poly(phosphonic acid)s 56, 311
itaconic acid polymers 56, 71, 75, poly(vinyl methyl ether-maleic acid) 71,
91-2, 103-4, 132-3 360
maleic acid polymers 56, 71, 75, 91-2, poly(vinyl phosphonate) cements see
103^4, 132-3 polyphosphonate cements
methacrylic acid polymer 70-1, 75, poly(vinyl phosphonic acid) 56-7, 311
360 Portland cements 1, 2, 5, 298
modified vinyl containing 3, 170-3 pottery 1
molecular mass 98, 103-4 precipitation cements 7
polymer preparation 97-8 propylene-2-methoxyphenol cement 321
preparation 97-8 protonic acids 6, 14—6
solid form use of 163 pulp capping 347
vinyl methyl ether maleic acid polymer pyruvic acid cement 6
71, 360
zinc polycarboxylate cements 103-4 Q nomenclature for silicates 114, lib,
polycarboxylates see polyalkenoates 120, 125, 129, 131, 137
polyelectrolyte cements 2-3, 90-169,
310-1, see also polyalkenoate Raman spectroscopy 198
cements and polyphosphonate rapid repair materials 3, 4
cements reaction cements 7
polyelectrolytes 31, 45-6, 56-85, 91-8, refractive index 63, 74—5
200-1, 310-1, see also poly(alkenoic refractory cements 197
acid)s, polyions, poly(phosphonic resin glass polyalkenoate cement 3, 4,
acid) 16^-175

395
Index
composition 170-3 turbidity measurements 78
light activation 171
properties 173 ultrasonic methods 74
setting reaction 170-2 underwater cement 2, 91
structure 172-3 Usanovich theory 18-20
rheometry 141, 374-8
road repairs 222 V-structure of water 37-40
runway repairs 222 vanillate cements see
EBA-methoxyhydroxybenzoate
salicylate cements 348 cements
salts, dissolution in water 41 vinyl polymerization 3
scanning electron microscopy 106, viscoelasticity 148-9, 341, 375
128, 226-30, 233, 329, 331 viscosity measurements 59, 141
setting measurements 374-8
silica gel 7, 9, 121, 139^0, 247, 307 water 30-55
silicate cements 7 in AB cements 30-1: hydration 31,
silicate glass 9 74-9, 139, 247, 249, 307; in cement
silicate minerals 6, 90, 114—6, 265 formation 7, 249; component of
silicic acid 7, 9, 121, 140, 244, 247 30, 48-51; ligands 101, 137-8;
silicophosphate cement 263-5 ' loosely bound' (evaporable)
slip casting 3,91, 169 49-50; as plasticizer 31, 51-2; as
SnF 2 113 reaction medium 247, 307; as
SnO cements 201-2, 312 solvent 30, 48, 249, 307; states
soft acids see HSAB theory 49-50;' tightly bound' (non-
soft bases see HSAB theory evaporable) 49-50
soil consolidation 90 as a base 49
solvation see hydration coordination to ions 47-8, 101, 137-8
Sorel's cements see magnesium and gelation 49-50, 72, 77-9, 83-4
oxychloride cement, zinc oxychloride hydration shells 50, 74-7
cement and hydrides 33-4
splint bandage 2, 91, 117, 168 hydrosphere 32
SrO polyacrylate cement 102 ion binding effect 73-4
sulphur impregnation of cements 297 structure 34—40: anomalous density
sustained release devices 3, 4, 157-8, 39; bond angles 32; bond lengths
222, 304 32; constitution 31-3; D-structure
syringate cements see 37-40; density and 7 3 ^ ; hydrogen
EBA-methoxyhydroxybenzoate bonding 38; hydronium ions 44;
cements I-structure 3 7 ^ 0 ; intrinsic 74;
nucleophilic attack 39; O-H bond
tannic acid cements 6, 308, 315 energy 33; orientated 73;
tartaric acid cements 6, 308, 315 refractive index 73-4; translational
tartaric acid in glass polyalkenoate motion 35; translational energy
cement 132—5 37; V-structure 37-8; see also ice
temperature measurements 147, 380-1 as a solvent 30, 40-8: dissolution of
tensile strength 370 salts 41; dissolution of polymers
titrimetric methods 59, 311 45; hydrophobic interactions 40-1;
tobermorite gel 140 solvation energies 41;
tooth material see dentine, enamel thermodynamics 40-1
trace element release 3, 4, 156-8, 222, working time 375
304
transition temperatures 130 X-ray diffraction (XRD) spectroscopy
translucency see optical properties 9-10, 33, 35, 47, 51, 105, 125-6, 130,
transmission electron microscopy 145 198, 202-3, 208-9, 224-31, 250,
tricarballylic acid (1,2,3- 283-6, 293, 323, 359, 367-8
propanetricarboxylic acid) cements
6,315 Y 2 O 3 cements 102, 312

396
Index

zinc chloride solutions 2, 6, 283-9 205-6; powder (ZnO) 205-6, see


zinc oxide 1 0 3 ^ , 205-6, 321-2, 329-31 also zinc oxide
active forms 321-2, 328-31 history 204-5
deactivation 205-6 phases: ZnO-P 2 O 5 -H 2 O 199, 207-9
defect structures 329 properties 214: biological 219;
densification 1 0 3 ^ , 205-6 erosion 216-7; exotherm 207;
ignition 103-4, 205-6, 329 factors affecting 218-9; film
magnesium oxide in 205-6 thickness 215; fluidity 215;
mineralizers 206 mechanical 215-6; phosphoric
preparation by thermal decomposition acid:water effect 218;
salts 328-9 powder: liquid effect 219;
preparation by zinc oxidation 328 preparation 214-5; setting
sintering 206, 329 contraction 215; setting time
water in 328-9,331 214-5, 219; strength 215; working
zinc oxide chelate cements 318-20, see time 215
also EBA cements, EBA divanillate setting reaction 205, 207-12:
cements, aluminium role 205, 207, 209-12;
EBA-methoxyhydroxybenzoate crystallite formation 205, 208-12;
cements, zinc oxide eugenol cements exotherm 207; hydration 211-2;
zinc oxide eugenol (ZOE) cements infrared spectroscopy 210; pH
321-37, 381 changes 207; phosphoric acid
applications 320 concentration effect 207; XRD
composition 321-2, 336: abietic acid 209-10
(rosin) 322, 334; accelerators structure 99-101,212-3:
322-3; antimicrobial agents in 322, microstructure 212-3; molecular
335; impression paste 335; 99-101
modified materials 336; reinforced zinc phosphates 199-200
336-7 zinc polyacrylate cement see zinc
history 320-1, 335 polycarboxylate cement
properties 333-7: biological 334-5; zinc polyalkenoate cement see zinc
creep 333; hydrolysis 332—4; polycarboxylate cement
mechanical 333, 336-7; setting zinc polycarboxylate cement 31, 45, 56,
contraction 333; setting time 90-1, 93, 103-116, 362, 366, 372,
328-9; zinc oxide type, effect 380-1
328-31 applications 103
setting reaction 322-31: electrical composition 103—4: liquids 103—4,
conductance 325-6; permittivity see also poly(alkenoic acid)s;
325-6; water role 324-9, 331; zinc modified materials 113; poly-
oxide type, effect 328-31 (alkenoic acid) liquid 103-4, see
structure 331-2 also poly(alkenoic acid); powder
zinc oxychloride cements 2, 31, 51, 283, (ZnO) 103-4, see also zinc oxide;
285-90 reinforcing fillers 113; stannous
history 2,283,285-6 fluoride 107
phases 286-90: ZnO-HCl-H 2 O history 103
290; ZnO-ZnCl 2 -H 2 O 51, properties 94-7, 106-18: adhesion
286-90 94-7, 111-3; biological 112;
setting 287-8 erosion 109-11; setting 106-8;
structure: thermogravimetry 288—9; mechanical 107-9; metal fluoride
XRD 286-7 effect 108
zinc oxysulphate cements 6 setting reaction 98-9, 105-6:
zinc phosphate cement 204-21, 381 electrical conductivity and
applications 204, 214 permittivity 366-7; infrared
composition 205-6: aluminium role spectroscopy 105, 361, 365
205, 207, 209-12; fluoride-containing structure 105-6: hydration states
220; modified 220; liquid 207, 105-6; microstructure 105-6;
see also phosphoric acid; MgO in molecular structure 99-101, 105

397
Index
zinc selenate solution 6 cements, zinc phosphate cement, zinc
zinc sulphate solution 6 polycarboxylate cement
ZnO cements 102, 201-2, 204, 311-2, ZnO-P 2 O 5 -H 2 O phases 199, 207-9
see also EBA cements, zinc oxide ZOE cements see zinc oxide eugenol
chelate cements, zinc oxide eugenol cements

398

You might also like