Vous êtes sur la page 1sur 8

The ofcial journal of the Australian Dental Association

Australian Dental Journal


Australian Dental Journal 2011; 56:(1 Suppl): 2330 doi: 10.1111/j.1834-7819.2010.01293.x

Glass-ionomer cement restorative materials: a sticky subject?


SK Sidhu*
*Queen Mary University of London, Barts and The London School of Medicine and Dentistry, Institute of Dentistry, London, United Kingdom.

ABSTRACT
Glass-ionomer cement (GIC) materials have been in clinical use since their inception 40 years ago. They have undergone several permutations to yield different categories of these materials. Although all GICs share the same generic properties, subtle differences between commercial products may occur. They have a wide range of uses such as lining, bonding, sealing, luting or restoring a tooth. In general, GICs are useful for reasons of adhesion to tooth structure, uoride release and being tooth-coloured although their sensitivity to moisture, inherent opacity, long-term wear and strength are not as adequate as desired. They are useful in situations where they are not disadvantaged by their comparatively lower physical properties, such as where there is adequate remaining tooth structure to support the material and where they are not subject to heavy occlusal loading. The last decade has seen the use of these materials being extended. However, they are likely to retain their specic niches of clinical application.
Keywords: Glass-ionomers, resin-modified glass-ionomers, high viscosity glass ionomers, clinical uses. Abbreviations and acronyms: ART = Atraumatic Restorative Treatment; HEMA = hydroxylmethacrylate; MMGIC = metal-modified glass-ionomer cement; RMGIC = resin-modified glass-ionomer cement.

INTRODUCTION Glass-ionomer cement (GIC) materials were invented four decades ago by Wilson and Kent in 1969 at the Laboratory of the Government Chemist in London, United Kingdom.1 These materials form part of the contemporary armamentarium for restorative dentistry largely due to their adhesive, tooth-coloured and uoride-leaching properties. They are used today in a variety of clinical situations as restorative, lining, luting and sealing materials; no other restorative material has such wide applications. After 40 years of practical use, they are reasonably well understood and researched. This paper collates some of the information relevant to our understanding and clinical use of these materials and will focus on the restorative applications. Some evidence is theoretical or scientic, while the rest may be anecdotal, amassed over a long period of use and observation. Basic composition and variants In simple terms, glass-ionomers are derived from organic acids and a glass component, and are referred to as acid-base reaction cements. The acid is generally an aqueous polymeric acid and the glass is usually a uoroaluminosilicate, although other non-uoride
2011 Australian Dental Association

glasses have been used. The additives may differ in products available. They are complex materials and no two commercial systems are chemically or mechanically identical. The setting reaction involves a conventional acid-base reaction initiated on mixing the components. The maturation of the resultant cement is relatively slow, with the rst 24 hours being important to maintain the water balance within the system. The standard protocol for the earlier GICs was to delay the nishing and polishing stages by at least 24 hours. The emergence of newer improved, rapidly maturing materials for posterior use has necessitated a revision of the earlier techniques. However, chemical maturation may not be achieved in some cements for 24 hours or more, and hence they should be allowed to mature completely under a protective coating to optimize their benecial properties.2 It is difcult to apply a blanket rule applicable to all GICs, especially in the absence of well-controlled data, therefore it may be advisable to delay the polishing process if at all possible. Finishing and polishing is often advised with rotary instruments such as exible discs, preferably lubricated with a medium such as petroleum jelly or bonding resin. It is presumed that this will prevent the restoration from dehydrating and maintain the all-important water balance in the system as they are water-based materials.
23

SK Sidhu Developments in the glass-ionomer category have led to the introduction of newer materials, resulting in considerable confusion as to what constitutes a true glass-ionomer cement material. A material that contains the components of a GIC alone is not necessarily one. A true GIC is a two-part system characterized by an acid-base reaction critical to its cure, and continuing uoride release. In an attempt to improve the physical characteristics of the original materials, additions of metal powders were introduced. The rst such suggestion was a silver alloy and GIC admixture, with the subsequent emergence of materials incorporating ne metal particles sintered onto the cement-forming glass to form ceramic-metal materials or cermets. These cements with alloy additives, whether or not fused to the glass, collectively are better referred to as metal-modied glass-ionomer cements (MMGICs). They are often described as packable or high viscosity or high powder:liquid ratio GICs, and commercial examples include Ketac-Silver (3M ESPE, Seefeld, Germany), Hi Dense (Shofu Inc, Kyoto, Japan) and Miracle Mix (GC Corp, Tokyo, Japan). They are also sometimes referred to loosely as metal-reinforced GICs, but this term has been viewed as somewhat of a misnomer. Although these cements are said to be superior in terms of physical properties compared to the conventional GICs, the literature does not always support this view. In general, their strength has been found to be less than satisfactory compared to other posterior restorative materials for use in high stress-bearing areas3 and they may not even be better than the conventional high viscosity GIC materials.4 They do not appear to perform well clinically in posterior teeth.5 Another such development was the introduction of the resin-modied glass-ionomer cement (RMGIC) materials patented in the late 1980s.6,7 This innovation was an attempt to help overcome the problems traditionally associated with the conventional materials, i.e. moisture sensitivity and low physical properties (particularly their early mechanical strength). They were perceived to be an improvement over the original materials while still maintaining the clinical advantages of the traditional GICs, such as adhesion and uoride release, offering some measure of protection against caries. In essence, the RMGICs are glass-ionomer cements with the incorporation of a small quantity of monomers as well as initiators involved in the polymerization reaction. The fundamental acid-base curing reaction is supplemented by a second polymerization reaction. This latter process may be initiated by light, as in the light-cured RMGICs, which have the ability to set without light activation although more slowly. Other versions of RMGICs such as luting cements are not dependent on light activation. In their simplest forms, they are GICs with the addition of a small
24

quantity (4.5 to 6%) of resin components, such as hydroxyethylmethacrylate (HEMA) or Bis-GMA, although the actual formulation may vary. Some of the water component of the conventional GIC is replaced by a water HEMA mixture. The initial set of these materials is due to the formation of a polymerization matrix while the acid-base reaction hardens and strengthens the matrix formed. The rst RMGIC to be developed and marketed was a lining cement (Vitrebond, 3M Dental, St Paul, MN, USA) but other versions were subsequently introduced. Their command set facility made them popular as liners and bases, and subsequently as restorative materials. Rapid acceptance of these materials by the dental profession saw subsequent similar materials appearing in the marketplace which were variations of the same theme. However, these latter materials are not all considered true GICs as they may not fulll the requirements of a GIC of having a typical acid-based glass-ionomer reaction, unlike the RMGICs which are considered as GICs. By denition, the RMGICs contain a basic ion-leachable glass, a water-soluble polymeric acid, organic monomer s and an initiator system.8 The material must be capable of auto-setting due to the acid-base reaction, even if designed to be light-cured. The arrival of the RMGICs did not make the conventional GICs obsolete; on the contrary, these latter materials were themselves undergoing exciting developments of their own. The need for a suitable material that could be used in conjunction with hand instrumentation in remote communities where there is no access to rotary instrumentation and dental care, was commissioned by the World Health Organization and this resulted in the development of the high viscosity GICs. A new restorative technique emerged, the Atraumatic Restorative Treatment (ART) technique,9 whereby decalcied tooth tissue can be removed using hand instruments only, followed by restoration with an adhesive auto-setting GIC. These material properties are essential to allow them to be used in any environment in any part of the world. The high viscosity GICs in their powder-liquid hand-mixed version are ideal for this as other restorative materials require some use of electrically-driven equipment. The earliest high viscosity material developed for this purpose was Fuji IX (GC Corp, Tokyo, Japan). The success of this formulation prompted other manufacturers to develop faster-setting GICs with better physical properties, not just for rural dentistry but for wider use. Today, these latter materials earn their pride of place in contemporary restorative dentistry and not only include Fuji IX but also Ketac Molar (3M ESPE) and ChemFil Molar (Dentsply DeTrey GmbH, Konstanz, Germany). The changes in the formulation over the traditional materials included a reduction in the size of the glass particles in the matrix offering improved
2011 Australian Dental Association

Glass-ionomer cement restorative materials physical properties and stiffer syringeable materials which allow some degree of packability. Application of ultrasonic excitation, which appears to increase the strength of conventional high viscosity GICs, can be used to set on command and may improve their survival rate clinically.10 In summary, the broad categories of GICs available today are the conventional, the resin-modied and the metal-modied GICs. The latter two categories were developed in an attempt to overcome the problems of moisture sensitivity and low mechanical properties associated with the conventional materials but at the same time to retain some of their clinical advantages. Improvements within the conventional GICs have produced a subgroup of high viscosity GICs. GICs are presented as hand-mixed powder:liquid materials or capsulated versions, and more recently paste systems. Hand-mixed materials require dispensing of the powder which should be mixed with the liquid in the correct powder:liquid ratio as specied by the individual manufacturer. Optimal physical properties are obtained when the powder:liquid ratio as recommended by the individual manufacturer are followed. In general, for the conventional GICs, increasing the powder:liquid ratio improves the strength but may be difcult to mix by hand. Conversely, lowering the ratio from that recommended by the manufacturer would impair the cements properties.11 Increasing the powder:liquid ratio results in an increase in viscosity and a drier mix which may reduce the ability of the material to effectively wet the substrate. This may have an effect on the bonding and ultimately the retention of the restoration.12 The use of capsulated materials reduces the possibility of errors in the powder:liquid ratio while mixing, as they are supplied in capsules containing premeasured amounts of powder and liquid separated by the liquid which is encased in a pillow. On activation of the capsule, this pillow is ruptured and the liquid is expelled through a narrow orice onto the powder. The activated capsule is then mixed mechanically according to the manufacturers instructions. It is usually advised that the rst few millimetres extruded from each capsule are discarded before usage. The angled nozzle of the capsules act as a syringe to facilitate direct placement into a cavity. The latest development in the dispensing arena of GICs is a paste-paste RMGIC, Ketac Nano (3M ESPE) introduced in 2007. This system uses a specially designed cartridge and material dispenser or doublebarrelled clicker that delivers dened portions of two pastes to be mixed by spatulation. This nano-ionomer material contains an ultra-ne glass powder designed for this purpose. The exact properties of this material will undoubtedly be known in the fullness of time.
2011 Australian Dental Association

General properties The features of clinical relevance of a restorative material, especially a GIC, include adhesive properties, marginal adaptation, biocompatibility, moisture sensitivity, uoride release, strength, and wear. In general, the GICs are often thought to suffer from drawbacks such as poorer physical and aesthetic properties when compared to other restorative materials. In addition, they are frequently said to be technique-sensitive due to their moisture sensitivity; however, this is not unique to the GICs when adhesives and resin composites are considered. The RMGICs have some advantages over the conventional materials: greater working time; command set on application of the relevant light source; earlier nishing; aesthetics closer to resin-based materials; better strength characteristics.13 However, they have not been proven to be superior to the conventional materials with regard to adhesion, resistance to water uptake, uoride release, solubility and biocompatibility.14,15 They appear to perform well in clinical trials based on retention, secondary caries and absence of postoperative sensitivity, but this is not necessarily true of their marginal characteristics, surface properties and colour stability.16 In addition, access for placement of the light-curing tip in some parts of the mouth may preclude use of these materials. Adhesive properties The adhesion of GICs appears to be via mechanical interlocking of cement in dentinal tubules and the development of an ion-exchange layer adjacent to the dentine.17 In practical terms, what does all this mean? This should translate into good retention within cavities and the added benets of prevention of secondary caries. Is there enough evidence to support this idea? The evidence for adhesion and retentive properties is typically conducted on non-carious cervical lesions, both in the laboratory and clinically, which may appear fundamentally logical but clinically irrelevant. The performance in this type of cavity may not necessarily be extrapolated to other clinical situations and carious surfaces. Nevertheless, the evidence in this type of clinical scenario seems to point towards effective bonding of GICs to tooth structure.18 Conditioning the dentine surface of the cavity prior to placement of the GIC is essential for encouraging adhesion of the cement to the substrate. The bond strength of most GICs to dentine has been shown to be signicantly higher if the dentine is pretreated.19 Regarding the bond strength of GICs to other restorative materials, overlaying with resin composite may be desirable and even necessary in some cases such as in posterior restorations. The shear bond strength of GICs to composites is said to be sufcient20 and that of the
25

SK Sidhu RMGICs to resin composite may be signicantly higher than the conventional GICs.21 This may be attributed to the formation of a catalyst-rich air-inhibited surface layer on the RMGIC, facilitating bonding by polymerizing to the resin composite above.13 Marginal adaptation and microleakage Marginal leakage can occur due to dimensional changes and lack of adaptation of the restoration to the cavity walls. Although the slower-setting conventional materials are thought to permit stress relief within the restoration, the RMGICs may exhibit more rapid setting contraction through the polymerization of the polymer component. When the RMGIC is extended to enamel margins, there may be considerable risk of enamel fracture.22 However, this is not necessarily borne out by research as the RMGICs appear to display substantially better adaptation to dentine than the conventional materials.23 It is possible that a propensity for water absorption by the HEMA content compensates for the initial setting contraction in the RMGICs. Biocompatibility: biological and pulpal effects In a review of the literature on biocompatibility of GICs, the authors concluded that most aspects of the GICs allow a reasonable margin of tolerance from a biocompatible standpoint.15 However, they emphasize that the biologic properties are product-specic. While the initial pulp reactions to some products appear to resolve in time, especially if there is a dentine barrier, the long-term effects of direct application of GIC to pulp tissue are largely unknown. Glass-ionomers have been much maligned in the past due to the fear of elution of ions such as aluminium which may have the potential for profound biological effects. Whether this actually occurs is highly debatable, as there would be far more reports of adverse reactions if this were the case. A recent paper reviewed the role of aluminium in GICs and concluded that this ion is leached in varying degrees; however, it is largely excreted and poses a negligible health hazard.24 Concern has been previously raised regarding the biocompatibility of the RMGICs in particular as they contain unsaturated groups.15 They cannot be considered biocompatible to the same extent as conventional GICs.25 Moreover, the RMGICs are considered to produce a polymerization exotherm and greater temperature rises than conventional GICs;26 however, this has not been shown to be a clinical issue. Moisture sensitivity The complex nature of the setting reaction of GICs is often blamed for the moisture sensitivity, especially in
26

the early stages and beyond. The need to maintain the water balance in GICs, particularly in the early phases of their maturity, has led to the recommendation that the surface of a newly placed glass-ionomer be protected from water loss and, equally importantly, from water gain. Various materials have been used including copal varnish, light-cured bonding resins, petroleum jelly and cocoa butter. Early protection with light-cured resins or commercial varnishes reduces their vulnerability to dissolution and deterioration of physical properties. The effect of a waterproof coating on the uoride release by a GIC system has been debated;2 however, it is doubtful if the resins used as coating agents remain on the surface long enough to prevent this, while their presence in the short-term is benecial in protection. Another advantage of using a coating agent is the potential to ll surface voids and defects, reducing the uptake of stains from food and drinks. The RMGICs are thought to be less prone to moisture sensitivity due to the resin network reducing the diffusion of water into the cement and hence protecting the cement from dissolution by early contact with water. However, this may not be the case as they appear to be susceptible to dehydration27 as well as having the potential to take up water from the environment;28 this may ultimately affect their properties such as strength and colour stability and could affect the bonding interface.29 One of the greatest advantages of the RMGICs over their conventional counterparts is earlier nishing and polishing; however, it has been suggested that the water balance in these materials is just as critical30 which may have implications for nishing soon after initial set. Fluoride release: the role of uoride and anticariogenicity The uoride-releasing properties of GICs is probably one of their greatest assets. It is assumed that the GICs have a caries-inhibitory effect which is due to their long-term and sustained uoride release. Fluoride is used as a ux during the manufacturing process of the glass powder and is not a matrix-forming species; this uoride is available for release from the set cement to inuence the immediate surrounding tooth tissue as well as any adjacent surface. However, the inherent uoride is depleted fairly quickly within the rst few months. Nevertheless, the cement has the capacity to take up more uoride from the ambient environment, depending on the concentration gradient. It is thought that this may continue for the life of the restoration and hence the GIC acts as a uoride reservoir. This rechargeability31 is particularly advantageous where there is a high caries rate. Hence, they are known as bioactive or smart materials as they are not passive,
2011 Australian Dental Association

Glass-ionomer cement restorative materials but instead react to the environment. The uoride release of the RMGICs is just as good, although the amount and rate of release by different restorative products may vary.32 The critical level of uoride released by a material for effective caries inhibition has not yet been established, often working on the premise that the more uoride released over a longer period, the better. The GICs are thought to be rechargeable uoride systems with their ability to be recharged by uoride exposure in solution, hence acting as reservoirs of uoride which may be released into saliva, plaque and dental tissues,33 over time and given the right conditions.31 It is thought that the uoride release from GICs is responsible for the bacterial inhibition associated with these materials (especially against Streptococcus mutans), although this has been recently disputed.34 Hence, the uoride release may not be the only mechanism of antibacterial action as other components (such as aluminium and strontium) have also been implicated.35,36 Perhaps this antibacterial property could account for the lower incidence (anecdotal) of gingival inammation immediately adjacent to cervical GIC restorations. The MMGICS appear to have a lower antibacterial effect compared to the RMGICs.37 In spite of laboratory and anecdotal evidence that GICs show cariostatic properties, this is often not borne out in clinical studies on the incidence of secondary caries.38,39 Clinical data are conicting as to whether GICs have an effect on secondary caries. This may be due to the methods for clinical evaluation of GICs which need to be appraised38 and does not explain the anecdotal evidence to support the effect of GICs on the immediate surroundings as well as adjacent surfaces. Is this a serendipitous outcome? Clinical uses of GICs The common attributes of adhesion and uoride release make these materials useful for a variety of clinical situations. While other restorative materials have enjoyed a largely constant number of indications for their use, the list of indications for GICs has steadily grown. Notwithstanding, a common indication for GICs as restorative materials is the non-carious cervical lesion, where the need for cavity preparation and mechanical retention is reduced (Figs 1a and 1b). These materials have particularly good potential in root caries, not uncommon in an ageing population, due to their adhesive qualities and uoride release (Figs 2a and 2b). Such lesions do not lend themselves to ideal cavity preparations, particularly those that present as encircling defects around the cervical margin.
2011 Australian Dental Association

Fig 1. RMGIC used to restore the cervical lesion in tooth 21. (Courtesy of Professor Martin Tyas, Australia.)

Fig 2. Carious cervical lesions before and after restoration with RMGIC. (Courtesy of Dr Sueo Saito, Japan.)
27

SK Sidhu Glass-ionomers have been traditionally used in non or minimal load-bearing situations. Their use in posterior teeth has been limited by their physical properties. However, the need for a tooth-coloured material with relatively easy handling properties prompted the development of the high viscosity GICs. The main reasons for this was the need to nd a replacement for the traditional amalgam as well as the need for the use of a material in what is now the well-established ART technique (see above) in areas where there is no access to rotary instrumentation and dental care, such as in rural communities and many developing countries, and is predicated on the use of a high viscosity GIC. Shortterm results show that this approach has increased the ratio of restorations to extractions in these populations.40 The question remains regarding the extent to which this technique can be extended to the wider population and is an effective public health measure. The high viscosity GICs may be selected for occlusal and approximal carious lesions in primary and permanent posterior teeth, provided that they can be conservatively prepared and restored within the limits of the relevant occlusion and not be subjected to heavy occlusal load. Glass-ionomers have not performed well in clinical studies on longevity of posterior restorations under load.41,42 Thus, in high load areas overlaying them with a more durable material such as resin composite would make clinical sense. In the cervical lining (open sandwich) technique, it is recommended that the glass-ionomer should be kept well below the contact area otherwise it may dissolve (Figs 3a and 3b); progressive loss of material in approximal areas, just below contact areas, was commonly observed in a sixyear study using a high viscosity GIC in approximal cavities, leading the authors to conclude that the presence of approximal contacts promotes disintegration of the GIC.43 What the GICs may be suited for are microcavity designs for conservative treatment of approximal lesions, which are essentially an occlusal approach to an approximal lesion. At present, the GICs are not designed to be placed under direct occlusal load so they should be overlaid with another more durable material such as a resin composite. The effectiveness of the tunnel approach to the traditional approximal preparation was reviewed and it was concluded that tunnel restorations restored with GICs are technically decient and have a limited life-span;39 they were not recommended as alternative preparations for approximal lesions. Glass-ionomers can be considered only as long-term provisional restorations in stress-bearing posterior cavities.44 Caution must be exercised in the use of any GIC in core build-ups if there is little coronal tooth structure present to support a core build-up,45,46 as their physical properties do not lend themselves to this; they may be used as space-llers until further improvements are made. Although the MMGICs appeared to be promising as core build-up materials,47 their physical properties and lack of adhesion to tooth structure make them less than ideal for this purpose.48 Also, it is advised against routinely preparing a tooth whose core includes a GIC (even if it is an RMGIC) at the same visit as placement, as the level of maturity would not have been fully established. In an era of minimum intervention dentistry, rather than replacement dentistry, perhaps more consideration could be given to repairs of open margins, marginal defects around castings, etc. in the absence of a compelling need to otherwise replace or remake a restoration. This should not, however, be a compromise accepted in the presence of inadequate dentistry. Glassionomers may be very useful in this indication. They can also be used to repair perforations (Figs 4a and 4b).

(a)

(b)

(a)

(b)

Fig 3. (a) Radiograph showing approximal dissolution of GIC where it had been extended to immediately below the contact area. (b) Ideal outline of the extent of the glass-ionomer when restoring an approximal lesion. (Courtesy of Professor Martin Tyas, Australia.)
28

Fig 4. (a) Resorptive defect in tooth 11. (b) Defect repaired with GIC. (Courtesy of Dr Justin Barnes, United Kingdom.)
2011 Australian Dental Association

Glass-ionomer cement restorative materials In summary, the GICs may be useful as liners and bases, for anterior approximal restorations, cervical restorations (both carious and non-carious), core buildups where there is sufcient remaining tooth structure, occlusal and small approximal restorations in deciduous teeth, microcavities, internal and tunnel preparations, temporary repairs of fractured teeth, temporary repairs of defective crown margins, retrograde root llings and root perforation repairs. CONCLUSIONS GICs are mainstream restorative materials that are bioactive and have a wide range of uses such as lining, bonding, sealing, luting or restoring a tooth. It is important to recognize that although the GICs share the same generic properties, subtle differences between commercial products may occur. The RMGICs appear to have properties intermediate to the conventional GICs and resin composites and are often considered a hybrid of the two materials. In general, GICs are useful for reasons of adhesion to tooth structure, uoride leaching and being toothcoloured, although their sensitivity to moisture, inherent opacity, long-term wear and strength are not as adequate as desired. They are useful in situations where they are not disadvantaged by their comparatively low physical properties, such as where there is adequate remaining tooth structure to support the material and where they are not subject to heavy occlusal loading. The last decade has seen the use of these materials being extended. However, they are likely to retain their specic niches of clinical application. REFERENCES
1. Wilson AD, Kent BE. Surgical cement. British Patent No. 1316129, filed in 1969, specification published in 1973. 2. Hattab FN, Amin WM. Fluoride release from glass ionomer restorative materials and the effects of surface coating. Biomater 2001;22:14491458. 3. Naasan MA, Watson TF. Conventional glass ionomers as posterior restorations. A status report for the American Journal of Dentistry. Am J Dent 1998;11:3645. 4. Yap AU, Cheang PH, Chay PL. Mechanical properties of two restorative reinforced glass-ionomer cements. J Oral Rehabil 2002;29:682688. 5. Holst A. A 3-year clinical evaluation of Ketac-Silver restorations in primary molars. Swed Dent J 1996;20:209214. 6. Antonucci JM, McKinney JE, Stansbury JW. Resin modied glass-ionomer dental cement. US Patent 7160856, 1988. 7. Mitra SB. Photocurable ionomer cement systems. European Patent Application 323120, 1988. 8. McLean JW, Nicholson JW, Wilson AD. Proposed nomenclature for glass-ionomer dental cements and related materials. Quintessence Int 1994;25:587589. 9. Frencken JE, Pilot T, Sangpaisan Y, Phantumvanit P. Atraumatic Restorative Treatment (ART): rationale, technique and development. J Public Health Dent 1996;56:135140.
2011 Australian Dental Association

10. Kleverlaan CJ, van Duinen RN, Feilzer AJ. Mechanical properties of glass ionomer cements affected by curing methods. Dent Mater 2004;20:4550. 11. Billington RW, Williams JA, Pearson GJ. Variation in powder liquid ratio of a restorative glass-ionomer cement used in dental practice. Br Dent J 1990 22;169:164167. 12. Wilder AD, Boghosian AA, Bayne SC, Heymann HO, Sturdevant JR, Roberson TM. Effect of powder liquid ratio on the clinical and laboratory performance of resin-modied glass-ionomers. J Dent 1998;26:369377. 13. Burgess JO, Barghi N, Chan DC, Hummert T. A comparative study of three glass ionomer base materials. Am J Dent 1993;6:137141. 14. Sidhu SK, Watson TF. Resin-modied glass ionomer materials. A status report for the American Journal of Dentistry. Am J Dent 1995;8:5967. 15. Sidhu SK, Schmalz G. The biocompatibility of glass-ionomer cement materials. A status report for the American Journal of Dentistry. Am J Dent 2001;14:387396. 16. Sidhu SK. Clinical evaluations of resin-modied glass-ionomer restorations. Dent Mater 2010;26:712. 17. Lin A, McIntyre NS, Davidson RD. Studies on the adhesion of glass-ionomer cements to dentin. J Dent Res 1992;71:1836 1841. 18. Peumans M, Kanumilli P, De Munck J, Van Landuyt K, Lambrechts P, Van Meerbeek B. Clinical effectiveness of contemporary adhesives: a systematic review of current clinical trials. Dent Mater 2005;21:864881. 19. Peutzfeldt A. Compomers and glass ionomers: bond strength to dentin and mechanical properties. Am J Dent 1996;9:259263. 20. Tyas MJ, Toohey A, Clark J. Clinical evaluation of the bond between composite resin and etched glass ionomer cement. Aust Dent J 1989;34:14. 21. Kerby RE, Knobloch L. The relative shear bond strength of visible light-curing and chemically curing glass-ionomer cement to composite resin. Quintessence Int 1992;23:641644. 22. Watson TF. A confocal microscopic study of some factors affecting the adaptation of a light-cured glass ionomer to tooth tissue. J Dent Res 1990;69:15311538. 23. Sidhu SK. Marginal contraction gap formation of light-cured glass ionomers. Am J Dent 1994;7:115118. 24. Nicholson JW, Czarnecka B. Review paper. Role of aluminum in glass-ionomer dental cements and its biological effects. J Biomater Appl 2009;24:293308. 25. Nicholson JW, Czarnecka B. The biocompatibility of resinmodied glass-ionomer cements for dentistry. Dent Mater 2008;24:17021708. 26. Kanchanavasita W, Pearson GJ, Anstice HM. Temperature rise in ion-leachable cements during setting reaction. Biomaterials 1995;16:12611265. 27. Sidhu SK, Sherriff M, Watson TF. The effects of maturity and dehydration shrinkage on resin-modied glass-ionomer restorations. J Dent Res 1997;76:14951501. 28. Small IC, Watson TF, Chadwick AV, Sidhu SK. Water sorption in resin-modied glass-ionomer cements: an in vitro comparison with other materials. Biomater 1998;19:545550. 29. Sidhu SK, Watson TF. Interfacial characteristics of resin-modied glass-ionomer materials:a study on uid permeability using confocal uorescence microscopy. J Dent Res 1998;77:1749 1759. 30. Sidhu SK, Pilecki P, Sherriff M, Watson TF. Crack closure on rehydration of glass-ionomer materials. Eur J Oral Sci 2004;112:465469. 31. Yan Z, Sidhu SK, Mahmoud GA, Carrick TE, McCabe JF. Effects of temperature on the uoride release and recharging ability of glass ionomers. Oper Dent 2007;32:138143.
29

SK Sidhu
32. Chan WD, Yang L, Wan W, Rizkalla AS. Fluoride release from dental cements and composites: a mechanistic study. Dent Mater 2006;22:366373. 33. Wiegand A, Buchalla W, Attin T. Review on uoride-releasing restorative materialsuoride release and uptake characteristics, antibacterial activity and inuence on caries formation. Dent Mater 2007;23:343362. 34. Poggio C, Arciola CR, Rosti F, Scribante A, Saino E, Visai L. Adhesion of Streptococcus mutans to different restorative materials. Int J Artif Organs 2009;32:671677. 35. Hayacibara MF, Rosa OP, Koo H, Torres SA, Costa B, Cury JA. Effects of uoride and aluminum from ionomeric materials on S. mutans biolm. J Dent Res 2003;82:267271. 36. Ngo HC, Mount G, McIntyre J, Tuisuva J, Von Doussa RJ. Chemical exchange between glass-ionomer restorations and residual carious dentine in permanent molars:an in vivo study. J Dent 2006;34:608613. 37. Friedl KH, Schmalz G, Hiller KA, Shams M. Resin-modied glass ionomer cements: uoride release and inuence on Streptococcus mutans growth. Eur J Oral Sci 1997;105:8185. 38. Randall RC, Wilson NH. Glass-ionomer restoratives: a systematic review of a secondary caries treatment effect. J Dent Res 1999;78:628637. 39. Wiegand A, Attin T. Treatment of proximal caries lesions by tunnel restorations. Dent Mater 2007;23:14611467. 40. Frencken JE. The ART approach using glass-ionomers in relation to global oral health care. Dent Mater 2010;26:16. 41. Kilpatrick NM. Durability of restorations in primary molars. J Dent 1993;21:6773. 42. Welbury RR, Walls AW, Murray JJ, McCabe JF. The 5-year results of a clinical trial comparing a glass polyalkenoate (ionomer) cement restoration with an amalgam restoration. Br Dent J 1991;170:177181. 43. Scholtanus JD, Huysmans MC. Clinical failure of class-II restorations of a highly viscous glass-ionomer material over a 6-year period: a retrospective study. J Dent 2007;35:156162. 44. Manhart J, Garc a-Godoy F, Hickel R. Direct posterior restorations: clinical results and new developments. Dent Clin North Am 2002;46:303339. 45. Cho GC, Kaneko LM, Donovan TE, White SN. Diametral and compressive strength of dental core materials. J Prosthet Dent 1999;82:272276. 46. Bonilla ED, Mardirossian G, Caputo AA. Fracture toughness of various core build-up materials. J Prosthodont 2000;9:1418. 47. Burke FJ, Watts DC. Cermetan ideal core material for posterior teeth? Dent Update 1990;17:364370. 48. Gu S, Rasimick BJ, Deutsch AS, Musikant BL. In vitro evaluation of ve core materials. J Prosthodont 2007;16:2530.

Address for correspondence: Dr Sharanbir K Sidhu Clinical Senior Lecturer Honorary Consultant Centre for Adult Oral Health Queen Mary University of London, Barts and The London School of Medicine and Dentistry Institute of Dentistry Turner Street, London E1 2AD United Kingdom Email: s.k.sidhu@qmul.ac.uk

30

2011 Australian Dental Association

Vous aimerez peut-être aussi