Вы находитесь на странице: 1из 12

Biomaterials 19 (1998) 467 478

Development of glass-ionomer cement systems


Dennis C. Smith
Faculty of Dentistry, University of Toronto, 124 Edward Street, Toronto, ON M5G 1G6, Canada

Abstract In the 1960s the idea of positive physico-chemical adhesion with tooth substance resulted in the invention of polyacrylic acid-based cements, rst the zinc polycarboxylate and, subsequently, the glass-ionomer cements. These materials were shown to undergo specic adhesion with hydroxyapatite and proved to have properties satisfactory for a variety of clinical applications. The key properties of the glass-ionomer cementsuoride release over a prolonged period and specic adhesion to enamel and dentine coupled with aesthetic qualitiesare related to their characteristics as aqueous polyelectrolyte systems. In order to improve toughness, speed of setting and resistance to dehydration, hybrid materials in which some of the water content of the glass-ionomer system was replaced by water-soluble polymers or monomer systems capable of ambient polymerization were formulated in the late 1980s. These materials, which have been termed resin-modied glass-ionomer cements, involve, ideally, the formation of an interpenetrating polymer network combining the acidbase cross-linking reaction of the metal ionpolyacid with the cross-linking polymerization of the monomer system or additive action of the polymers. In the predominantly resin materials there is little polyelectrolyte character and it is controversial whether such materials should be categorized as glass-ionomer cement systems. The specic advantages of these materials over traditional glass-ionomer systems and over composite restorative systems remain to be fully documented. Studies of adsoption to hydroxyapatite of typical monomers using X-ray photoelectron spectroscopy (XPS) and time-of-ight secondary ion mass spectrometry (TOF SIMS) indicate that resistance to water displacement decreases as hydrophobicity increases. 1998 Published by Elsevier Science Ltd. All rights reserved Keywords: Glass-ionomer; Polyelectrolyte; Cement; Resin; Bonding

1. Introduction The development of amalgam, gold and procelain restorative materials in the rst half of the nineteenth century stimulated the development of dental cements as luting and lining materials and as more aesthetic restorative materials. The zinc oxychloride cement of Sorel (1855) was displaced by the zinc oxidephosphoric acid cement of Pierce in 1879 [1]. This oxyphosphate cement superseded the oxychloride and oxysulphate cements because of its lower irritation to the pulp and greater durability. The work of Ames [2] and Fleck [3] around the turn of the century established the modern-day zinc phosphate cement. At about the same time as Dr Pierce, Dr J. Foster Flagg originated the zinc oxide eugenol cement (1875) which soon became popular because of its

Tel.: (416)-979-4903, Ex. 4470; fax: (416)-979-4936.

obtundent eect. By the end of the century these cements were used to lute gold and porcelain crowns and inlays to the tooth in addition to serving as temporary lling materials and cavity bases. The use of cements became widespread following the popularization of a technique for casting gold restorations by Taggart in 1907. A more aesthetic lling material, the silicate cement, was developed by Thomas Fletcher in England in 1873 but the material did not become popular until Steenbock of Berlin introduced an improved version in 1904 [4]. Later in 1908 Shoenbeck [5] developed a material that contained uoride. The silicate cement was used principally for llings because of its translucence. Hybrid silico-phosphate cements were developed for cementing applications. Thus by the end of the rst quarter of this century these three basic types of cement, zinc phosphate, zinc oxideeugenol and silicate cement, were established for the bonding of inlays, crowns, posts, bridges and orthodontic bands on or in the tooth and as cavity linings,

0142-9612/98/$19.00 1998 Published by Elsevier Science Ltd. All rights reserved. PII S 0 1 4 2 - 9 6 1 2 ( 9 7 ) 0 0 1 2 6 - 9

468

D.C. Smith / Biomaterials 19 (1998) 467 478

bases and lling materials. Over the next 50 years although these materials underwent considerable technical improvement they remained, in principle, unchanged. The eect of variables such as powderliquid ratio and mixing conditions on properties such as strength and solubility became well understood and clinical usage was designed to optimize such physical parameters and to minimize biological eects such as pulpal response associated with acidity. Reasonable performance was obtained from these materials by specifying precise manipulative techniques though few studies of clinical durability were published. Not withstanding the rapid progress of restorative dentistry that was facilitated by the development of these cements, the need for new and improved materials has been long recognized. In 1902, Fleck [3], one of the pioneers of zinc phosphate cement, remarked, prophetically, that a new departure cement must have a new departure liquid as well as powder. The ideal cement will be found outside the phosphate class. However, research on new cements was not successful until more than 60 years later.

2. Resin cements: adhesion to the tooth The so-called self curing acrylic resin materials were developed in the 1940s by German chemists [6]. Reports [7] after World War II disclosed the use of tertiary amines as accelerators for redox polymerization with benzoyl peroxide. Subsequent development of this concept led to other systems [8] and a variety of acrylic lling materials in the 1950s. The major problem with these acrylic materials was polymerization shrinkage; poor colour stability, low modulus, high thermal expansion and lack of adhesion to the tooth were other disadvantages. The combination of shrinkage plus lack of adherence to the cavity walls resulted in leakage and bacterial penetration with a high incidence of dentine caries. These deciences led to various investigations directed towards the development of adhesive resin restorative materials. A major step forward in the late 1940s was achieved by Hagger [9] working for the Amalgamated Dental Company in London who invented a cavity seal for Sevriton lling material that contained glycerophosphoric acid dimethacrylate and methacrylic acid and which could be polymerized rapidly using a sulphinic acid initiator system. McLean and Kramer [10, 11] investigated this system in the early 1950s and were able to demonstrate that the glycerophosphoric acid dimethacrylate increased adhesion by penetrating the dentine surface forming what is now called the hybrid zone. This was demonstrated by an altered staining reaction of the dentine surface. Similar results were obtained in 1956 by Buonocore et al. [12] who also demonstrated that an adhesive resin containing

the Sevriton cavity seal showed increased bonding to dentine that was rst pretreated with 7% HCl. All these bonds showed signicant degradation on storage in water for 6 months. Other dental manufacturers developed similar acidic cavity primers containing methacrylic acid and acidic amides [13]. Other types of phosphate monomers were developed in Japan in the 1970s [14]. The signicant iatrogenic problems associated with the use of acrylic lling materials stimulated developments on two fronts; rstly, exploration of the possible research approaches to adhesive dental restorative materials and secondly, the evolution of lled resin materials (ceramicresin composites). The reduction of polymerization shrinkage and improvement in physical properties through the incorporation of llers had received attention earlier in attempts to improve acrylic denture base materials. Similar acrylic lling materials containing aluminosilicate glass llers were formulated in the 1950s [8]. Improved properties were obtained with these materials when the silicate glass particles were precoated with polymer and/or primed with silane [15]. Such materials became available commercially and one material (TD71) was reported on favourably by McLean in 1962 [16]. Fluoride release from these materials was small but may have contributed to the clean surface and margins and longevity observed in some long-term cases (Fig. 1). These materials were dicult to handle and were superseded by the composites based on BIS-GMA originated by Bowen [20]. Nevertheless, the concept of a material having the good properties of a silicate cement but utilizing a matrix more resistant to dissolution is evident in this early development. In 1960, the NIH in its survey of the status of research in dentistry determined that research on adhesive dental materials was one of three areas (the others were saliva

Fig. 1. Acrylic resinsilicate glass composite (TD71) after 12 years in the mouth (courtesy Dr J.W. McLean).

D.C. Smith / Biomaterials 19 (1998) 467 478

469

and neurophysiology) in which research should be greatly expanded and encouraged. As a result a workshop on Adhesive Restorative Dental Materials was held in September 1961 under the chairmanship of Phillips and Ryge [17]. A large number of basic and applied aspects of the subject were discussed with many creative suggestions that were to be a stimulus for future work. Further workshops were held in 1963 [18] and 1973 [19]. A major emphasis in the 1961 Workshop was on adequate wetting of enamel and dentine surfaces and the need for chemical groups in bonding agents that could compete with water. Anity tests led Bowen [20] to suggest that carboxyl groups and calcium chelating groups appeared to have greater anity for calcied tissue surfaces than did water. Although Buonocore et al. [12] had demonstrated greatly improved bonding of acrylic systems to enamel after etching with phosphoric acid, this concept did not become established until the early 1970s when BIS-GMA based materials became widely available. Experience with the enamel acid-etch technique showed the need for adequate wetting and penetration into the tooth surface with resin systems. About the same time the idea of positive physico-chemical adhesion with tooth substance to provide strong bonding was also established. Over the last 20 years these concepts of wetting and adhesion have resulted in two main streams of development. One was concerned with aqueous-based systems that could readily wet the tooth surface and undergo chemical interaction with tooth substance. The other approach involved use of polymerizable systems containing potentially reactive monomers having carboxyl or phosphate groups.

3. Polyelectrolyte cements When the author began to investigate the problem of adhesion to dental tissues in 1961 following publication of the Workshop Proceedings it was evident that hydrophilic materials capable of wetting and reacting with hydroxyapatite (HA) and/or the collagenous phase of tooth tissue (dentine) were required for durable bonding. In relation to the presence of HA in both enamel and dentine, compounds containing groups capable of chelating or complexing to calcium seemed most promising. Support for this approach was evident in the literature; for example the work of Leach and Putnam [21] in 1962 had demonstrated that a variety of mono- and dicarboxylic acids showed no appreciable chemical absorption to HA at pH 4 and 7.3 in contrast to citric acid which showed strong absorption. At this time there was a developing interest in watersoluble polymers [22], many of which had chelation or complexation properties and could form association complexes. In addition to citric acid and other polycarboxylic acids, other models for multidentate ligands for

calcium included tetracycline and its derivatives, diketones such as acetonyl acetone and polyphosphates [23]. These and similar compounds were studied both as potential adhesive ligands for calcium and as potential cement formers [23]. Much of this work was unsuccessful and was not published. However, the work did suggest the polymeric complexing agents were possible adhesive materials. Reactive polymers that were capable of bonding to collagen were also investigated including polyurethane and s-triazine systems. Dichlorotriazines were found [24] to react rapidly with the organic components of dentine and dichlorotriazinyl monomers were synthesized and investigated [25, 26]. Unfortunately although good initial bonding could be obtained the bonds were hydrolytically unstable and the materials were suspect toxicologically. These studies indicated also that water-based adhesive systems were desirable [26]. As a result of this experience the concept of polyelectrolyte cements was developed in 1963. These materials were based on reactions of metal oxides with reactive water-soluble polymers. The author had made the rst zinc polyacrylate cement using zinc oxide and a polyacrylic acid solution in that year and by 1964 had developed it into a usable system. Polyacrylic acid was chosen because of its known ability to complex calcium [27] and its formation of hydrogen bonds with organic polymers [21] comparable to collagen. Polyacrylic acid was particularly suitable since it was of simple structure and low toxicity (indeed the salts had antibacterial effects), and had been studied extensively for its metal ion complexing eects, for example, by Gregor [28, 29]. Considerable technical development followed, and following a patent application, information on the zinc polyacrylate cement was rst released in 1967 [25] followed by formal publication in 1968 [30]. Technological improvement and clinical trials followed and the material became available commercially in 1968. It is clear that the general concept of a polyelectrolyte cement allowed considerable elaboration and variation. Obviously, as disclosed in the patent literature [31], various metal oxides and/or llers could be used as well as variety of copolymers in the aqueous cement liquid. It was found that major changes in cement properties resulted from variations in the cement components. Thus inclusion of magnesium oxide in the cement powder (as was common with zinc phosphate cements) gave watersensitive materials because of non-specic binding [32]. Further, a low molecular weight and narrow distribution in the polyacrylic acid was important to avoid cobwebbing in the cement mix [31, 33]. This was achieved by the use of a chain transfer agent [31]. Thus there was a considerable technical basis for the polyacrylate cements which was developed further in the next decade as discussed elsewhere [26, 34]. Materials containing llers and uorides and copolymers such as poly(acrylic aciditaconic acid) became commercially available.

470

D.C. Smith / Biomaterials 19 (1998) 467 478

Fig. 2. Adsorption of polyalkenoic acids on hydroxyapatite: ;, polymaleic acid; , poly(acrylic acid); , poly(acrylicitaconic) acid; # , poly(ethylene maleic acid); , poly(vinyl methyl ether maleic acid). From Ellis et al. [41] with permission.

In addition to their low toxicity and good physical properties when correctly formulated, the major new feature of the polyacrylate cements was their adhesion to tooth tissue. Adhesion to enamel and dentine were demonstrated in vitro [35, 36] and in vivo, initially in orthodontic situations [37]. Evidence gradually accumulated supporting the concept of ion binding to the HA phase of enamel and dentine [34, 38]. Various studies on adsorption of polycarboxylic acids to HA have been carried out subsequently, but the precise mode of binding remains unclear. Earlier studies by the author and colleagues [39, 40] suggested that the bond to Ca was labile. Ellis et al. [41] found that adsorption of polyalkenoic acids to HA was variable in relation to structure (Fig. 2). Thus adsorption does not appear to correlate with adhesion properties. For low molecular weight material Misra [42] has also suggested ionic and hydrogen bonding coupled with phosphate displacement may be involved in the adhesion of polyacrylic acid to HA. In this context the polar character of polyelectrolytes such as polycarboxylic acids leads to wetting of many surfaces. The interaction is inuenced by the exibility of the molecule, the degree of uncoiling related to hydration and neutralization and the shape as expressed by monomer composition [34]. Polyacrylic acid is a particularly simple exible molecule from these aspects and this probably largely accounts for its unique behaviour.

basic chemistry of dental silicate cements. By modication of the Al O /SiO ratio in the silicate glass Wilson    and Kent [39] were able to produce usable cements with polyacrylic acid. Novel, more reactive glasses of high uorine content were developed and the key discovery of the eect of tartaric acid in improving the setting properties [45] resulted in the rst practical glass-ionomer cement (GIC) in 1972 [46]. Jurecic [47] had found that gelation of concentrated polyacrylic acid solutions on standing due to intra-chain hydrogen bonding could be prevented by the use of acrylic aciditaconic acid copolymers. The use of this principle [48] allowed restorative glass-ionomer cements to be produced. An alternative approach to this problem was to use dried polyacrylic acid blended with the glass so requiring only the addition of water or tartaric acid solution for the cement mix [49]. However, this formulation has been used commercially only in relatively recent years. The ensuing stages in the development of the glass-ionomer cements have been discussed in detail by Wilson and McLean [46]. The further elaboration of this system has been presented recently in a series of papers [50].

5. Composition considerations The evolution of the GICs over the last decade has resulted in changes in both the glass powder component and the polycarboxylic acid liquid. Very recently, the acidbase setting reaction between these two components has been modied by introduction of water-soluble polymers and polymerizable monomers into the composition. As a result there are signicant dierences in the composition and, thus, properties of the available commercial materials intended for the various specic applications. This background has been reviewed elsewhere [51]. As indicated above most of the research and development on GIC has concerned materials comprised of a calcium uoro-alumino-silicate glass powder and an aqueous solution of a poly(acrylic aciditaconic acid) copolymer containing tartaric acid. Modications in both components have been made in various brands for both patent and practical reasons.

6. Glass composition 4. Development of glass-ionomer cements In spite of their novel properties zinc and other metal oxide cements were opaque and unaesthetic. Formulations using zinc-containing glass ceramics and silicate cement powders were developed [31] but they did not set rapidly or satisfactorily when mixed with polyacrylic acid solutions. Similar observations were made by Wilson et al. [43] who had carried out extensive studies on the The original ion leachable glasses [49] were based on SiO AlO CaF AlPO Na AlF composition. Ac      cording to Wilson and McLean [46] the Al O /SiO    ratio is required to be 1 : 2 or more and the F content can be up to 23%. Later glasses in commercial materials contained more sodium and less uoride [46]. Typical chemical compositions are given in Table 1 [46]. More recently glasses containing Sr, Ba and Zn have been used in commercial material.

D.C. Smith / Biomaterials 19 (1998) 467 478 Table 1 Composition of ionomer glasses (from Wilson and McLean [46] with permission) Species Composition A 41.9 28.6 1.6 15.7 9.3 3.8 % Mass B 35.2 20.1 2.4 20.1 3.6 12.0

471

SiO  Al O   AlF  CaF  NaF AlPO 

The eect of variations in the glass composition has been discussed by Wilson and McLean [46], Barry et al. [52], Wilson et al. [53] and by Wood and Hill [54]. Two modications to these basic glasses are of commercial importance. Firstly, longer working time in the cement and decreased water sensitivity of the set cement can be obtained by depleting the outer 10100 m of the powder particles of calcium by treating with acids such as HCl [55]. Secondly, the powdered glass can be blended with metal powders such as silver, gold, platinum or palladium [56]. The resulting fused and ground product has been claimed to yield GIC llings with greater wear resistance than the Ca-depleted glasses referred to above [56]. Little research has been reported on other types of glasses. Bertenshaw et al. [57] developed alkaline aluminoborate glasses that were found to yield strong cements of appropriate setting characteristics with polyacrylic acid. Neve et al. [5860] have recently described such materials having improved properties though strengths are lower than conventional cements. In another approach Darling and Hill [61] have developed high-strength hydrolytically stable cements based on alumino-zinc-silicate glasses.

Fig. 3. Types of carboxylic acid units used in glass-ionomer (polyalkenoate) cement liquids. From Wilson and McLean [46] with permission.

7. Liquid composition As noted previously, the liquid component of the original glass-ionomer cements was aqueous polyacrylic acid. A low molecular weight was required to achieve a high concentration without gelation. The latter problem was overcome by using an acrylic aciditaconic acid copolymer [47, 48]. A variety of other unsaturated carboxylic acids and monomers has been proposed and patented for copolymers with acrylic acid but the principal additional copolymer materials in practical use appear to have been acrylic acidmaleic acid [62, 63] and acrylic acid3-butene-1,2,3-tricarboxylic acid [64]. The introduction of dicarboxylic or tricarboxylic acids (Fig. 3) into the polymer chain not only prevents gelation of the liquid but provides greater reactivity because of the

increased number of carboxyl groups per chain unit and the higher acidity. Increased chain cross-linking probably also occurs in the nal set cement leading to better physical properties. In the long term, however, there is some evidence that increased brittleness may occur [65]. By using homopolymers of polymaleic acid or polyitaconic acid (or their copolymers) even more reactive liquids can be made which alone or in admixture with polyacrylic acid will set rapidly with traditional silicate glasses to give products of higher strength [66]. However, these polyacids show poor adhesion to calcied tissues [63, 67]. A similar tendency can be found in the copolymers with acrylic acid which bond less strongly to enamel and dentine than polyacrylic acid. Nevertheless, if optimal proportions are present, e.g. a mole ratio of about 1 : 1 in acrylic acidmaleic acid copolymers, a more reactive liquid with adhesion comparable to polyacrylic acid can be obtained [63]. The molecular weight and distribution of the polyacid inuence the viscosity characteristics of the liquid. Higher molecular weight copolymers improve physical properties of cements [6870] and the limitation of viscosity increase in the cement liquid and mix may be at least partly overcome by incorporating the dried polyacrylic acid or copolymer powder with the glass to produce a material which is mixed with water or a tartaric acid solution. However, water-admix materials appear to have lower fatigue resistence [71]. The inuence of polyacid molecular weight on properties of GIC has been recently re-examined by Wilson et al. [72]. Darling and Hill [61] cite data showing that for polyacrylic acids of comparable number average molar mass, the fracture

472

D.C. Smith / Biomaterials 19 (1998) 467 478

Fig. 4. Structure of the repeating unit for: (a) poly(vinyl phosphonic acid); (b) poly(acrylic acid). From Ellis et al. [6] with permission.

toughness (K ) of zinc polycarboxylate cements is con'! siderably higher than glass-ionomer cements. In both cases K is substantially increased at higher molar mass. '! 8. Setting modiers Various salts were found to aect the setting rate of polycarboxylate cements [25, 26, 43, 68] but Wilson and co-workers [41, 73, 74] found that tartaric acid was particularly eective in controlling the setting reaction of glass-ionomer cement. Various metal ion chelating agents can be used but the (#) isomer of tartaric acid is unique in both prolonging working time and increasing the setting rate [46, 73]. These eects are attributed to increased attack on the surface of the glass particles and the formation of stable metal (uoro) ion complexes [75]. Up to 10% of this form of tartaric acid is incorporated in most if not all glass-ionomer cement compositions. Polyphosphonates were investigated, inter alia, by Wilson et al. [53] as modiers for the GIC setting reaction. Recently, Ellis et al. [76] have studied cements based on solutions of poly(vinyl phosphonic acid) (Fig. 4). These cements have comparable properties to conventional GIC but also faster setting, improved translucence and low early solubility. Biocompatibility properties have not been published yet. A more recent alternative approach by Kao et al. [77] utilizes N-acryloyl substituted amino acids as a route to improved performance in GIC.

9. Polymer-modiers Hybrid materials in which some of the water content of the glass-ionomer system was replaced by water-soluble monomer systems capable of ambient free radical polymerization were developed by Antonucci and colleague in 1986 [78, 79]. In this work monomers such as hydroxyethyl methacrylate (HEMA) and polyethylene glycol dimethacrylate were used with chemical redox initiator systems such as hydrogen peroxide/cupric ion or benzoyl peroxide/sulphinate salt. Alternatively visible light curing systems based on camphorquinone/tert-

amino methacrylate were also formulated. Thus such materials have a dual-setting mechanism involving the acidbase reaction of the polyacid with the glass plus the polymerization reaction. Antonucci and Stansbury [80] examined the feasibility of incorporation of water-soluble or compatible polymers into GIC alone or in combination with polymerizable systems. Such water-soluble polymers included poly(ethylene oxide or glycol), gelatin, poly(vinyl alcohol) and poly(N-vinyl pyrrolidone). Ideally, all these systems involve the formation of an interpenetrating polymer network combining the acidbase cross-linking reaction of the metal ionpolyacid with the cross-linking polymerization of the monomer system or additive action of the polymers. Antonucci and Stansbury [80] found that with both polymerizable systems and added polymer products with higher diametral tensile strengths and, in the case of a light-cured composition, higher compressive strengths could be obtained. The test specimens still showed generally brittle-type fracture but some evidence of plastic deformation was observed. A light-cured hybrid glass-ionomer cement system was developed by Mitra [81, 82] in which the powder component is composed primarily of a radiopaque uoro-alumin silicate glass powder containing a photosensitizer. The liquid is an aqueous solution of a polyacrylic acid copolymer having pendent methacrylate groups together with approximately 10% of HEMA and a photoinitiator. This material is commercially available as Vitrebond (3M) and several other materials based on similar principles have also appeared (e.g. XR Ionomer, Kerr; PhotoBond Applicap, Espe; Fuji LC, GC Corp.). Several curing systems are possible with these materials such as camphorquinone with tert-amines or sulphinates and combinations such as persulphateascorbic acid. Materials that utilize several such systems to provide light- and chemically activated polymerization in addition to the acidbase reaction have recently become available. Such cements have been described as dualcure or tri-cure to indicate a continuing polymerization after the light source has been removed. Some manufacturers, however, have created confusion by describing the materials that set by light-activated polymerization and acidbase reaction as dual cure [83] whereas the term cure has traditionally been used to refer to a polymerization reaction. These polymer-modied materials have the advantages of high early strength and water resistance and command set. Improved adhesion to dentine results probably through a combination of both chemisorption from the polyacrylic acid component and formation of a hybrid layer from the penetration of the HEMA component or other hydrophilic monomers. Extension of this latter bonding principle has led to materials containing a preponderance of polymeric components described as

D.C. Smith / Biomaterials 19 (1998) 467 478

473

GIC resin or resinomers. The classication of these materials as GIC is controversial since low water content may minimize or negate the acidbase reaction typical of GIC. The expression of polyelectrolyte characteristics requires a water matrix.

matrix. The latter is assumed to be an interpenetrating polymer network consisting of the ionic metal polyacrylate (polysilicate) hydrogel entangled with poly(hydroxyethyl methacrylate) hydrogel. It is possible that hydrophilic and hydrophobic domains are present. The elucidation of these structures and their eect on factors such as uoride release await further research.

10. Setting reaction considerations The setting reaction of the GIC is complex and may vary with composition. In general, it has been represented (Wilson and McLean [46]) as an acidbase reaction between the polyacid liquid and the glass in which Ca and Al ions are released by attack on the surface of the glass particles and ultimately cross-linking of the polyacid chains into a network. This ion release is facilitated by tartaric acid that readily forms complexes with these ions which form a molecular structure with the polyacrylate chains [61]. Initial setting (gelation) is now regarded as due to chain entanglement as well as weak ionic cross-linking which corresponds with the viscoelastic behaviour of the freshly set material. Work by Wasson and Nicholson [84] has conrmed earlier work of Cook [85] that, contrary to previous belief, there is no sequential release of Ca and Al ions. Instead, these and other species are liberated together with dierential rates of reaction in matrix formation. As the cement matures over the rst 24 h and beyond, progressive cross-linking occurs possibly with hydrated Al ions since the sensitivity to moisture of the set cement decreases and the percentage of bound water and glass transition temperature increase [46]. The nal set structure has been presented as a complex composite of the original glass particles sheathed by a siliceous hydrogel and bonded together by a matrix phase consisting of hydrated uoridated calcium and aluminium polyacrylates [46, 74]. Wasson and Nicholson [84] suggested that a silicate matrix is also slowly formed in the GIC in addition to the polyacrylate structures. Additional studies by Wasson and Nicholson [86] and Darling and Hill [61] provide further evidence that a hydrated silicate network may be a secondary setting reaction in GIC. Electron microscopic studies by Hatton and Brook [87] have conrmed this secondary setting reaction, their analytical data demonstrating that silicon is present in the matrix phase as well as aluminium and calcium. They suggested that silicate species may be involved in the setting reaction. Anderson and Dahl [88] have demonstrated that signicant release of Al occurs from both conventional and light-cured GIC at the freshly set stage with dissolution of the matrix. This is consistent with the picture of increasing stability and strength of GIC over time [65] arising from increased Al cross-linking in both polyacrylate and silicate networks. The structure of the light-cured polymer-reinforced GIC is a similar composite of glass particles and hydrogel 11. Recent development Expansion of the concept of incorporation of polymerization and polyelectrolyte salt formation mechanisms within one molecule or a combination of molecules has led to a spectrum of materials [89] which range from classical acidbase GIC to modied polymerceramic (glass) composites with little polyelectrolyte character. The objectives sought in these latter materials are to improve the strength, toughness and resistance to dissolution whilst retaining the fundamental attributes of adhesion, uoride release and biocompatibility associated with the GIC system. The need to provide adequate water in the system for polycarboxylate reactions to occur, i.e. adequate hydrophilicity, has led to diculties in some materials in balancing this requirement against the hydrophobicity required in resin cement systems. As is well known, resin cement systems, even those containing monomers having potentially functional polar groups such as carboxylate or phosphate, lack intrinsic adhesion to dentine unless a conditioning pretreatment to improve surface penetration is employed. For example, Leung and Morris [90] have demonstrated by Raman spectroscopy that 4-META in a methacrylate medium reacts too slowly with enamel for salt formation and primary bonding is more likely due to improved wetting, enhanced surface penetration and micromechanical interlocking. Further, Ruse and Smith [91] have shown previously that acid-demineralized surfaces of dentine contain very little calcium or phosphorus thus reducing still further the possibility of calcium salt formation on subsequent application of functional resinbonding systems. Treatment of dentine by acidic primers containing hydrophilic solvents or monomers is thus required to also enhance penetration facilitating formation of the so-called hybrid layer or resin diusion zone. The performance of resin-modied GIC in terms of intrinsic adhesion behaviour to dentine is thus an indication of the contribution of polyelectrolyte character and/or penetration enhancement with micromechanical interlocking to their mechanistic behaviour. Another aspect of resin-modied systems is signicant shrinkage on setting arising from monomer polymerization [92, 93]. The current status of resin-modied glass-ionomer cements has been reviewed by Sidhu and Watson [94] including the vexed question of the generic name.

474

D.C. Smith / Biomaterials 19 (1998) 467 478

Classication of these materials is made more confusing by the multiplicity and complexity of current formulations. Thus Vitremer (3M) luting cement is based on the Vitrebond (3M) copolymer developed by Mitra [81]. The copolymer is based on an acrylicitaconic acid chain having amide-linked methacrylate groups. The formulation also contains HEMA and tartaric acid and water. An acidbase reaction occurs on mixing but is overtaken by a polymerization reaction, both light and chemically initiated. Although there must be less COOH and reduced chain mobility compared to a traditional GIC, there is evidently enough interaction with tooth tissue and enhanced penetration to provide adequate bonding without a need for primer treatment. We [95] have shown recently that this system interacts chemically with dentine but bonding is reduced by a delay between application and light-curing. Another current material, Fuji Duet (GC Corp.), based on a poly(acrylic aciditaconic acid) copolymer is stated to be water based and resin reinforced. It contains tartaric acid and HEMA. It requires, however, the use of a 10/2 citric acid/ferric chloride conditioner in order to realize its bonding potential thus suggesting a greater proportion of non-ionic monomers and less polyelectrolyte character. Between these materials which lay emphasis on the GIC characteristics and products that are essentially polymer composites containing F-containing llers such as GIC glass are several hybrid materials that contain relatively large amounts of acid monomers in polymerizable systems. These approaches have been discussed by Hammesfahr [96]. Examples of these materials include Advance(Dentsply-Caulk) Hybrid Ionomer Cement and Dyract(Dentsply-Caulk). In Advance (Dentsply-Caulk) a GIC glass is mixed with a liquid containing polymerizable carboxylic acids, water and diluent monomer. The material sets by interaction between the glass and the monomeric carboxylic acid monomers in the liquid and by a chemically initiated (self-curing) reaction. Thus this approach aims to form the polyacid structure in situ. The novel carboxylic acid monomer employed is termed OEMA monomer (Fig. 5). The system as supplied is said to provide bond strength to dentine at conventional GIC levels but higher bond strength requires use of the Probond (Dentsply-Caulk) acid phosphate primer suggesting limited tooth interaction prior to setting of the composition. The degree of interaction with the glass before setting and the eect of setting contraction due to polymerization of the monomer component as well as changes in structure and cross-linking in the cement over time are not clear at present. Sucient interaction occurs, however, to provide signicant uoride release. Dyract (Dentsply-Caulk) is described as a light-cured polyacid modied composite resin rather than a modied GIC; it illustrates nevertheless the trend towards

Fig. 5. Structure of OEMA monomer present in Advance (DentsplyCaulk) resin-modied glass-ionomer luting cement (courtesy P. Hammersfahr).

Fig. 6. Structure of TCB monomer present in Dyract (Dentsply-Caulk) polyacid-modied composite resin (courtesy P. Hammersfahr).

polyacid formation during the setting process. This onecomponent system is based on an Sr/Al glass and an anhydrous urethane dimethacrylate (UEDMA) system containing an acidic monomer TCB (Fig. 6). This carboxylic acid monomer undergoes, in principle, an equilibration reaction with the glass after formulation. Bonding of the material depends on the use of the Probond primer. After polymerization, diusion of water into the hydrophobic resin is said to allow metal ion cross-linking and weak ionic bonding with tooth tissue. A low but sustained F release is claimed. The extent of water sorption and the long-term stability over time is unclear although it is stated that the material expands slightly over time. These approaches, while interesting, introduce additional questions into the performance of modied GICs. The use of a polyacid prepolymer rather than polymerizable acid monomers is more ecient with respect to reduced shrinkage and improved cross-linking. As noted previously with 4-META [92], reaction of acidic monomers with calcic material is slow especially in any anhydrous or non-polar media. The possibility of formation of heterogeneous structures after polymerization with hydrophilic and hydrophobic domains would seem probable leading to a need for assurance of dimensional stability and retention of physical properties over time. Further information is needed on biocompatibility with respect to leached material and the possibility of sensitization. Several recent studies provide some information in these areas. Attin et al. [92] found that curing shrinkage of resin-modied glass ionomers was greater than for

D.C. Smith / Biomaterials 19 (1998) 467 478

475

a hybrid composite and a conventional GIC. The resinmodied materials showed volumetric expansion after water storage. Feilzer et al. [93] determined that the conversion of shrinkage stresses into expansive stresses in the resin-modied products may prove to be detrimental in the long term. Flexural fatigue behaviour data [97] suggest that failure of the resin phase may limit the performance of a resin-modied material and that all the glass ionomers showed lower fatigue resistance than the composites studied. In general, the data presently available suggest that in the resin-modied glass-ionomer materials improved strength, fracture toughness and lower moisture sensitivity have been gained at the expense of increased overall dimensional change, reduced direct adhesion, reduced polyelectrolyte character and F release, questionable long-term stability and a lesser biocompatibility. The known long-term performance of carefully placed GIC [98] indicates that there are two distinguishing eective characteristics of the original (or classical) polyacrylate-based GIC: long-sustained uoride release and intrinsic adhesion to tooth tissue. The clinical signicance of the level and duration of uoride release is still contentious and will be addressed elsewhere. Clinical studies which measure eects such as uoride ion uptake on restored and adjacent teeth and in plaque [99] as well as eects on oral ora [100] seem desirable rather than in vitro studies. With respect to intrinsic adhesion, durable bonding of polyacrylic acid-based cements to enamel and dentine in vitro and clinically was demonstrated many years ago [3537]. As mentioned previously, the mechanisms have not yet been completely claried. To examine the durability of the bond relative to the structure of the polyacid

component, Dr Rana Sodhi and I have carried out as yet unpublished surface spectroscopic studies of the stability of the chemisorption of monomers on clean sintered hydroxyapatite (HA) surfaces. In X-ray photoelectron spectroscopic (XPS) investigations, for example, the C peak in the energy spectrum for the HA shows (Fig. 7) evidence of CH binding at 285 eV together with several minor peaks for oxygeneated components indicative of the ubiquitous presence of C contaminants. In another specimen having a layer of polyacrylic acid adsorbed from solution and dried, a characteristic peak

Fig. 8. High-resolution detail of C peak in XPS spectrum from dried lm of polyacrylic acid adsorbed onto clean hydroxyapatite showing carboxyl peak (compare Fig. 7).

Fig. 7. High-resolution detail of C peak in XPS spectrum of clean sintered hydroxyapatite surface.

Fig. 9. Corresponding spectrum to Fig. 8 after boiling specimen for 2 h showing signicant retention of carboxylate function.

476

D.C. Smith / Biomaterials 19 (1998) 467 478 [5] Schoenbeck F. Process for the production of tooth material. US Patent No. 897160, 1908. [6] Kulzer DRP. Application No. 85578-IVe/3g, 1943; French Patent No. 883679, 1953. [7] Blumenthal LM. Recent German developments in the eld of dental resins. Fiat Final Report HO 1185. Oce of Technical Services, US Dept of Commerce, Washington DC, 1947. [8] Glenn JF. Composition and properties of unlled and composite resin restorative materials. In: Smith DC, Williams DF, editors. Biocompatibility of dental materials, vol 3. Boca Raton: CRC Press, 1982:97130. [9] Hagger O. Swiss Patent No. 278946; UK Patent No. 687299, 1951. [10] McLean JW, Kramer IRH. A clinical and pathological evaluation of a sulphinic acid activated resin for use in restorative dentistry. Brit Dent J 1952;93:25569, 2913. [11] McLean JW, Kramer IRH. Alterations in the staining reaction of dentin resulting from a constitutent of a new self-polymerizing resin. Br Dent J 1952;93:1503. [12] Buonocore M, Wileman W, Brudevold F. A report on a dentin composition capable of bonding to human dentin surfaces. J Dent Res 1956;35:84651. [13] Glen JF, Clancy G. Dental cavity primer. US Patent No. 2794016, 1957. [14] Kadoma Y, Nakabayashi N, Masuhara E. Kobunshi Ronboushu, 1978;35:423. [15] Carter JA, Smith DC. Some properties of polymer coated ceramics. J Dent Res 1967;46:1274. [16] McLean JW. Anterior lling materials in Europe. Dent Prog 1962;2:1819. [17] Phillips RW Ryge G, editors. Proceedings of a workshop on Adhesive Restorative Dental Materials, Bethesda, MD, National Institutes of Health, 1961. [18] Austin RH, Wilsdorf HGF, Phillips RW, editors. Adhesive Restorative Dental Materials. II. NIDR Workshops, NIH, Bethesda, MD, Public Health Service Publication C1494, 1965. [19] Moskowitz HD, Ward GT, Woolridge ED, editors. Dental adhesive materials. New York: NY University College of Dentistry, 1973. [20] Bowen RL. Proceedings of a workshop on Adhesive Restorative Dental Materials, Bethesda, MD, National Institutes of Health. 1961;17791. [21] Leach SA, Puttnam NA. Infra-red studies of the interaction of weak acid anions with hydroxyapatite. J Dent Res 1962;41:716. [22] Davidson R, Sittig M, editors. Water soluble resins. New York: Reinhold, 1962. [23] Smith DC. Some factors involved in the development of adhesive dental lling materials. In: Alner DJ, editor. Aspects of adhesion, vol. 3. London: University of London Press, 1967:95112. [24] Crossland L, Hardwick JL, Smith DC. Caries experience in rats following tooth brushing with halogenated triazines. Caries Res 1967;1:2758. [25] Smith DC. A new dental cement. Br Dent J 1967;123:5401. [26] Smith DC. Approaches to adhesion to tooth structure. In: Silverston IJ, Dogon LM, editor. St. Paul, MN: North Central, 1975. [27] Wall FT, Drenan JW. Gelation of polyacrylic acid by divalent cations. J Pol Sci 1951;7:8390. [28] Gregor H. Complexing of ions by polyacrylic acid. J Phys chem 1955;59:349. [29] Gregor H. Complexation constants of polyacrylic acid with Mg, Ca, Mn, Co and Zn. J Phys Chem 1955;59:9901. [30] Smith DC. A new dental cement. Br Dent J 1968;125:3814. [31] Smith DC. Improvements in surgical cements. British Patent No. 1139430, 1969. [32] Jacobson AL. Congurational eects of binding of magnesium to polyacrylic acid. J Polym Sci 1962;57:321.

for COOH is present at 289 eV (Fig. 8). It was found that even after boiling the HA specimens in water for 2 h, considerable amounts of polyacrylic acid remained adsorbed to the HA (Fig. 9). In contrast, other monomers present in dentine-bonding agents and in resin-modied glass ionomers were partly or completely lost after boiling. Data that were conrmatory of these ndings were obtained using time-of-ight secondary ion mass spectrometry (TOF SIMS) techniques. These results suggest that increased hydrophobicity and reduced chain exibility and mobility in functional monomers and polymers reduce intrinsic bonding to HA. Further work on these mechanisms is indicated in relation to durability of bonding.

12. Future outlook Clinical experience has delineated the practical advantages and disadvantages of the glass-ionomer cement system. This has resulted in improved formulations and more controlled techniques. Some light-cured polymerreinforced materials appear to have substantial benets in most of the established applications while retaining the advantages of uoride release and adhesion. It is not clear at the present time if the various compositions in commercial materials that represent a gradation from wholly acidbase materials of traditional type to materials closely resembling resin composites and utilizing dentine-bonding systems for adhesion can be considered as comprising one class of materials. The contribution of polyelectrolyte character to adhesion and uoride release in the various materials requires further elucidation. It remains to be seen if the less traditional more hydrophobic materials will exhibit the same longevity and properties including biocompatibility as the original GIC. Biocompatibility questions include the issues of hypersensitivity to dimethacrylates and specically to HEMA [101] and signicance of oestrogenecity of bis-phenol A products such as BIS-GMA [102]. Further exploitation of polyacrylate and other polyelectrolyte cements is clearly possible and continued development of copolymers and additives should result in even more eective materials.

References
[1] Pierce CH. Discussion. Pennsylvania association of dental surgery. Dent Cosmos 1870;21:696 7. [2] Ames WVB. A new oxyphosphate for crown setting. Dent Cosmos 1892;34:392 4. [3] Fleck DJ. The chemistry of oxyphosphates. Dent Items Int 1902;24:906. [4] Steenbock P. Improvements in and relating to the manufacture of a material designed for the production of cement. UK Patents Nos. 15176, 15181, 1954.

D.C. Smith / Biomaterials 19 (1998) 467 478 [33] Elgood EJ, Heath NS, OBrien BJ, Solomon DH. J Appl Polym Sci 1964;8:8817. [34] Smith DC. Composition and characteristics of dental cements. In: Smith DC, Williams DF, editors. Biocompatibility of dental materials, vol. 2. Boca Raton: CRC Press, 1982:143200. [35] Mizrahi E, Smith DC. The bond strength of zinc polycarboxylate cement. Br Dent J 1969;127:4104. [36] Smith DC, Williams PD. Method of measuring the adhesion of restorative materials to enamel and dentin. J Dent Res 1967; 46:1274. [37] Mizrahi E, Smith DC. Direct cementation of orthodontic brackets to dental enamel. Br Dent J 1969;127:3715. [38] Crisp S, Prosser HJ, Wilson AD. An infra-red spectroscopic study of cement formation between metal oxides and aqueous solutions of poly(acrylic acid). J Mater Sci 1976;11:3648. [39] Wilson AD, Kent BE. The glass-ionomer cement. A new translucent dental lling material. J Appl Chem Biotechnol 1971; 21:31320. [40] Peters WJ, Jackson RW, Smith DC. Studies of the stability and toxicity of zinc polyacrylate cements. J Biomed Mater Res 1974;8:5360. [41] Ellis J, Jackson AM, Scott RP, Wilson AD. Adsorption of carboxylate cements to hydroxyapatite. III. Adsorption of polyalkenoic acids. Biomaterials 1990;11:37984. [42] Misra DN. Adsorption of low molecular weight sodium polyacrylate on hydroxyapatite. J Dent Res 1993;71:141822. [43] Wilson AD, Kent BE, Clinton D, Miller RP. The formation and microstructure of dental silicate cements. J Mater Sci 1972; 7:22038. [44] Peters WJ, Jackson RW, Iwano K, Smith DC. The biological response to zinc polyacrylate cement. Clin Orthop 1972; 88:22833. [45] Wilson AD, Crisp S. poly(carboxylate) cement. UK Patent No. 1422337, 1976. [46] Wilson AD, McLean JW. Glass-ionomer cement. Chicago: Quintessence, 1988. [47] Jurecic A. Acrylic acid copolymers for dental use. US Patent No. 3741926, 1973. [48] Crisp S, Wilson AD. Poly(carboxylate) cements. UK Patent No. 1484454, 1977. [49] Wilson AD, Kent BE. Surgical cement. UK Patent No. 1316129, 1973. [50] [Special Issue] Glass ionomer cements. Clin Mater 1991; 7:273346. [51] Smith DC. Glass ionomer cements: a prospect. In: Proceedings of the Symposium on Aesthetic Restorative Materials. American Dental Association, 1991; Chicago: American Dental Association, 1993:4959. [52] Barry TL, Clinton DJ, Wilson AD. The structure of a glass ionomer cement and its retention to the setting process. J Dent Res 1979;58:1072. [53] Wilson AD, Prosser HJ, Groman DM, Sayers GS. Cementforming compositions. UK Patent Application No. AB 2182044, 1986. [54] Wood D, Hill R. Structureproperty relationships in ionomer glasses. Clin Mater 1991;7:30112. [55] Schmitt W, Purrmann R, Jochum P, Gasser O. Calcium depleted aluminium uorosilicate glass powder for use in dental or bone cements. US Patent No. 4376835, 1983. [56] McLean JW, Gasser O. Powdered dental material and process for the preparation thereof. US Patent No. 4527979, 1985. [57] Bertenshaw B, Combe EC, Tidy DC, Laycock JNC. Surgical cement composition containing aluminoborate glass and a polymer containing recurring carboxylic or carboxylate groups. US Patent No. 4174334, 1979. [58] Neve AD, Piddock V, Combe EC. Development of novel dental cements I. Formation of aluminoborate glasses. Clin Mater 1992;9:712.

477

[59] Neve AD, Piddock V, Combe ED. Development of novel dental cements. II. Cement properties. Clin Mater 1992;9:1320. [60] Neve AD, Piddock V, Combe ED. The eect of glass heat treatment on the properties of a novel polyalkenote cement. Clin Mater 1993;12:1136. [61] Darling M, Hill R. Novel polyalkenoate (glass ionomer) dental cements based on zinc silicate glasses. Biomaterials 1994; 15:299306. [62] Tezuka C, Karasawa Y. Setting solution for dental glass ionomer cements. US Patent No. 4089830, 1978. [63] Schmitt W, Purrman R, Jochum P, Gasser O. Mixing component for dental glass ionomer cements. US Patent No. 4360605, 1982. [64] Suzuki N. Dental cement compositions. US Patent No. 3962267, 1976. [65] Mitra SB, Kedrowski BL. Long term mechanical properties of glass ionomers. Dent Mater 1994;10:7882. [66] Schmitt W, Purrmann R, Jochum P, Zahler WD. Self-hardening compound particularly for dental-medical application. US Patent No. 3882080, 1974. [67] Smith DC. Glass ionomer cements. In: Kawahara H, editor, Proceedings of the International Congress on Implantol Biomater in Stomatol. Tokyo: Ishiyaku, 1980:2654. [68] Smith DC. A review of the zinc polycarboxylate cements. J Can Dent Ass 1971;37:2230. [69] Wilson AD, Crisp S, Abel G. Characterization of glass-ionomer cements. 4. Eect of molecular weight on physical properties. J Dent 1977;5:11720. [70] Prosser HJ, Powis DR, Wilson AD. Glass-ionomer cements of improved exural strength. J Dent Res 1986;65:1468. [71] Nakajima N, Watkins JH, Arita K, Hanoaka K, Okabe T. Mechanical properties of glass ionomers under static and dynamic loading. Dent Mater 1996;12:307. [72] Wilson AD, Hill RG, Warrens CP, Lewis BG. The inuence of polyacid molecular weight on some properties of glass-ionomer cements. J Dent Res 1989;68(2):8994. [73] Crisp S, Lewis BG, Wilson AD. Characterization of glassionomer cements. 5. The eect of tartaric acid in the liquid component. J Dent 1979;7:30412. [74] Prosser HJ, Powis DR, Brant P, Wilson AD. The characterisation of glass-ionomer cements. 7. The physical properties of current materials. J Dent 1984;12:23140. [75] Nicholson JW, Brookman PJ, Lacy OM, Wilson AD. Fourier transform infrared spectroscopic study of the role of tartaric acid in glass ionomer cements. J Dent Res 1988;67(12):14514. [76] Ellis J, Anstice M, Wilson AD. The glass polyphosphonate cement: a novel glass ionomer cement based on poly(vinyl phosphonic acid). Clin Mater 1991;7:3416. [77] Kao ECC, Culbertson WM, Dong X. Preparation of glass ionomer cement using N-acroyl-substituted amino acid monomersevaluation of physical properties. Dent Mater 1996; 12:4451. [78] McKinney JE, Antonucci JM. Wear and microhardness of two experimental dental composites. J Dent Res 1986;66:11349. [79] Antonucci JM. Formulation and evaluation of resin modied glass ionomer cements. Transactions of the 13th Annual Meeting. Soc Biomater, New York, 1987:225. [80] Antonucci JM, Stansbury JW. Polymer modied glass ionomer cements. J Dent Res 1989;68(Special issue):251, Abstr 555. [81] Mitra SB. Adhension to dentin and physical properties of a lightcured glass-ionomer liner/base. J Dent Res 1991;70(1):727. [82] Mitra SB. In vitro uoride release from a light-cured glassionomer liner/base. J Dent Res 1991;70(1):758. [83] Wilson AD. Glass-ionomer cementorigins, development, future. Clin Mater 1991;7:27582. [84] Wasson EA, Nicholson JW. Studies on the setting chemistry of glass ionomer cements. Clin Mater 1991;7:28993.

478

D.C. Smith / Biomaterials 19 (1998) 467 478 [94] Sidhu SK, Watson TF. Resin modied glass ionomer materials. Am J Dent 1995;8:5967. [95] Titley KR, Smith DC, Chernecky R. SEM observations of the reactions of the components of a light-activated glass polyalkenoate (ionomer) cement on bovine dentin. J Dent 1996; 24:4116. [96] Hammesfahr PD. Development in resionomer systems. In: Hunt P, editor. Glass ionomers: the next generation. Philadelphia, 1994:4755. [97] Braem MJA, Lambrechts P, Gladys S, Vanherle G. In vitro fatigue behaviour of restorative composites and glass ionomers. Dent Mater 1995;11:13741. [98] Mount GJ. An atlas of glass ionomer cements, second edn. London: Martin Dunitz, 1994. [99] Seppa L, Salmeakivi S, Forss H. Enamel and plaque uoride following glass ionomer application in vivo. Caries Res 1992; 26:3404. [100] Koch G, Hatibovic-Kofman S. Glass ionomer cements as a uoride release system in vivo. Swed Dent J 1990;14:26773. [101] Kanerva L, Estlander T, Jolanki R, Tarvainen K. Occupational allergic contact dermatitis caused by exposure to acrylates during work with dental prostheses. Contact Derm 1993;28:26875. [102] Olea N, Pulgar R, Perez P et al. Estrogenecity of resin-based composites and sealants used in dentistry. Environ Health Perspect 1996;104:298305.

[85] Cook WD. Degradative analysis of glass ionomer polyacrylate cements. J Biomed Mater Res 1983;17:101527. [86] Wasson EA, Nicholson JW. New aspects of the setting of glassionomer cements. J Dent Res 1993;72:4813. [87] Hatton PV, Brook IM. Characterization of the ultrastructure of glass ionomer (polyalkenoate) cements. Br Dent J 1992; 173:2757. [88] Anderson OH, Dahl JE. Aluminum release from glass ionomer cements during early water exposure in vitro. Biomaterials 1994;15:8828. [89] Burgess J, Norling B, Summitt J. Resin ionomer restorative materialsthe new generation. In: Hunt P, editor. Glass ionomers: the next generation. Philadelphia, 1994:7581. [90] Leung Y, Morris MD. Characterization of the chemical interactions between 4-MET and enamel by Raman spectroscopy. Dent Mater 1995;11:1915. [91] Ruse ND, Smith DC. Adhesion to dentinsurface characterization. J Dent Res 1991;69:16103. [92] Attin T, Buchalla W, Kielbassa AM, Hellwig E. Curing shrinkage and volumetric changes of resin-modied glass ionomer restorative material. Dent Mater 1995;11:35962. [93] Feilzer AJ, Kakaboura AI, de Gee AJ, Davidson CL. The inuence of water sorption on the development of setting shrinkage stress in traditional and resin-modied glass ionomer cements. Dent Mater 1995;11:18690.

Вам также может понравиться