Sie sind auf Seite 1von 11

REVIEW

Australian Dental Journal 1999;44:(1):1-11

Resin-ionomer restorative materials for children: A review


K. M. Y. Hse, BDS, MDS* S. K. Leung, BDS S. H. Y. Wei, BDS(Hons), DDS, MS, MDS, FRACDS, FDSRCS(Eng), FADI, FHKAM(Dental Surgery)
prime candidates, resin composite and glass ionomer restoratives appear to be viable alternatives.1,2 Stainless steel crowns remain as the most durable restorations in primary molars, often surviving until the tooth exofoliates.3,4 The order of survival rates of restorations in paediatric dentistry from highest to lowest is: stainless steel crowns, amalgam, resin composite and glass ionomer cements.4 Resin composite appears to be as durable as amalgam in the short term, particularly with respect to the maintenance of good anatomic form, but it has a high long-term failure rate,3 mainly due to discolouration, loss of retention and secondary caries because there is no cariostatic action.5 Moreover, placement of a high quality composite restoration requires excellent moisture control which is more difficult to achieve with children.1 Glass ionomer cements were first introduced to the dental profession by Wilson and Kent in 1972. Their main characteristics are an ability to chemically bond to enamel and dentine with insignificant heat formation or shrinkage; biocompatibility with the pulp and periodontal tissues; fluoride release producing a cariostatic and antimicrobial action; less volumetric setting contraction; and a similar coefficient of thermal expansion to tooth structure. These advantages have made them successful as luting cements and lining mat e ri a l s. However, as a restorative material, their sensitivity to moisture and low mechanical strength and wear resistance make them the least durable. This may be adequate for primary teeth because they will exofoliate in a number of years.6,7 Recently, there have been rapid developments in the field of hybrid resin-ionomer restorative materials. These include the light-cured glass ionomer cements and the compomers. The light-cured glass ionomers were developed chiefly to overcome the problems of moisture sensitivity and low early mechanical
1

Abstract Hybrid restorative materials comprising resins and components of conventional glass ionomers have been widely introduced and accepted by the dental profession in recent years. These include the resinmodified glass ionomer cements and the polyacidmodified resin composites or compomers. They are developed in an attempt to overcome the problems of traditional restoratives, such as moisture sensitivity and reduced early strength, while at the same time maintaining their clinical advantages of command setting, adhesion to tooth structures, adequate strength to occlusal load, fluoride release and aesthetics. This paper reviews the development, composition and properties of these new materials. Their clinical performance appears to be promising and they should be considered as good alternatives to amalgam and other conventional restorative materials in the future.
Key words: Resin-modified glass ionomer cements, polyacid-modified resin composites, compomers, primary teeth. (Received for publication December 1996. Revised May 1998. Accepted May 1998.)

Introduction Dental amalgam has been the restorative material of choice for many decades. However, in recent years, there has been increasing awareness about the safety of dental amalgam, mainly in respect to possible mercury toxicity that may affect human health and the environment. These concerns have helped the dental profession to focus on the need to develop alternative restorative materials. Among the

*Acting Senior Dental Officer, Pamela Youde School Dental Clinic , Department of Health, Hong Kong Government, Hong Kong. Part-time Lecturer in Paediatric Dentistry, Faculty of Dentistry, The University of Hong Kong. Chair Professor of Paediatric Dentistry, Faculty of Dentistry, The University of Hong Kong.
Australian Dental Journal 1999;44:1.

strength, at the same time maintaining their clinical advantages. The fundamental acid-base curing reaction is supplemented by a second light-curing polymerization reaction which increases the strength and solubility resistance. They are commonly refered to as resin-modified glass ionomer cements.2,8 Their command set facility and fluoride releasing properties have made them very popular as liners and bases, especially for use under composite restorations. The compomers, as the name implies, combine the characteristics of both composites and glass ionomers into a single component. However, this group of restorative materials probably should not be classified in the same category as the resinmodified glass ionomers and a new term known as polyacid-modified resin composites appears more appropriate.8 The rapid acceptance of these new resin-ionomer hybrid restorative materials by the dental profession was largely due to their ease of use. This paper aims to provide information about the current status of these resin-ionomer restorative materials, their properties and clinical performance. Resin-modified glass ionomer cements Composition A conventional glass ionomer cement comprises a powder and a liquid component. The original ionleachable glasses, which form the basis of the powder, were based on an SiO 2-Al 2O3-CaF 2AlPO4-Na3AlF6 composition with a fluoride content of up to 23 per cent.9 The liquid component of the glass ionomer cement is usually an acrylic aciditaconic acid copolymer which has a lower viscosity than the originally used 50 per cent polyacrylic acids, thus improving the ease of mixing, providing longer working time and an increased post-set hardening rate.10 About 10 per cent tartaric acid is also incorporated into the liquid so as to prolong the working time and increase the setting rate.11 The simplest forms of resin-modified glass ionomer cements contain the addition of a small quantity of resin component such as hydroxyethyl methacrylate (HEMA) or Bis-GMA in the liquid of the conventional glass ionomers.12 Some of the water component of the conventional glass ionomer cement is replaced by a water/HEMA mixture.13 More complex materials have been developed by modifications of the polyacid with side chains that can be polymerized by a light-curing mechanism. Up to 18-20 per cent of additional resins are added to the liquid and, depending on the powder/water ratio of the mixture, about 4-5 per cent of the final cement mass can be regarded as extra resins.14 It is then possible to light-cure, resulting in an immediate setting reaction in the resins which will provide an
2

umbrella effect and protect the ongoing acid-base reaction within the cement. Chemistry of setting reaction When the powder and liquid in conventional glass ionomers are mixed together, an acid-base reaction occurs between the polyalkenoate acid and ionleachable glass, resulting in a plastic paste which then hardens to a solid mass. The final set structure is a complex composite of the original glass particles, sheathed by a siliceous hydrogel and bonded together by a matrix phase of hydrated fluoridated calcium and aluminium polyacrylates.15,16 In the resin-modified glass ionomer cement, the setting reaction is said to be a dual mechanism.13 The usual glass ionomer acid-base reaction begins on mixing the material, followed by a free radical polymerization reaction which may be generated by either photoinitiators or by chemical initiators or both. If chemical initiators are included, then the polymerization reaction will begin on mixing as well. The acid-base reaction in this modified cement system is known to slow down as some of the water has been replaced by HEMA.13 Finally, two matrices are formed: a metal polyacrylate salt hydrogel and a polymer. The initial set of the resin-modified glass ionomer cement is the result of the formation of polymer matrix and the acid-base reaction serves to harden and strengthen the formed matrix.13 A true resin-modified glass ionomer cement must be capable of setting without being photocured, that is, it will set under conditions where no polymerization reaction occurs, with the acid-base reaction still being active. 8,14,17 In summary, a true resin-modified glass ionomer cement material is a two-part system which is characterized by an acid-base reaction that is critical to its cure, a diffusion-based adhesive between the tooth surface and the cement and continuing fluoride release.18 Typical examples of true resin-modified glass ionomer cements are shown in Table 1.19 Clinical properties Adhesion Glass ionomer cements bond chemically to enamel and dentine with insignificant heat formation or shrinkage of material during the hardening reaction.20 An ion-enriched layer is developed at the interface between the cement and the tooth structure so that the cement can firmly adhere to both enamel and dentine without signs of marginal leakage.21 Resin-modified glass ionomer cements adhere to dentine in the same way as a conventional cement. Secondary ion mass spectrometry depth profiles have the ion-exchanged process between the lightcured cement and the dentine surface.22 Laboratory
Australian Dental Journal 1999;44:1.

Table 1. Typical commercial examples of resin-modified glass ionomer cements and polyacidmodified composite resins
Category Resin-modified glass ionomer cements Clinical characteristics Material Vitrebond Vitremer XR-ionomer Zionomer Fuji Lining LC Fuji II LC Photac-Bond Photac-Fil Dyract Dyract AP Compoglass Compoglass F Compoglass Flow Ionoseal Ionosit Fil Ionosit Baseliner Ionosit Seal VariGlass VLC Hytac Aplitip F2000 Manufacturer 3M, St Paul, USA 3M, St Paul, USA Kerr, Romumus, USA Denmat, Santa Maria, USA GC Corporation, Tokyo Japan GC America, Chicago, USA ESPE, Seefeld, Germany ESPE, Seefeld, Germany Dentsply De Trey, Kanstanz, Germany Dentsply De Trey, Kanstanz, Germany Vivadent, Schaan, Liechtenstein Vivadent, Schaan, Liechtenstein Vivadent, Schaan, Liechtenstein Voco, Cuxhaven, Germany DMG, Hamburg, German y DMG, Hamburg, Germany DMG, Hamburg, German y LD Caulk, Milford, USA ESPE, Seefeld, Germany 3M, St Paul, USA Application liner/base restorative liner/base liner/base liner/base restorative liner/base restorative restorative restorative restorative restorative restorative fissure sealant restorative liner/base fissure sealant restorative restorative restorative

.Usually two-paste system .Can harden without


light-curing

.Possess properties of
true glass ionomer s Polyacid-modified resin composites (compomers)

.Often one component with an adhesive system .Can only be hardened


through light-curing

.May lack the typical features


of true glass ionomer s

shear bond strength of the resin-modified cement to dentine is significantly higher than that of conventional glass ionomer cement and the bond is a stable one. Adhesion studies on human dentine have also shown high adhesion values.23,24 It may be because of the slowness of the acid-base reaction in the modified cement that the polyacid is available for a longer period, resulting in the formation of a stronger adhesive bond.13 Resin-modified glass ionomers have the advantage of being able to directly bond to resin composite. They produce a catalyst rich air-inhibited layer which can polymerize with the composite, making them useful in glass ionomer/composite laminate restorations.13,24-26 Studies have shown that the adhesion can be increased by acid etching the enamel27 or by application of dentine bonding agents prior to the placement of the restorative.28,29 Biocompatibility Polyacrylic acids are weak acids but their high molecular mass and chain entanglement make it difficult for them to penetrate through the dentinal tubules.30 Moreover, dentine is an excellent buffer and the polyacrylic acids are readily precipitated by the calcium ions in the tubules.31 Being similar to conventional glass ionomer cement, the resinmodified glass ionomers are also highly biocompatible to the pulp. The improved adhesion to dentine significantly reduces marginal leakage at the tooth/restoration interface, displaying substantially better adaptation and seal to the cavity preparation than conventional glass ionomer materials.24,32-34 It is now apparent that calcium hydroxide lining does not have a therapeutic effect on pulp tissue.35-37 It only serves to isolate the pulp from bacterial insult and is conducive to healing. Glass ionomer cement
Australian Dental Journal 1999;44:1.

can achieve the same purpose by its ability to closely adhere to tooth structure without microleakage and thus essentially isolate the lesion. The presence of a sub-lining would in fact decrease the surface area for adhesion with the inherent risk of encouraging leakage.38 Mechanical strength Conventional glass ionomer cement is a very weak material and it lacks early mechanical strength. The final set structure shows a dramatic increase in compressive strength but is rather brittle and comparatively low in tensile strength39 and has low abrasion resistance40 making it unsuitable for high stress-bearing areas such as posterior teeth. Moreover, it is adversely affected by both moisture contamination and dehydration.41 Water absorption causes the cement to lose its translucency and results in erosion of the weakened surface while dehydration at this stage results in surface crazing.42 Therefore, protection of the cement during setting and finishing procedures is required to ensure maximum translucency and mechanical strength. Inclusion of the resin component into the conventional glass ionomers allows rapid development of strength and more resistance to early moisture contamination. The set cement has improved diametral tensile strength, compressive strength and elastic modulus, when compared with its conventional counterparts.12,13,24,43-45 The resinous component renders it tougher and less brittle. Fluoride release The glass powder particles contain up to 23 per cent fluoride, some of which will be released from the glass, mainly in the form of sodium fluoride. Many investigators have demonstrated the ability of
3

glass ionomer to increase the fluoride content in enamel and dentine adjacent to restorations.46-50 The uptake of fluoride would increase its resistance to acid demineralization and prevent caries formation around restorations.48,51-53 Fluoride release from glass ionomers also has an antimicrobial action against Streptococcus mutans in plaque.54-56 Several laboratory and clinical studies have clearly demonstrated the ability of the resin-modified glass ionomers to release fluoride.24,57,58 The fluoride release from and uptake by the resin-modified products was higher than or the same as that of conventional glass ionomers59-61 and has no adverse effect on the bond strength.50 Clinical use Dentine pretreatment It has been proposed that a brief 10 second application of a 10 per cent polyacrylic acid on the prepared dentine surface will remove the gross debris without opening up the dentinal tubules or demineralizing the tooth surface. 21 However, there is still no clinical evidence that such conditioning will enhance the adhesion of glass ionomer. Previous studies have shown an increase in bond strength of the conventional glass ionomer cements with dentine pretreatment,62,63 while more recent laboratory studies have found no such effect.28,64,65 For the resin-modified glass ionomer cement, it has been suggested that the cement is adhesive in nature and does not require surface conditioning of tooth structure, probably due to the HEMA component.42 Recent studies have found no significant difference between various pretreatment methods on marginal gap formation at the tooth/restoration interface of a resin-modified glass ionomer.66,67 Handling and manipulation The resin-modified glass ionomer cement has improved setting characteristics. There is a longer working time because the resin slows down the acidbase reaction. It sets sharply once the polymerization reaction is initiated by light. This minimizes disruption of the ionomeric component and reduces the effects of moisture contamination in the early stage of setting. It is also easier to apply since the consistency remains constant until light-cured, and exact positioning of the base may be undertaken without haste. A recent study has found that most of the current resin-modified glass ionomers have greater curing shrinkage than the conventional chemically-cured cements.68 Incremental placement techniques should always be used to ensure complete curing at depth and to minimize polymerization shrinkage.18 Most manufacturers state that immediate polishing can be
4

carried out after light-curing. However, the setting reaction will continue slowly for at least 24 hours and the best result can be obtained if finishing is delayed. When immediate polishing is required, care must be taken not to overheat the restoration as this may cause excessive drying and cracking and may prevent setting of the ionomeric component.69 It is recommended to place a layer of resin seal or a fissure sealant coating over the polished surface to cover up the exposed porosities and to reinforce the surface in the short term.14,43,69,70 Problems of resin-modified glass ionomer cements In summary, the resin-modified glass ionomer cements are superior to the conventional glass ionomer cements by providing a longer working time with a command set upon light-curing, easier clinical procedures and manipulation, improved mechanical strength and aesthetics towards that of resin composite. This makes them a highly desirable alternative to amalgam for restoring primary teeth. However, compared with resin composites, they are generally more difficult to handle because they require skillful mixing techniques in order to give the correct consistency, otherwise the paste may be too sticky during placement or harden too quickly before contouring can be finished. Moreover, their overall strength and aesthetic properties are still inferior to that of resin composites. When compared with the conventional glass ionomers, there may be difficulties for intraoral placement of the curing tip to different parts of the mouth especially in small children. Finally, most of these products lack the long-term research needed to evaluate their clinical performance and wear resistance and their life expectancy still remains unknown. Polyacid-modified resin composites (compomers) Recently, other resin-ionomer hybrid restoratives have been marketed as multipurpose materials or are resins that may release fluoride but have only limited glass ionomer properties. One such new material is the compomer which contains the major ingredients of both composites (resin component) and glass ionomer cements (polyalkenoate acid and glass fillers component) except for water.17 However, in contrast with the resin-modified glass ionomers, they have a limited dual setting mechanism. The dominant setting reaction is the resinous photopolymerization and no acid-base reaction can occur until later when the material absorbs water. The name compomer means that the material possesses a combination of the characteristics of both composites and glass ionomers,71 but actually it
Australian Dental Journal 1999;44:1.

shows minimal glass ionomer reactions. In fact, a more preferred nomenclature of polyacid-modified resin composites has been suggested but is less widely accepted.8,17,72-75 Typical examples of these new products are shown in Table 1.17,75 All of them have in common the following characteristics: single paste light-curing materials with glass particles as fillers and at least two different resins for the matrix, including a light-curable monomer like urethane dimethacrylate (UDMA) or Bis-GMA.75 The clinical performance of the polyacid-modified resin composites or compomers has not been fully evaluated since they have only been marketed for a few years. Among them, Dyract (Table 1) was one of the earliest compomer restoratives available and it has been used in many clinical trials for comparison with the resin-modified glass ionomer cements. Many new compomer restorative materials are available which are claimed to have better physical properties, lower wear, more fluoride, higher quality of marginal seal and smoother surface than the first and second generations of compomers. These include Dyract AP, Compoglass F, Compoglass Flow, F2000 and Hytac Aplitip (Table 1). Most of the information regarding the composition, physical properties and performance of the polyacid-modified resin composites is based on short-term clinical reports, abstracts, laboratory studies or from the product information leaflets supplied by the manufacturers.71 Results of long-term clinical trials are not yet available. Composition Taking Dyract as an example, such a system is characterized by a unique single-component compomer restorative and a newly developed primer/adhesive liquid for enhanced adhesion to tooth tissues and improved seal of the cavity. The filling material consists of two resins forming the matrix of the final paste. The urethane dimethacrylate (UDMA) monomer is well-established for its tissue compatibility in light-cured lining materials. The ingredient that significantly contributes to the innovative character of Dyract is TCB resin. It consists of a new monomer of dual functionality, made up of a butane tetracarboxylic acid backbone with a polymerizable hydroxyethylmethacrylate (HEMA) side chain. The resultant new monomer contains two methacrylate groups as well as two carboxyl groups. The former can cross-link with other methacrylate terminated resins when initiated through radical polymerization while the later groups can undergo acid-base reaction to form a salt with metal ions and water. Considering the recently available compomer, Dyract AP, the ingredients of the restorative are basically similar to the original Dyract. The organic
Australian Dental Journal 1999;44:1.

matrix of the restorative has been modified by adding a small amount of a highly cross-linking monomer which brings an enormous increase in hardness and strength of the matrix almost equal to that of a hybrid resin composite, Spectrum TPH, and considerably higher than the original Dyract.76 In F2000 compomer restorative, the resin matrix is comprised of three monomers: the dimethacrylate functional oligomer (CDMA oligomer) derived from citric acid, the hydroxypropylene dimethacrylate which is commonly known as glyceryl dimethacrylate (GDMA) and a high molecular mass hydrophilic polymer. The CDMA oligomer has a greater ratio of methacrylate groups to carboxyl groups which allows greater cross-linking of the resin matrix. GDMA is chemically and functionally similar to HEMA with a hydrophilic hydroxyl group which acts as a diluent for the CDMA and copolymerizes with the oligomer. The high molecular mass hydrophilic polymer is an essential and unique ingredient in the F2000 compomer formulation. It rapidly takes up a controlled amount of fluid from the oral cavity which facilitates the transport of fluorides. Due to its large size and flexibility, it acts as a rheology modifier that contributes to the clinical handling characteristic of the compomer.77 Of equal importance for the final properties of the restorative system is the reactive silicate glass filler. Dyract, the first generation compomer, contains solely glass ionomer fillers. The finely milled glass, with a mean particle size of 2.5 m, accounts for 72 per cent (m/m) of the composition and also contains 13 per cent (m/m) of fluoride. In the second generation product Compoglass, the fillers are a combination of methacrylate monomers and conventional glass ionomer fillers with a mean particle size of 1.5 m, which produce an additional stability to the cross-linkage, with improvements in physical properties.78 In the recently marketed compomers, great variations exist among the formulation, particle size and loading of the glass filler. In F2000, the fluoro-aluminosilicate glass filler has an average particle size of about 3 m and a maximum of about 10 m. A small amount of colloidal silica is added with the filler and contributes to a loading of about 84 per cent by mass. Ground Ca-Al-Znfluoroglass is the filler used in Hytac Aplitip. It has a mean particle size of 5 m and accounts for 66 per cent of the composition by mass. Glass fillers with smaller size, 1 m or less, are found in other recently marketed compomers. Dyract AP, had its strontiumfluoro-silicate glass filler of mean particle size reduced to 0.8 m from 1.5 m and a loading of

Dentsply De Trey, Kanstanz, Germany. 5

about 73 per cent by mass. The Ba-Al-fluorosilicateglass filler with a particle size of 1 m accounts for 60 per cent by mass of Compoglass F compomer restorative. The general reduction of filler particle size in the newer compomers has resulted in an improvement in wear and strength and better polishing and fluoride release in this type of restorative material. When Dyract was developed, the prime objective was to build adhesiveness into the system at a level sufficiently high to make acid etching unnecessary. The primer/adhesive consists of three resins: PENTA, a patented dipentaerythritolpentacrylate phosphoric acid, which contains an acidic monomer made up of phosphoric acid with a polymerizable methacrylate group attached and is responsible for the formation of ionic bonds to the inorganic part of the tooth; TGDMA and an elastomeric resin which are specially synthesized to determine the level of cross-linking among the different monomers and, more importantly, the elasticity of the cured primer/ adhesive. Acetone acts as a solvent that carries the resins, helps to wet the tooth surface and assists the penetration of the resin in the dentine surface. Recently, a new bonding agent, Prime & Bond 2.1,, has been developed and used with Dyract AP compomer restorative material. It retains the ease of use of original Dyract-PSA but has an increased ability to absorb the stresses caused by chewing and temperature fluctuations. Major ingredients of the new bonding agent are similar to the original Dyract-PSA, with the exception of the presence of cetylamine hydrofluoride to deliver an additional amount of fluoride to teeth.76 Other bonding systems also have been developed to be used for individual compomer restorative materials by different product m a n u fa c t u r e rs. For example, Syntac Singlecomponent is used with Compoglass F, F2000 compomer primer/adhesive is used with F2000, Hytac OSB is used with Hytac Aplitip. Different from the previously developed primer/adhesive systems, the F2000 primer/adhesive and Syntac Single-component bonding adhesive are hydrophilic in nature. They contain resin monomer (HEMA), methacrylate modified polyacrylic acid and maleic acid in an aqueous solution of water. Thus, they are less volatile and highly suitable for use on moist dentine surfaces. Acetone is the solvent in the Prime & Bond 2.1 and Hytac OSB bonding adhesive. All the bonding systems in the recently marketed compomer restorative materials require light-curing. With these adhesive systems, adequate bond strength to tooth substance is obtained clinically

without the need for traditional acid etching procedures. Chemistry of setting reaction Restorative materials such as Dyract AP can only be hardened through photopolymerization. There are two stages of setting reaction. The first stage is the dominant free radical polymerization identical with that occuring in resin composite. Upon lightcuring, the polymerizable molecules of UDMA and the patented TCB resins are interconnected into a three dimensional network which is reinforced by means of the enclosed filler particles. At this stage, the existing carboxyl groups on the TCB molecules still remain inactive since Dyract AP is an anhydrous formulation and therefore no ion-exchange process takes place. Exclusion of water is essential in preventing premature setting of the material in the container but also ensures that setting occurs only by photopolymerization.17 After the initial set, the polymerized bulk of Dyract AP begins to absorb water in the moist environment of the mouth. With the presence of water, Dyract AP now contains all the ingredients that are necessary to initiate an ionic acid-base reaction as in glass ionomers. The acidic conditions in Dyract AP, by vitrue of the carboxyl group on the TCB molecules, cause metal cations to be liberated from the reactive silicate glass, which eventually leads to the formation of hydrogels in the resin structure of the compomer, although the rigidity of the set material at this stage means that the extent to which such a reaction can occur is limited.17 This additional acid-base reaction results in further crosslinkages of the entire matrix. Depending on the size of the restoration, the absorption of water will continue for several months until the filling material has reached its maximum level of water content, which is estimated to be approximately three per cent of water (m/m) at the most in Dyract.71 The volumetric change following water absorption is considered to be low and insignificant.68 Setting reactions of all the recently marketed compomers are also based on the dominant light-initiated free radical polymerization followed by a later acid-base reaction. Clinical Properties Adhesion In the Dyract AP restorative system, two different mechanisms are responsible for the formation of adhesive bonds to the cavity wall. One of these is the self-adhesive property of the restorative itself. Fifty per cent of the reactive units of the patented TCB monomer in the restorative consists of hydrophilic carboxyl (-COOH) groups. Such polyelectrolytes
Australian Dental Journal 1999;44:1.

,Dentsply De Trey, Kanstanz, Germany. 6

can bond to both enamel and dentine without acid etching. The functional carboxyl groups can form ionic bonds with the calcium ions of the tooth surface. It is suggested that some secondary valence bonding like hydrogen bonding may occur as well. The second mechanism is adhesion to the tooth surface through the primer/adhesive system. The hydrophilic phosphate group of the PENTA resin in the adhesive will form ionic bonds with the calcium ions of the hydroxyapatite. In addition, when lightcured, the three methacrylate-based resins in the adhesive will undergo free radical addition polymerization. The cross-linked resins form a reinforced zone, similar to the hybrid zone, of the surface dentine, and make both the enamel and dentine compatible for the actual restorative.71,76 The formation of hybrid layers at the dentine interface is further supported by a scanning electron microscope study investigating the mechanism by which the primer/adhesive of a recently marketed compomer bonds to tooth substance. 77 In the study, the etching on the conditioned enamel surface appeared more shallow. Partial opening of the dentinal tubules was found on dentine but not to the extent typically seen with phosphoric acid treatment. Additional mechanical bonding with resin tags formation was found in the dentinal tubules which confirmed the penetration of the compomer primer/adhesive. After placement of the restorative over the cured adhesive, reactions between the methacrylate groups of the coupling agent and the restorative are assumed to take place. The manufacturer demonstrated that if the Dyract primer/adhesive is not applied prior to the placement of the restorative, the adhesion to enamel and dentine will be reduced by a factor 2 and 4 respectively.71 Dyract has a significantly higher bond strength to dentine than other resin-modified glass ionomer cements and chemically cured glass ionomer.71,79-81 When comparing the dentine bond strength of this first generation product, a higher value is obtained in recently marketed compomers.77,82 The bond strength to dentine of Hytac Aplitip was found to be comparable to a commercially available resin composite, TPH.83 However, the bond strength of these compomers to enamel is definite and more modest. Adhesion of Dyract to enamel is enhanced with the use of acid-etching technique. A higher bond strength can be obtained when the enamel has been etched by 10 per cent or 35 per cent phosphoric acid prior to the placement of the primer/ adhesive.27,84,85 Perhaps this procedure should be a routine step to further increase the bond strength of the restorative system. Strength and wear performance In order to withstand high chewing forces in the oral cavity, a filling material intended for long-term
Australian Dental Journal 1999;44:1.

use needs to have high compressive and flexural strength. The initial and long-term compressive strength, diametral tensile strength and transverse strength of Dyract are found to exceed that of the conventional and resin-modified glass ionomers.71,81,86 A recent study on Dyract also showed that its compressive strength, flexure strength and abrasion resistance were not significantly different from that of a hybrid resin composite.87 Laboratory studies on Dyract AP revealed that the value of compressive strength and flexure strength were almost equal to that of the hybrid resin composite, Spectrum TPH, and considerably higher than original Dyract.76 Substance loss or wear is one of the most critical parameters that significantly affects the longevity of a dental restorative. The manufacturer performed a laboratory three-body abrasion test on Dyract and the results indicated that the total substance loss corresponding to five years was approximately 200250 m, about twice the amount recorded for Prisma APH, a posterior resin composite.71 The same test was carried out to evaluate the wear resistance of the recently marketed compomers. Similar wear resistance was demonstrated among them and a higher wear resistance was found when compared with the first generation products.76,77,88-90 They were significantly more wear resistant than the resinmodified and chemically cured glass ionomer cements.77,89 Laboratory studies on Dyract AP and Hytac Aplitip even showed a wear value similar to that of resin composites.76,89 The introduction of smaller, submicrometre filler in newer compomers is probably the reason for the improvement in their wear resistance. Long-term clinical studies are more reliable to determine the wear performance of restorative materials. Recent clinical studies on Dyract placed in Class I and II cavities in primary molars have found an overall wear of 100 m in six months, and 190 m in 12 months.91,92 In the latter, the authors also reported that after two years of clinical service, visible wear without exposure of dentine had been detected in most of the restorations. A much lower wear value for Dyract was obtained in other studies. In one study, the wear value was only 20.8 m in six months, about one-fifth that of the first study.93 In another controlled clinical study, the wear value of Dyract was 43.3 m in six months and 72.7 m in 12 months which was about three times the wear of a hybrid resin composite, Prisma TPH, being investigated in the same study.94 In a later study, the 24 months wear values of Dyract and TPH were 113 m and 63.9 m respectively, which were considered to be low and clinically acceptable.95 The reasons for

LD Caulk, Milford, USA. 7

such differences in the wear value are unknown but may be related to different methods of measurement and criteria used. Little is known about the clinical wear performance on the recently marketed compomer restorative materials. Fluoride release The release of fluoride ions from Dyract compomer is well documented.71,96-99 Laboratory investigations on the release of fluoride ions lasting for more than 12 months have been carried out by the manufacturer. Their results indicated that even after one year, Dyract not only continued to release fluoride ions but, most importantly, maintained the same rate of diffusion. Secondary ion mass spectrometry in prepared and filled third molars has shown fluoride uptake into the enamel which was the highest at the contact zone and in the adjacent layer of some 20 m in thickness. The concentration dropped as the sites were moved further away from the cavity wall and eventually reached the level of the natural fluoride concentration of sound enamel at a depth of 75 m. The increase in concentration of fluoride in the adjacent tooth structure was equal to that of traditional glass ionomer with proven anticariogenic properties.71 Recently, studies have found that the release of fluoride by Dyract was significantly less than resinmodified glass ionomer cement or other fluoride releasing resin composite.59,96,100-102 Moreover, unlike glass ionomer cement, it was not affected by fluoride replenishment.59 Similar to glass ionomer cement, it acted as a fluoride reservoir. When it was exposed to fluoride ion sources, such as toothbrushing with a fluoridated dentifrice, fluoride would be adsorbed and slowly released into the surroundings after the ion source was removed.96-98 This may be an effective caries preventive measure for an adjacent tooth. It is further supported by a laboratory study investigating the influence of fluoride release on Dyract that showed an initial inhibitory effect on the Streptococcus mutans growth.99 However, the antibacterial action decreased significantly over time. The cumulative fluoride release of the recently marketed compomers was found to be higher than the first generation products.76,77,82 One newer compomer, Compoglass F, has as much as 50 per cent more fluoride release than its original, Compoglass.82 The increase in fluoride release is partially due to the finer particle size of the fluoride glass contained in the newer compomer restoratives and the incorporation of additional fluoride in some of the primer/adhesive systems.76,82 A laboratory study on Dyract showed that it could release more fluoride in an acidic solution of pH 3.101 Another newer compomer could also demonstrate similar effect when immersed in a medium of pH 4.5 without
8

adversely affecting the strength of the material.77 This could be assessed as a beneficial anticariogenic effect arising from the presence of a low pH soft drink causing a high fluoride outflow each time a potential carious site was challenged. However, the glass ionomer material was essentially unaffected by the change of pH.77 In addition, the caries inhibition effect of compomer restorative material was higher than the conventional type of resin composite.103,104 It appears, therefore, that Dyract or other polyacidmodified resin composites are probably effective in terms of cariostatic properties. Optical properties The aesthetic qualities of a material are determined by its colour and opacity. The colour is obviously important for the overall match of a filling to its surroundings if the material is to look at all natural. Dyract AP is available in a range of twelve different shades which follows the Vita shade guide76 and, in general, most of the recently marketed compomers have a broader range of shades for selection than the first generation compomers. However, only F2000 and Compoglass F compomers have specialty shades for primary teeth.77,82 Less obviously, the opacity of a material must also be correct and Dyract AP has an optimal range for opacity of about 40-45 per cent, similar to that of the resin composite materials.76 Furthermore, the determined radiopacity of Dyract AP is 2.5 times that of dentine which has a radiopaque value of 2. Thus for Dyract AP, it should be around 5 which is slightly higher than that of enamel (3.5). This value is considered to be desirable for radiographic detection of recurrent caries105 and offers an easy method for documentation of dental work. Handling and manipulation Ease of manipulation is another advantage of the compomer restoratives. Similar to resin composites, they are supplied in compules which require no mixing. The gun provided allows easy dispensing directly to cavities and surfaces. The manufacturer of Hytac Aplitip compomer has even designed a fine aplitip system in conjunction with its small applicator that allows a better view of the working area. Since the adhesive can provide sufficient bond strength for retention, no acid etching procedure is required prior to placement of the restorative. The consistency makes it easy to apply and contour without stickiness and, therefore, less time will be required for final finishing. These properties are especially beneficial in treating children because restorations usually can be completed much faster and within the tolerance of the child patient. As with other light-curing restoratives, polymerization shrinkage is a problem. A recent study has
Australian Dental Journal 1999;44:1.

shown that curing shrinkage is similar to that of the conventional hybrid resin composites.68 Therefore, placement in increments of 3 mm or less is recommended for Dyract AP, 2 mm or less for other newer compomers, and then each to be cured for at least 40 seconds. Finishing can be undertaken immediately after curing using fluted tungsten carbide finishing burs or polishing discs.71,76,77,82,90 Clinical performance Reports of 6 month, 12 month and 24 month evaluations of the clinical performance of Dyract as a restorative in primary molars have been published recently. Direct evaluation results using the United States Public Health Service (USPHS) system and indirect replica techniques have shown promising results.91-94,106 In these studies, baseline assessments have shown restorations of excellent quality. No post-operative sensitivity was reported at any time throughout the study period. In a controlled clinical study comparing the performance of Dyract with a resin composite, Prisma TPH, on restoration of primary teeth, Dyract was comparable to TPH in colour match, marginal integrity, anatomical form changes and formation of secondary caries after one year. The only significant difference was marginal discolouration and wear, in which TPH was better than Dyract. In this study, an overall failure rate after one year was 1.7 per cent.94 The two year results of the same study demonstrated an additional statistically significant difference in the colour match between the two materials in which Dyract was inferior to TPH.95 The cumulative failure rate for two years was only 1.9 per cent which was comparable with another study with a failure rate of 2.5 per cent.92 However, the wear of these restorations in two years was unlikely to necessitate their replacement. The results of these short-term clinical reports have shown that the polyacid-modified resin composite or compomer is a reliable restorative material for use in primary teeth. Unfortunately, in some studies, carefully designed control groups were not included and, therefore, their clinical performance could not be directly compared with those of another material used for similar purpose. The results of these studies tended to be over-optimistic, especially when the research was directly supported by manufacturers. Conclusion Variations in composition and chemistry among the commercial products marketed under the same group of hybrids or resin-ionomer restorative materials may directly affect their properties and clinical characteristics. They may or may not have the typical features of true glass ionomers such as chemical adhesion to tooth structures and long-term fluoride release. Therefore, they should be used
Australian Dental Journal 1999;44:1.

carefully, closely following the instructions of the manufacturers because different handling methods may influence their clinical behaviour. The development of both resin-modified glass ionomers and polyacid-modified resin composites has greatly enhanced the efficacy and effectiveness of restoring carious teeth. The extent of success of these restorative materials is very promising as more long-term controlled clinical trials evaluating their clinical performance become available. In reviewing their advantages and clinical characteristics, they appear to be extremely suitable alternatives to the conventional restorative materials. They also have a particular role in the restoration of primary teeth. References
1. Christensen GJ. Restoration of pediatric posterior teeth. J Am Dent Assoc 1996;127:106-108. 2. Williams P. Goodbye amalgam, hello alternatives? J Can Dent Assoc 1996;62:139-144. 3. Kilpatrick NM. Durability of restorations in primary molars. J Dent 1993;21:67-73. 4. Papathanasiou AD, Curzon MEJ, Fairpo CG. The influence of restorative material on the survival rate of restorations in primary molars. Pediatr Dent 1994;16:282-288. 5. Qvist V, Qvist J, Mjr IA. Placement and longevity of toothcolored restorations in Denmark. Acta Odontol Scand 1990;48:305-311. 6. Walls AWG, Murray JJ, McCabe JF. The use of glass polyalkenoate (ionomer) cements in the deciduous dentition. Br Dent J 1988;165:13-17. 7. Welbury RR, Walls AWG, Murray JJ, McCabe JF. The 5-year results of a clinical trial comparing a glass polyalkenoate (ionomer) cement restorative with an amalgam restoration. Br Dent J 1991;170:177-181. 8. McLean JW, Nicholson JW, Wilson AD. Proposed nomenclature for glass ionomer dental cements and related materials. Quintessence Int 1994;25:587-589. 9. Wilson AD, Kent BE. The glass-ionomer cement: A new translucent dental filling material. J Appl Chem Biotech 1971;21:213. 10. Crisp S, Ferner AJ, Lewis BG, Wilson AD. Properties of improved glass ionomer cement formulations. J Dent 1975;3:125-130. 11. Wilson AD, Crisp S and Ferner AJ. Reactions in glass ionomer cements. IV) Effect of chelating co-monomers on setting behaviour. J Dent Res 1976;55:489-495. 12. Mathis RS, Ferracane JL. Properties of a glass ionomer/resincomposite hybrid material. Dent Mater 1989;5:355-358. 13. Wilson AD. Resin-modified glass ionomer cements. Int J Prosthodont 1990;3:425-429. 14. Mount GJ. (a) Glass ionomer cements and future research. Am J Dent 1994;7:286-292. 15. Smith DC. Composition and characteristics of glass ionomer cements. J Am Dent Assoc 1990;120:20-22. 16. Hatton PV, Brook IM. Characterisation of ultrastructure of glass-ionomer (poly-alkenoate) cement. Br Dent J 1992;173:275-277. 17. McCabe JF. Resin-modified glass-ionomers. Abstracts: 1st European Union Conference on Glass-ionomers. Coventry: University of Warwick, 1996:24-28. 18. Sidhu SK, Watson TF. Resin-modified glass ionomer materials: A status report for the American Journal of Dentistry. Am J Dent 1995;8:59-67. 19. Sidhu SK, Watson TF. Resin-modified glass-ionomer materials. Part I: Properties. Dent Update 1995;22:429-432. 20. Croll TP. Glass ionomers for infants, children and adolescents. J Am Dent Assoc 1990;120:65-68.
9

21. Mount GJ. Adhesion of glass-ionomer cement in the clinical environment. Oper Dent 1991;16:141-148. 22. Lin A, McIntyre NS, Davidson RD. Studies on the adhesion of glass-ionomer cements to dentine. J Dent Res 1992;71:18361841. 23. Mitra SB. Adhesion to dentine and physical properties of a light-cured glass ionomer liner/base. J Dent Res 1991;70:72-74. 24. Burgess JO, Barghi N, Chan DCN, Hummert T. A comparative study of three glass ionomer base materials. Am J Dent 1993;6:137-141. 25. Fortin D, Vargas MA, Seift EJ. Bonding of resin composites to resin modified glass ionomer s. Am J Dent 1995;8:201-204. 26. Friedl KH, Schmalz G, Hiller KA, Mortazavi F. Marginal adaptation of composite restorations versus hy b ri d ionomer/composite sandwich restorations. Oper Dent 1997;22:21-29. 27. Cortes O, Garcia-Godoy F, Boj JR. Bond strength of resin-reinforced glass ionomer cements after enamel etching. Am J Dent 1993;6:299-301. 28. Bell RB, Barkmeier WW. Glass-ionomer restoratives and liners: Shear bond strength to dentine. J Esthet Dent 1994;6:129-134. 29. Hinoura K, Miyazaki M, Onose H. Dentine bond strength of light-cured glass-ionomer cements. J Dent Res 1991;70:15421544. 30. McLean JW. Glass ionomer cements. Br Dent J 1988;164:293300. 31. Hume WR, Mount GJ. In vitro studies on the potential for pulpal cytotoxicity of glass-ionomer cements. J Dent Res 1988;67:915-918. 32. Crim GA. Marginal leakage of visible light-cured glass ionomer restorative materials. J Proshet Dent 1993;69:561-563. 33. Hallet KB, Garcia-Godoy F. Microleakage of resin-modified glass ionomer cement restorations: An in vitro study. Dent Mater 1993;9:306-311. 34. Sidhu SK. Marginal contraction gap formation of light-cured glass ionomers. Am J Dent 1994;7:115-118. 35. Lee SJ, Walton RE, Osborne JW. Pulp response to bases and cavity depths. Am J Dent 1992;5:64-68. 36. Leinfelder KF. Changing restorative traditions: The use of bases and liners. J Am Dent Assoc 1994;125:65-67. 37. Cox SJ, Suzuki S. Re-evaluating pulp protection: Calcium hydroxide liners vs cohesive hybridization. J Am Dent Assoc 1994;125:823-831. 38. Mount GJ. Glass ionomer cements: Past, present and future. Oper Dent 1994;19:82-90. 39. Prosser HJ, Powis DR, Brant P and Wilson AD. The characterisation of glass ionomer cements. 7. The physical properties of current materials. J Dent 1984;12:231-240. 40. Moore BK,Platt J, Phillips RW. Abrasion resistance of glass ionomer restorative materials. J Dent Res 1984;63:Spec Iss:276:Abstr 946. 41. Mount GJ, Makinson OF. Glass ionomer cements: Clinical applications of the setting reaction. Oper Dent 1982;7:134-141. 42. McLean JW. Clinical application of glass ionomer cements. Oper Dent 1992;17:Suppl 5:184-190. 43. Cho E, Kopel H, White SN. Moisture sensitivity of resin-modified glass-ionomer materials. Quintessence Int 1995;26:351-358. 44. Li Jianguo, von Beetzen, Sundstrom F. Strength and setting behaviour of resin-modified glass ionomer cements. Acta Odontol Scand 1995;53:311-317. 45. Uno S, Finger WJ, Fritz U. Long-term mechanical characteristics of resin-modified glass ionomer restorative materials. Dent Mater 1996;12:64-69. 46. Swartz ML, Phillips RW, Clark HE. Long term fluoride release from glass ionomer cement. J Dent Res 1984;63:158-160. 47. Retief DH, Bradley EL, Denton JC, Switzer P. Enamel and cementum fluoride uptake from a glass ionomer cement. Caries Res 1984;18:250-257. 48. Forss H, Seppa L. Prevention of enamel demineralization adjacent to glass ionomer filling materials. Scand J Dent Res 1990;98:173-178.
10

49. Skartveit L, Tveit AB, Totdal B, Ovrebo R, Raadal M. In vivo fluoride uptake in enamel and dentine from fluoride containing material. J Dent Child 1990;57:97-100. 50. Mitra SB. In vitro fluoride release from a light-cured glass ionomer liner/base. J Dent Res 1991;70:75-78. 51. Hicks MJ, Flaitz CM, Silverstone LM. Secondary caries formation in vitro around glass ionomer restorations. Quintessence Int 1986;17:527-532. 52. Tyas MJ. Cariostatic effect of glass ionomer cement: A five-year clinical study. Aust Dent J 1991;36:236-239. 53. Varpio M, Noren JG. Artificial caries in primary and permanent teeth adjacent to composite resin and glass ionomer cement restorations. Pediatr Dent 1994;16:107-109. 54. Forss H, Jokinen J, Spets-Happonen S, Seppa L, Luoma H. Fluoride and mutans streptococci in plaque grown on glass ionomer and composite. Caries Res 1991;25:454-458. 55. Loyola-Rodriguez JP, Garcia-Godoy F, Lindquist R. Growth inhibition of glass ionomer cements on mutans streptococci. Pediatr Dent 1994;16:346-349. 56. Svanberg M, Mjr IA, Orstavik D. Mutans streptococci in plaque from margins of amalgam, composite and glass ionomer restorations. J Dent Res 1990;69:861-864. 57. Hatibovic-Kofman S, Koch G. Fluoride release from glass ionomer cement in vivo and in vitro. Swed Dent J 1991;15:253258. 58. Kupietzky A, Houpt M, Mellberg J, Shey Z. Fluoride exchange from glass ionomer preventive resin restorations. Pediatr Dent 1994;16:340-345. 59. Forsten L. Resin-modified glass ionomer cements: fluoride release and uptake. Acta Odontol Scand 1995;53:222-225. 60. Momoi Y, McCabe JF. Fluoride release from light-activated glass ionomer restorative cements. Dent Mater 1993;9:151-154. 61. Forss H. Release of fluoride and other elements from lightcured glass ionomers in neutral and acidic conditions. J Dent Res 1993;72:1257-1262. 62. Powis DR, Folleras T, Merson SA, Wilson AD. Improved adhesion of a glass ionomer cement to dentine and enamel. J Dent Res 1982;61:1416-1422. 63. Swift EJ. Glass ionomer: A review for the clinical dentist. Gen Dent 1986;34:468-471. 64. White GJ, Beech DR, Tyas MJ. Dentine smear layer: An asset or a liability for bonding? Dent Mater 1989;5:379-383. 65. Tyas MJ. The effect of dentine conditioning with polyacrylic acid on the clinical performance of glass ionomer cement. Aust Dent J 1993;38:46-48. 66. Sim TPC, Sidhu SK. The effect of dentine conditioning on light-activated glass ionomer cement. Quintessence Int 1994;25:505-508. 67. Pachuta SM, Meiers JC. Dentin surface treatments and glass ionomer microleakage. Am J Dent 1995;8:187-190. 68. Attin T, Buchalla W, Kielbassa AM, Hellwig E. Curing shrinkage and volumetric changes of resin-modified glass ionomer restorative materials. Dent Mater 1995;11:359-362. 69. White SN. Light-cured glass ionomers. J Can Dent Assoc 1994;22:39-43. 70. Williamson RT. Protection of glass ionomer cements during the setting reaction. J Prosthet Dent 1995;73:400-401. 71. Dentsply De Trey. Dyract product profile. Kanstanz: Dentsply De Trey, 1994. 72. Smales RJ, Koutsikas P. Occlusal wear of resin-ionomer restorative materials. Aust Dent J 1995;40:171-172. 73. Smith DC. Evolution and development of glass ionomer cement systems. Abstracts: 1st European Union Conference on Glassionomers. Coventry: University of Warwick, 1996:1-2. 74. Guggenberger R, May R, Stefan KP. New trends in glass ionomer chemistry. Abstracts: 1st European Union Conference on Glass-ionomers. Coventry: University of Warwick, 1996:3-5 75. Meyer J-M, Cattani-Lorente M-A. Compomers - Between glass-ionomer cements and composites. Abstracts: 1st European Union Conference on Glass-ionomers. Coventry: University of Warwick, 1996;32.
Australian Dental Journal 1999;44:1.

76. Dentsply De Trey. Dyract AP technical manual. Kanstanz: Dentsply De Trey, 1997. 77. 3M. F2000 compomer restorative system. Technical product profile. St Paul: 3M, 1997. 78. Vivadent. Compoglass product documentation. Schaan, Liechtenstein: Vivadent, 1995. 79. Aboush YEY, Torabzadeh H. Bond strength of a compomer to enamel and dentin. J Dent Res 1994;73:Spec Iss:812:Abstr 208. 80. Triana R, Prado C, Garro J, Garcis-Godoy F. Dentin bond strength of fluoride-releasing materials. Am J Dent 1994;7:252254. 81. Kielbassa AM, Wrbas KT, Hellwig E. Initial tensile bond strength of resin-modified glass ionomers and polyacid-modified resins on perfused primary dentin. J Dent Child 1997;64:183-187. 82. Vivadent. Compoglass F, Compoglass Flow. Scientific documentation. Schaan, Liechtenstein: Vivadent, 1998. 83. Schuh H, Richter R, Watts DC. Shear bond strength of a compomer adhesive system to hard dental tissues. J Dent Res 1997;76:314:Abstr 2401. 84. Triolo PT, Barkmeier WW, Los SA. Bonding efficacy of a compomer using different conditioning procedures. J Dent Res 1995;74:107:Abstr 761. 85. Abate PF, Bertacchini SM, Polack MA, Macchi RL. Adhesion of a compomer to dental structures. Quintessence Int 1997;28:509-512. 86. Uno S, Finger WJ, Fritz U. Long-term mechanical characteristics of resin-modified glass ionomer restorative materials. Dent Mater 1996;12:64-69. 87. Attin T, Vataschki M, Hellwig E. Properties of resin-modified glass-ionomer restorative materials and two polyacid-modified resin composite materials. Quintessence Int 1996;27:203-209. 88. Frey O, Sogolowsk W. Correlation of abrasion resistance and mechanical properties of compomers. J Dent Res 1997;76:75: Abstr 495. 89. De Gee AJ, Feilzer AJ, Werner A, Davidson CL. Wear performance of polyacid modified resin composites. J Dent Res 1997;76:74:Abstr 486. 90. ESPE. Hytac product dossier. Seefeld: ESPE, 1997. 91. Peters MCRB, Roeters FJM, Frankenmolen FWA. Clinical evaluation of Dyract in primary molars. Am J Dent 1996;9:83-88. 92. Roeters FJM, Frankenmolen FWA. Two-years clinical evaluation of class I and II compomer restorations in deciduous molars. Programme and Abstracts: 15th Congress International Association of Pediatric Dentistry. Sweden: IAPD, 1995:27:Abstr S8. 93. Krejci I, Gebauer I, Hausler T, Lutz F. Kompomere Amalgamerasatz fur Milchzahnkavitaten? (Composite polymers - An amalgam substitute for deciduous tooth cavities?) Scheweiz-Monatsschr-Zahnmed 1994;104:724-730. 94. Hse KMY, Wei SHY. Clinical evaluation of compomer in primary teeth; 1-year results. J Am Dent Assoc 1997;128:10881096.

95. Leung SK, Wei SHY, Hse KMY. Clinical evaluation of compomer in primary teeth: 2-years results. 13th Annual Scientific Meeting. Hong Kong: The University of Hong Kong, 1998:Abstr 24. 96. Suljak JP, Hatibovic-Kofman S. A fluoride release-adsorptionrelease system applied to flouride-releasing restorative materials. Quintessence Int 1996;27:635-638. 97. Nunez A, Burgess JO, Chan DCN. Fluoride release and uptake of six fluoride releasing restorative materials. J Dent Res 1997;76:Spec Iss:Abstr 2485. 98. Rasmussen TE, Froerer JJ, Hollis RA, Christensen RP. Long term fluoride release from compomers and flowable resins. J Dent Res 1997;76:Spec Iss:Abstr 2487. 99. Friedl KH, Schmalz G, Hiller KA, Shams M. Resin-modified glass ionomer cements: long term fluoride release and influence on Streptococcus mutans growth. Eur J Oral Sci 1997;105:81-85. 100. Cardenas HL, Madhure Baldwin V, Burgess JO, Chan DCN. Short term fluoride release of restorative materials. J Dent Res 1995;74:107:Abstr 768. 101. Lavis JF, Peters MCRB, Makinson OF, Mount GJ. Changes to Dyract restorative resin immersed in various media. Am J Dent 1997;10:133-136. 102. Stassinakis A, Hugo B, Hiraoda J, Hotz P. Fluoride release from light cured glass ionomers and composites in comparison to conventional glass ionomer cements. J Dent Res 1995;74:435:Abstr 273. 103. Erlenbaugh AM, Donly KJ. Caries inhibition at margins of fluoride releasing restorative mat e ri a l s. J Dent Res 1995;74:247:Abstr 1887. 104. Dionysopoulos P, Kotsanos N, Papadoyiannis I, Konstantinidis A. Artificial secondary caries around some new F-containing restoratives. Abstracts: 3rd Congress of the European Academy of Paediatric Dentistry. Belgium: EAPD, 1996:31:Abstr 6. 105. Matsumura H, Sueyoshi M, Tanaka T, Atsuta M. Radiopacity of dental cements. Am J Dent 1993;6:43-45. 106. Papagiannoulis L, Kakaboura A, Pantaleon PH, Kavadia K. Clinical evaluation of a compomer in class II restorations in deciduous teeth. Abstracts: 3rd Congress of the European Academy of Paediatric Dentistry. Belgium: EAPD, 1996:33:Abstr 10.

Address for correspondence/reprints: Professor S. H. Y. Wei, Paediatric Dentistry, Faculty of Dentistr y, The University of Hong Kong, Prince Philip Dental Hospital, 34 Hospital Road, Hong Kong.

Australian Dental Journal 1999;44:1.

11

Das könnte Ihnen auch gefallen