13.07.2015 Views

Quality of dental restorations. FDI Commission Pro... - ResearchGate

Quality of dental restorations. FDI Commission Pro... - ResearchGate

Quality of dental restorations. FDI Commission Pro... - ResearchGate

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

123that continues over time in an individualavoiding fluoride result insecondary caries. However, fromclinical experience we know thatan incident <strong>of</strong> interest, such as theexample here using secondarycaries, can <strong>of</strong>ten be caused by morethan one set <strong>of</strong> sufficient causes.Thus, different causal pathwaysmay exist in different situations.Causal pathways (alternativelytermed causal web or cause-andeffectrelationships) involve theactions <strong>of</strong> risk factors acting individually,in sequence, or togetherthat result in an incidence <strong>of</strong> interest.These pathways may varywith different sets <strong>of</strong> risk factors.Understanding these pathways isnecessary to devise preventivecountermeasures or interventions toavoid a specified outcome, and thecountermeasure may be unique tothe pathway.Causal relationship can be determinedusing various levels <strong>of</strong>evidence. In theory, all informationregarding a hypothetical causalrelationship can be labelled evidenceand must therefore be appraised.However, formal requirements areneeded to address validity <strong>of</strong>evidence, and this is applied onscientific data. Inferences <strong>of</strong> causalrelationship are directly associatedwith study design. For many clinicalquestions, randomised controlledtrials (RCTs) are regarded asthe strongest evidence for causalrelationship. However, many determinants<strong>of</strong> aetiology, diagnosis andprognosis can for various reasonsonly be estimated indirectly usingcross-sectional, cohort or casecontrolstudy designs. In thesesituations, inference must beassumed on the basis <strong>of</strong> how findingssatisfy different criteria <strong>of</strong>causation.Statistical issuesClinical studies may be classified asexperimental or observational.Only studies with experimentaldesigns can be considered inductive,that is, can give an indication<strong>of</strong> a cause-effect relationshipbetween different factors or variableswith a certain degree <strong>of</strong>uncertainty. All other methodsinvolve limitations through bias orconfounding. However, data fromobservational studies should notbe regarded as unimportant orincorrect. Hypotheses are <strong>of</strong>tengenerated first on the basis <strong>of</strong>observational studies, and are thentested for validity under morerigorous experimentally designedconditions.Certain requirements must befulfilled to qualify as an experimentalstudy. These are the presence <strong>of</strong>control groups, predefined allocation<strong>of</strong> variables, and standardisedevaluation procedures and criteriafor the evaluation <strong>of</strong> outcomes. Theallocation <strong>of</strong> variables is randomisedif possible, in order to make evenstronger statistical inferences, i.e. arandomised controlled trial. Thespecific aim <strong>of</strong> the study and theformulation <strong>of</strong> a hypothesis shouldbe documented. When these criteriaare not met, or when observationsare made <strong>of</strong> phenomena that arenot manipulated by the investigator,a clinical study is classified asobservational.Relatively few clinical studies inrestorative dentistry fulfil the criteria<strong>of</strong> an experimental design 5,16,17,63 .The majority <strong>of</strong> clinical studieswhere an association has beenreported between clinical variablesand restoration performance havebeen observational studies. Thisis because although the studieswere experimentally designed toobtain information on differencesbetween, for example materials orcommercial products, the observationsand descriptions <strong>of</strong> theinfluence <strong>of</strong> other factors were notobtained by the manipulation <strong>of</strong>these factors.Many clinical studies are carriedout according to recommendationsoutlined by various national orinternational acceptances programmeguidelines, for example <strong>FDI</strong> 64 ,NIH 65 and ADA 66–68 . These guidelinesare designed to address ‘safety’and ‘efficacy’, that is attempts toscore performance as passing orfailing and not to rank clinicalperformance. This is why many <strong>of</strong>the criteria are based on a passinglevel <strong>of</strong>, for example 85–90 percent alpha scores according to theUSPUS criteria after one to threeyears, and only distinguish betweenunacceptable and acceptable performance.For different reasons theseguidelines do not require controls,do not test for placebo effects anddo not have statistical powers greatenough to answer anything otherthan simple experimental questions.The most commonly tested hypothesisis whether a new material orproduct has been comparable to aspecific traditional material.Many <strong>of</strong> these trials are carriedout in research environments, asopposed to general practitioners’practices. The operators are <strong>of</strong>tenselected and trained to ensure optimalhandling. Furthermore, thepatients are <strong>of</strong>ten <strong>dental</strong> students,<strong>dental</strong> school staff or dentists withabove average oral hygiene 12,13,16,17,69 .Controlling operators and theirworking environment, patients, andsize and intra-oral location <strong>of</strong> the<strong>restorations</strong> reduces confounding,when comparing different materialsor products. However, datafrom such studies do not reflectthe situation in ‘real-world’ <strong>dental</strong>practice 70 . This is especially apparentwhen technique-sensitivematerials are involved 1 . In generalpractice, treatment times areconstrained, the diagnostic thresholdsfor replacement may vary withthe patient load, and there are noeconomic incentives to producehigher clinical standards aboveacceptable 61 . In general, there ispublic concern that there is lack <strong>of</strong>data on clinical performance <strong>of</strong>restorative materials and on thequality <strong>of</strong> service provided bydentists in general practice, andespecially on the interaction betweenclinical performance <strong>of</strong> <strong>restorations</strong>and quality <strong>of</strong> service 71 .It is evident from the literaturethat there are disagreements concer-<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


124ning the material, operator andpatient effects on restoration quality.One <strong>of</strong> the major issuesappears to be the statistical treatment<strong>of</strong> data. It is difficult toconduct clinical studies, with theaim <strong>of</strong> establishing a numericalrelationship between one specificrisk factor and the technical excellence,restoration service period orreplacement reasons. The mainreason is that the clinical performance<strong>of</strong> a restoration is dependenton many known and unknownclinical variables that are difficultto control or record. It is alsodifficult, if not impossible, toassure independence among manyclinical variables that affect restorationclinical performance. Currently,there do not appear to be anygenerally acceptable, valid statisticaltechniques for isolating theinfluence <strong>of</strong> a single variable;indeed, many <strong>of</strong> the variables maynot be independent. In full knowledge<strong>of</strong> this situation, clinicalresearchers employ various strategieswhen designing trials aimingto clarify parameters <strong>of</strong> restorationquality (Figure 3).Restoration quality has beenaddressed in both prospective andretrospective longitudinal studies aswell as in cross-sectional studies.Data from prospective and retrospectivelongitudinal studies can beused for constructing survivalcurves, proportions <strong>of</strong> <strong>restorations</strong>with varying technical excellence asa function <strong>of</strong> time and reasons forreplacement.Current restorative materialshave excellent physical and mechanicalproperties. <strong>Pro</strong>spective clinicalstudies therefore need to beextended for many years and/orinclude large numbers <strong>of</strong> <strong>restorations</strong>before any strong statisticalinferences can be made. Longobservation periods are associatedwith problems such as patient dropouts,patient representativity andchanges in the clinician’s diagnosticabilities or understanding <strong>of</strong> replacementcriteria. Finally, also ethicalreasons may occasionally restrict theFigure 3. Parameters <strong>of</strong> the quality <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>. The horizontal axis representstime, while the vertical axis is the proportion <strong>of</strong> <strong>restorations</strong> remaining over time. The boldline represents the proportion <strong>of</strong> <strong>restorations</strong> remaining in situ (that is ‘survival’), and theseare in the category ‘excellent’ (for example ‘alpha’), ‘acceptable’ (‘beta’) or ‘defective’(‘charlie’). Replacements may be ‘true’ reasons, that is bulk fractures, secondary caries,marginal deficiencies, discolouration, etc, or because <strong>of</strong> faulty diagnosis, inclusion intolarger <strong>restorations</strong>, primary caries on other surfaces, etc. Arbitrarily lines depict theproportion <strong>of</strong> <strong>restorations</strong>/replacements that fall into the carious categories. The horizontalline at 50% marks the median survival time. The intersections between this line and the otherlines represent from left to right when 50% <strong>of</strong> the remaining <strong>restorations</strong> are: excellent,acceptable and remaining in situ.possibility <strong>of</strong> conducting prospectiveclinical studies.Retrospective studies are basedon analysing patient records or acombination <strong>of</strong> patient records andquality evaluation <strong>of</strong> <strong>restorations</strong>.A frequent problem with manyretrospective studies is that little orno information is available onpossible reasons for replacement.Several studies have revealed thatreplacements are not alwaysexplained by restoration failure 1 ,and even if they are, retrospectivedata give no indication regardingwhether the failures are relateddirectly to the restoration, to therestorative process or to externalfactors 72 .Cross-sectional clinical studieshave either been presented asreplacement studies or recordings<strong>of</strong> data from patients’ records.Other data have been derived fromassessment <strong>of</strong> technical excellence<strong>of</strong> <strong>restorations</strong> in situ or inextracted teeth, or from detailedstudies <strong>of</strong> failed <strong>restorations</strong>.Cross-sectional studies that focuson mean age <strong>of</strong> failed <strong>restorations</strong>identify the ‘geometric centre’ <strong>of</strong>the area above the survival curve inFigure 3, while studies that focuson the age <strong>of</strong> remaining <strong>restorations</strong>identify the centre below thiscurve.In replacement studies, theprevious history and age <strong>of</strong> the<strong>restorations</strong> is <strong>of</strong>ten unknown.Although the type <strong>of</strong> materialusually is recognised, specific tradenames or batch numbers areseldom recorded. A characteristic<strong>of</strong> the study method is that theevaluation criteria are not explicit,which leaves the diagnoses to theoperators involved in the study. Theresults do not indicate any causalrelationships, and they are probablyinfluenced by factors such associoeconomy, patient demographyand the dentist:patient ratio. Thesame arguments are applicablewhen interpreting results fromcross sectional studies. Although theevaluation criteria may <strong>of</strong>ten beaccurately described, the history andclinical parameters at the time <strong>of</strong>restoration placement remainsunknown.International Dental Journal (2001) Vol. 51/No.3


128cement, composite resin/resinmodifiedglass-ionomer cementand resin-modified glass-ionomercement/glass-ionomer cement.Several studies have reportedcomparisons between ‘sandwich’versus homogeneous <strong>restorations</strong>,but the conclusions are conflicting.Patient factorsEmpirical observations indicate thatnot only surface, but also bulk andmarginal discolouration vary amongpatients. However, very few studieshave identified specific patientfactors that may influence the opticalcharacteristics <strong>of</strong> <strong>restorations</strong>.Marginal discolouration alongveneers made from compositeresin on 87 maxillary anteriorteeth in 23 young patients wasmore common among smokerscompared to non-smokers in alongitudinal study over five years 102 .In a study <strong>of</strong> 52 pairs <strong>of</strong> ClassIII micr<strong>of</strong>illed composite resin<strong>restorations</strong> after eleven years,surface discolouration was most<strong>of</strong>ten recorded among smokers,and significant correlations werefound between the patients’consumption <strong>of</strong> alcoholic beveragesand body and surfacediscolouration 88 .AdaptationTraditionally, the terms used todescribe adaptation have variedwith the examination method, thetype <strong>of</strong> restoration and the nature<strong>of</strong> the restorative material. Horizontaldiscrepancies on smooth andapproximal surfaces have <strong>of</strong>tenbeen termed ‘overhangs’, while theterm ‘marginal ditching’ has beenused to describe defects alongmargins on the occlusal surfaces <strong>of</strong>teeth containing mainly amalgam<strong>restorations</strong>. The terms over- andunder-extension, with additionaldescriptors <strong>of</strong> the cement marginmorphology, for example ‘marginalwear’ and ‘cement excess’, haveusually described the adaptation <strong>of</strong>indirect <strong>restorations</strong>. Finally, theterm ‘gap’ has also been used formany years to infer a lack <strong>of</strong> adaptationbetween the materials andtooth tissues. However, this term israther ill defined and non-specific.Discrepancies measured alonga horizontal or a vertical axistangential to the interface haveoccasionally been interpreted assynonymous to marginal adaptation.Such discrepancies can beassessed clinically, but will necessarilyonly express the adaptationalong the margin on the toothsurfaces. Several methods havebeen used to assess adaptation <strong>of</strong>the entire restoration-tooth interface,but all these methods aredestructive 103 .The clinical evaluation <strong>of</strong> marginaldiscrepancies is questionable,explained by a lack <strong>of</strong> reliablediagnostic skills <strong>of</strong> clinicians. It hastherefore been argued that thescientific community must acceptthat <strong>dental</strong> restorative materials willbe misjudged during the process<strong>of</strong> evaluating the marginal qualities<strong>of</strong> <strong>restorations</strong> 104 .Material factorsThe <strong>dental</strong> literature containsnumerous papers in whichmarginal degradation has beenevaluated as a function <strong>of</strong> materialcomposition. The prevailing materialin these studies has been amalgam,and there is consensus that highcopperalloys are clinically superiorto low-copper alloys 105 .Inter-group differences in respect<strong>of</strong> the marginal degradation <strong>of</strong>amalgam <strong>restorations</strong> <strong>of</strong>ten appearafter a short time, and remainconstant 29,69,106–109 . This signifies thatat least one process that results inmarginal fractures occurs duringthe first year after placement <strong>of</strong>the restoration. This hypothesis,however, does not identify orexclude other aetiological factorsthat may be associated withmarginal fractures, including creep,mercuroscopic expansion, biomechanicalrelationships, bulk andcrevice corrosion or fatigue rupture.There may also be differencesbetween composite resins, Bryantet al. 110 having reported thatparticular types <strong>of</strong> marginal defectsare commonly associated withspecific types <strong>of</strong> composite resins.In general, <strong>restorations</strong> <strong>of</strong> micr<strong>of</strong>illedcomposite resins show moremarginal degradation compared to<strong>restorations</strong> <strong>of</strong> other types <strong>of</strong>composite resins.Operator factorsJokstad 29 reported an influence <strong>of</strong>the operator on the performance<strong>of</strong> 468 amalgam <strong>restorations</strong> <strong>of</strong>five alloys placed in 210 patientsafter five years. The five dentistswere all able to obtain superiormarginal adaptation with the bestalloys, and contrary to the findings<strong>of</strong> Mahler and Marantz 21 , all fivealloys performed equally well forthe five operators. Further, it wasevident that one operator alsoobtained satisfactory marginaladaptation with a low-copperalloy. It was concluded that themain operator variables influencingthe marginal adaptation werethe final condensation <strong>of</strong> the amalgamand the treatment <strong>of</strong> thesurface and margins.Mahler and Marantz 21 reportedon <strong>restorations</strong> <strong>of</strong> four amalgamalloys placed by four operators.The amalgams were chosen on thebasis <strong>of</strong> their marginal fracturebehaviour as found in an earlierstudy, ranging from little to extensivefracture. Following placement<strong>of</strong> the <strong>restorations</strong>, three-yearevaluation <strong>of</strong> marginal fracture wasundertaken using a linear ratingscale. It was found that the operatorinfluenced the marginal fractureindex, but in different waysdepending on the alloy. For thealloys with the most and leastmarginal fracture, there was nooperator difference. However, forthe two intermediate alloys, therewere large differences among theoperators. Overall, the associationwith alloy was stronger than withoperator, and it was thereforeInternational Dental Journal (2001) Vol. 51/No.3


129recommended that non-gamma-2amalgams should be used.Cavity designStratis and Bryant 111 carried out atwo-year study <strong>of</strong> 111 Class I andII amalgam <strong>restorations</strong> placed byone operator, and reported that acombination <strong>of</strong> modification <strong>of</strong> theocclusal cavo-surface angle andfinishing <strong>of</strong> the restoration had aninfluence on the marginal fractureat two years.Kreulen et al. 112 reported a photographicevaluation <strong>of</strong> the margins<strong>of</strong> 245 Class II amalgam <strong>restorations</strong>placed by three dentists. Theprinciple variable influencingmarginal adaptation was the dentist.In addition, improvement <strong>of</strong> themarginal adaptation by an occlusalbevel was discussed compared withnon-bevelled margins.Investigations published before1992 on the possible relationshipbetween marginal fracture andClass I and Class II cavity preparationsfor amalgam has beenreported in a previous paper, andwill not be discussed further 113 .Fukushima et al. 114 reported astudy on the early marginal breakdown<strong>of</strong> 432 posterior compositeresin <strong>restorations</strong>. It was determinedthat smaller cavities, greaterbulk <strong>of</strong> material at the margin(especially in functional cusp areas),and well-finished margins withoutoverfilling seem to reduce theoccurrence <strong>of</strong> marginal fracture.Material handling and proceduresSuccessful bonding is associatedwith several technique-relatedfactors. The use <strong>of</strong> adhesives istechnique sensitive because <strong>of</strong>complex multi-step applicationtechniques 38,115 . Careful management<strong>of</strong> the status <strong>of</strong> the collagenmeshwork is important to preventit from disintegration or collapseand thereby ensuring optimal resinpenetration 116 .A clinical technique, commonlyreferred to as ‘wet bonding’, hasbeen recommended especially foradhesive systems that utilise waterfree,acetone-containing primers.However, recent research hasrevealed that bonding systems thatutilise water-based primers appearto bond with equal effectiveness todry and wet dentine 117 . Adhesivesystems using acetone-based primersrevealed a higher techniquesensitivity 118 , whereas adhesivesystems containing water-basedprimers appear to be less techniquesensitive,as far as the remainingwetness <strong>of</strong> the acid-etched dentinesurface is concerned 119 .Clinically, the exact timing <strong>of</strong>the different stages <strong>of</strong> bonding asrecommended by the manufactureris <strong>of</strong>ten very difficult. For example,parts <strong>of</strong> the dentine may beetched for the same amount <strong>of</strong>time as the enamel because theprecise differentiation between thetwo substrates is not always possible.Excessive etching may result indemineralisation depths that aregreater than monomers can effectivelypenetrate 115,119 , and causesevere collapse <strong>of</strong> the collagenmeshwork 119 . The depth <strong>of</strong> demineralisationis dependent on etchingtime and phosphoric acid concentration120 , while the thickness <strong>of</strong> themonomer penetration, or ‘hybrid’,layer is a function <strong>of</strong> conditioningtime 121 . However, the implications<strong>of</strong> these variables on long-termclinical outcomes remain uncertain.IsolationDesiccation <strong>of</strong> the demineraliseddentine causes collapse <strong>of</strong> thecollagen meshwork, which impedesthe proper infiltration <strong>of</strong> theprimer 121 . Therefore, a wet bondingtechnique is recommended 122,123 .However, there is a wide range <strong>of</strong>interpretations <strong>of</strong> ‘wet’ 124,125 , withno clear guidelines in manufacturers’directions for use. While thenegative consequences <strong>of</strong> excessiveair-drying are well documented foracetone-based systems 122,123 , theresults for water-based systems arevariable 125 .Contamination <strong>of</strong> the etchedenamel surface with saliva prior tothe placement <strong>of</strong> a resin-basedmaterial significantly reduces thebond strength to enamel. Studieson the influence <strong>of</strong> saliva contaminationon dentine bonding arevariable. Although the tolerance <strong>of</strong>modern adhesives to saliva contaminationhas improved, reductionsin bond strengths may be anticipatedafter saliva contamination. Itis therefore important to preventsaliva contamination after application<strong>of</strong> the primer 126 .The effect <strong>of</strong> using rubber damremains uncertain. In an experimentalstudy comparing shear bondstrength <strong>of</strong> 36 composite resin<strong>restorations</strong> placed either withcotton rolls or under rubber dam,no significant differences werenoted between the two groups 127 .In another study using the sameprotocol, microleakage was assessed.This study concluded that the use<strong>of</strong> rubber dam isolation resulted inless microleakage at the enamelresininterface 128 .Patient factorsIntraoral locationAs for wear, bite force is probablya significant aetiological factorregarding the extent <strong>of</strong> materialdeterioration. As a result <strong>of</strong> relativelyhigh bite forces in the firstmolar region, it can be assumedthat more marginal fractures willoccur in this location.In a study <strong>of</strong> 88 composite resin<strong>restorations</strong> placed by nine <strong>dental</strong>students, it was observed after twoand three years that marginal integritywas significantly better inpremolars compared to that inmolars 100,101 .Berg and Derand 129 reporteddata on 51 out <strong>of</strong> originally 115porcelain inlays made with theCerec technique after five years in46 patients. No significant differencesin marginal ditching weredetected between molars andpremolars.Jokstad 29 did not find any strongrelationship between marginaldegradation and intraoral locationafter five years observation <strong>of</strong> ClassII amalgam <strong>restorations</strong>. Only thelower premolars showed less<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


130marginal degradation compared tothe other tooth groups. Osborneand Gale 130 reported that themarginal fracture <strong>of</strong> high-copperamalgam <strong>restorations</strong> could not berelated to the intraoral locationafter fourteen years service. Interactionsbetween tooth position andwidth indicated that lowerpremolars with conservative <strong>restorations</strong>exhibited the least fractureat the margins, and upperpremolars with a wide preparationexhibited the most. Osborne andGale 131 reported that the marginaldegradation <strong>of</strong> 429 Class II amalgamsafter two years was less inlower premolars than in the otherposterior teeth.Goldberg et al. 132 studied 475<strong>restorations</strong> over 1.5 years. UsingANOVA analyses, these investigatorscompared marginal fracturescores among different subgroups,categorised by intraoral location andfound more fractures in molars thanin premolars.Oral environmentDerand 133 assessed <strong>restorations</strong> <strong>of</strong>four amalgam alloys in 163 teethafter 2.5 years. The patients weredivided into three levels <strong>of</strong> bitingforce. It was found that marginalfractures increased with increasingbite force for the conventionalalloys, but that the relationship wasnot significant for the three nongamma-2alloys.Restoration failure andclinical factorsTechnical excellence <strong>of</strong> <strong>restorations</strong>deteriorates in clinical service andmay or may not be linked to restorationfailure. Certain investigators,including those <strong>of</strong> Harris 134 andOwens 135 , reflect a common opinionthat materials themselves are<strong>of</strong>ten the least <strong>of</strong> the problems andthat most restoration failures canbe attributed to poor attention todetail in cavity preparation andmaterial handling. However, thisview can at best be regarded asexpert opinion, and is linked to thefailure criteria. In a survey <strong>of</strong> 571International Dental Journal (2001) Vol. 51/No.3dentists, perceived causes <strong>of</strong> restorationfailure were ranked bypatient-related factors (45 per cent),dentist-related (35 per cent) andmaterial choices (20 per cent) 136 . Anamusing secondary finding was thatthese estimates were for <strong>restorations</strong>in general, while the percentageswere 48, 26 and 26 per centrespectively when dentists addressedthe causes <strong>of</strong> failure <strong>of</strong> the <strong>restorations</strong>they had made themselves.The following sections presentthe effects <strong>of</strong> the numerousdependent and independent variablesinfluencing the quality <strong>of</strong><strong>restorations</strong>, notably the operator;the operative techniques and instrumentused; the material; the location;type; size; initial and short-termtechnical excellence <strong>of</strong> the restorationand patient factors.General performanceMaterial factorsDifferences in composition andphysical propertiesSubtle differences in physical propertieswithin specific materialgroups may be considered to havea small effect on clinical performancein general practice settings. Aparticular exception is perhaps compositeresin cements for indirect<strong>restorations</strong>, where the micr<strong>of</strong>illedcements seem to be superior tocements with larger fillers 14,136 ,although conflicting results havebeen reported 138 . Two longitudinalmulticentre studies involving 24dentists in seven clinics revealed onlyminor differences amongst sixcomposite resins after three andfive years 139,140 . Dunne and Millar 141reported the evaluation <strong>of</strong> 315porcelain labial veneers in 96patients, fitted up to five yearspreviously in two teaching hospitals.Increased problems and failurerates were associated with veneerswhere inappropriate luting agentswere employed, that is, lutingcements not dedicated to veneercementation. In studying over 1,544amalgam <strong>restorations</strong>, 1,213 <strong>restorations</strong>remained after 15 fifteenyears. The conclusions from thisstudy was that the type <strong>of</strong> amalgamalloy used had no associationwith restoration survival 108,109 . VanNoort and Davis 142 observed in afive-year prospective study thesurvival <strong>of</strong> 2,399 Class III and 1,093Class V chemically-activated anteriorcomposite resin <strong>restorations</strong> in26 general <strong>dental</strong> practices, that thedifferences in clinical performancesbetween six materials was small. Ina cross-sectional study, 75 privatepractitioners evaluated 1,147 twoto-fouryears old anterior <strong>restorations</strong><strong>of</strong> 25 different materialsaccording to the CDA system. Withthe exception <strong>of</strong> one compositeresin, no obvious differences in thequality <strong>of</strong> <strong>dental</strong> <strong>restorations</strong> wereobserved 33 .Operator factorsAmong various clinical factorsaffecting restoration performance,an operator association is frequentlydetected in multicentre and crosssectionalstudies (Table 1). Controlledclinical studies are usually designedto avoid such operator effects, anddifferent research groups haveemployed various strategies (seesection on statistical issues) tocontrol these effects.Most <strong>of</strong> the relevant papers<strong>of</strong>fer no explanation <strong>of</strong> theobserved variation in operatoreffect, although some authors stressthe necessity <strong>of</strong> specific training for<strong>dental</strong> personnel using new materials.It is also possible that anindirect patient association may haveinfluenced an apparent operatorassociation. Whether the experience<strong>of</strong> the operator can be associatedwith technical excellence is uncertain.For example, Hawthorne andSmales 150 reported that the survivalfor composite resin <strong>restorations</strong>was best for the most recent graduates.By contrast, Shaini et al. 151commented that the poorestresults were seen in relation to theinexperienced operators. Of course,appropriate clinical training andexperience are necessary prerequisitesfor favourable clinical outcomes.


132I inlays was lower than that for theother types <strong>of</strong> inlays, that is two-,three- and four-surface inlays.In study <strong>of</strong> over 1,544 amalgam<strong>restorations</strong>, 1,213 <strong>restorations</strong>remained after fifteen years. Itappeared as if the three-surface<strong>restorations</strong> had less favourableclinical outcomes compared totwo-surface <strong>restorations</strong> after 15years. Also, the provision <strong>of</strong> a 90 ocavosurface angle combined witha cavity wall finish reduced the risk<strong>of</strong> failure <strong>of</strong> amalgam <strong>restorations</strong>,compared to larger cavosurfaceangles 108,109 .Prati et al. 161 reported three-yeardata on 116 Class III and V polyacid-modifiedcomposite resin(‘compomer’) <strong>restorations</strong>. Theyfound no statistical differences withrespect to the USPHS criteriabetween the Class III and V <strong>restorations</strong>.In a cross-sectional appraisal <strong>of</strong>520 cast <strong>restorations</strong> in 56 patientsmade between one to 40 yearspreviously, the <strong>restorations</strong> includingmore than two surfaces wereassociated with less favourableoutcomes <strong>of</strong> quality and survival 162 .The outcome after three years<strong>of</strong> 446 <strong>restorations</strong> placed byone dentist and two <strong>dental</strong> nursesin 282 patients using the ARTtechnique was reported byPhantumvanit et al. 148 . The survivalwas lower for occlusal surface<strong>restorations</strong> compared to those inother surfaces.Friedl et al. 163,164 carried out across-sectional study in which 102dentists provided informationabout 3,375 composite resin and5,240 amalgam <strong>restorations</strong>. Thefailed <strong>restorations</strong> with foursurfaces had a lower median agecompared to the other types <strong>of</strong><strong>restorations</strong>.Jokstad et al. 165 reported on across-sectional study <strong>of</strong> 8,310<strong>restorations</strong> a marked associationbetween the age <strong>of</strong> the <strong>restorations</strong>and both the types and the size <strong>of</strong>the <strong>restorations</strong>.Data from a five-year prospectivestudy <strong>of</strong> the <strong>of</strong> 2,399 Class IIIand 1,093 Class V chemically-activatedanterior composite resin<strong>restorations</strong> assembled from 26general <strong>dental</strong> practices indicatedthat the overall probability <strong>of</strong>survival at five years was 10 percent higher for Class V <strong>restorations</strong>than for Class III <strong>restorations</strong> 142 .Smales and Gerke 166 evaluated700 anterior composite resin<strong>restorations</strong> over four years. Significantlymore failures occurred inClass IV and V preparations. Ofall failures, 81 per cent were fromClass V preparations, which mayreflect undue reliance on dentinebondingsystems for restorationretention in premolar non-cariouscervical lesions.Fritz et al. 167 reported the longtermoutcome <strong>of</strong> 2,717 cast<strong>restorations</strong> provided for 548patients during 1960–1989. Relativelyminor differences in the 15year survival were noted for foursizes <strong>of</strong> casts: single surface (65 percent), two-surface (60 per cent),three-surface (68 per cent) andinlays and onlays (70 per cent).Investigations published before1992 on the association betweenClass I and Class II cavity preparationsfor amalgam and <strong>restorations</strong>urvival has been reported in aprevious paper and will not bediscussed further 113 .Wilson and Norman 147 reportedfive-year findings <strong>of</strong> an 11-centretrial <strong>of</strong> a posterior composite resin.The findings were based on datacollected from 649 (68 per cent)<strong>of</strong> the 958 <strong>restorations</strong> originallyplaced. Chi-square analyses indicatedthat <strong>of</strong> the independentvariables investigated, size <strong>of</strong>restoration had the greatest associationwith clinical performance.In a study <strong>of</strong> 950 anteriorcomposite resin <strong>restorations</strong> oversixteen years there were significantlymore failures with the Class IVcompared to the Class III <strong>restorations</strong>153 .Bentley and Drake 168 reported astudy <strong>of</strong> 1,207 <strong>restorations</strong> placedby students in 70 patients. Singlesurface<strong>restorations</strong> lasted longerthan multi-surface <strong>restorations</strong>.Moreover, multisurface <strong>restorations</strong>including the occlusal surfacesurvived significantly longer thandid those including the facial orlingual surface.Material handling andproceduresFactors associated with materialhandling and procedures that mayaffect the incidence <strong>of</strong> margin failures<strong>of</strong> amalgam <strong>restorations</strong>include trituration time 106 , use <strong>of</strong>rubber dam 169,170 , condensationtechniques 171 , and the carving,burnishing and polishing techniques172 . The effect <strong>of</strong> burnishingamalgam restoration margins isdifficult to estimate, and nearlyimpossible to quantify since the‘surface treatment’ is influenced byfactors such as burnishing load,direction <strong>of</strong> the strokes, number<strong>of</strong> strokes, beginning time aftertrituration and the size <strong>of</strong> theburnisher 172 .To what extent cavity varnishesand their thicknesses promotemarginal failures is unknown. It isconceivable that some varnishesmay be incorporated into the amalgamalong the margin, and therebyreduce the strength in these areas.Thus, there is a theoretical possibilitythat the type or amount <strong>of</strong>varnish may be related to marginfracture. However, clinical data onsuch a relationship is sparse, andexisting data are not conclusive. Inone longitudinal study it wasrevealed that application <strong>of</strong> avarnish or silver suspension did notinfluence the risk <strong>of</strong> long-termrestoration failure 107–109 .Letzel et al. 171 assessed the type<strong>of</strong> condensation instrument, togetherwith patient and operator on theperformance <strong>of</strong> a single amalgamalloy over 2.5 years. The authorsreported an association betweenfailure and the patient and theoperator, but there were too fewfailures to establish an associationwith condensation instrument.However, the authors did not indicatehow ‘failure’ was assessed.International Dental Journal (2001) Vol. 51/No.3


133IsolationThe debate about the necessity <strong>of</strong>using rubber dam in operativedentistry has been ongoing withvariable intensity for many decades.In this context it should perhaps beemphasised that working withoutrubber dam does not necessarilyallow saliva contamination duringoperative procedures.Raskin et al. 157 reported a tenyearlongitudinal study <strong>of</strong> 100Class I and II composite resin<strong>restorations</strong> in a group <strong>of</strong> selected,predominantly young patientsunder highly controlled conditions.The method <strong>of</strong> isolation was notfound to significantly influence clinicalperformance and survival.Dunne and Millar 141 reportedthe evaluation <strong>of</strong> 315 porcelainlabial veneers in 96 patients, fittedup to 63 months previously in twoteaching hospitals. The use <strong>of</strong> rubberdam could not be associatedwith survival.The consequence <strong>of</strong> using eitherrubber dam or cotton roll isolationon clinical deterioration wasreported by Smales 169,170 . In onestudy, 546 polished amalgam and148 anterior enamel-bonded compositeresin <strong>restorations</strong> were evaluatedover periods <strong>of</strong> up to 15 years. Heconcluded that although a statisticallysignificant difference wasfound between the two isolationmethods for marginal fracture <strong>of</strong>the composite resins, the clinicalrelevance <strong>of</strong> this difference wasquestionable. In the second paperthe survival <strong>of</strong> the <strong>restorations</strong> wasrelated to the possible influence <strong>of</strong>six other clinical parameters. Therewere no clinically significant differencespresent in the initial highquality <strong>of</strong> the <strong>restorations</strong>, or intheir later survivals, which could bedirectly related to the use <strong>of</strong>rubber dam.In a survey where the clinicalhandling properties <strong>of</strong> glassionomercements were addressed,Knibbs and Plant 173 attributed themain cause <strong>of</strong> unsatisfactory <strong>restorations</strong>in deciduous teeth made by17 general <strong>dental</strong> practitioners topoor handling <strong>of</strong> the material, principallyby moisture contamination.Van Dijken and Horstedt 174assessed 35 patients who receivedone hybrid and one micr<strong>of</strong>illedcomposite resin restoration placedin anterior teeth with and withoutrubber dam. After one year themarginal adaptation was investi-Table 2Clinical studies reporting a relationship between restoration performance and intraoral locationReference Restorations/ Restoration types Obs. period General performancepatients (years)LongitudinalKöhler et al., 2000 87 63/45 Posterior composite 5 No statistically significant difference betweenpremolars and molars, and between maxillary andmandibular teethLundin and Koch 2000 155 117/65 Posterior composite 10 Restorations in premolars had a higher failurerate than in molarsPyk and Mejare, 1999 158 242/142 Glass cermets – tunnel 3.5 Failure occurred about five times as <strong>of</strong>ten inmolars as in premolarsRaskin et al., 1999 157 100/ Posterior composite 10 Location not found to influence survivalDonly et al., 1999 176 72/18 Gold cast and composite 7 The acceptable <strong>restorations</strong> were located mainlyinlayin the premolarsPrati et al., 1998 161 116/ Class III and V polyacid- 3 No association with respect to the USPHS criteriamod. compositeand intra-oral locationGeurtsen and Schoeler, 1209/ Class I and II composite 1–4.5 More <strong>restorations</strong> with rating Alpha in premolar1997 177 teeth compared to molar teethGruythuysen et al., 1213/ Class I and II amalgam 15 The type <strong>of</strong> tooth had no association with1996 108 survivalSmales and Gerke, 700/ Class III and V composite 4 More failures occurred in premolar teeth com-1992 166 pared to other locationsJokstad, 1992 29 468/ Class II amalgam 10 No effects <strong>of</strong> intraoral location detectedRetrospectivePelka et al., 1996 162 520/56 Cast 1–40 Molars had less favourable outcomes <strong>of</strong> qualityand survival compared to premolarsDrake et al., 1990 61 1207/70 All types 1–20 No statistically significant differences wereBentley and Drake,observed. Mandibular incisive <strong>restorations</strong> lasted1986 168 longer than maxillaryCross-sectionalMcDaniel et al., 2000 178 706/ Class I and II amalgam ns Mandibular first (36%) and second (20%) molarsaccounted for most fractures among cuspalcoverage<strong>restorations</strong>Jokstad et al., 1994 165 10091/575 All types >10 The restoration age is possibly influenced byintra-oral locationKerschbaum et al., /1841 Fixed prostheses 1–15 An anterior placement and the lower jaw associ-1991 179 ated with a lower survival<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


134Table 3Clinical studies reporting a relationship between restoration performance and gender and age.Reference Restorations/ Restoration types Obs. period General performancepatients (application) (years)LongitudinalPyk and Mejare, 1999 158 242/142 Glass cermet (tunnel) 3.5 Success rate was not related to patient agePlasmans et al., 1998 145 300/ Amalgam (complex) 8.5 Restorations more prone to failure in patients >30yrs old than in younger onesPrati et al., 1998 161 116/ Class III/V polyacid – 3 No association between USPHS criteria andmodified compositepatient age and genderPhantumvanit et al., 1996 148 446/282 Glass-ionomer (ART) 3 No differences between ART <strong>restorations</strong> inchildren and adultsGruythuysen et al., 1996 108 1213/ Class I and II amalgam 15 20% <strong>of</strong> the study <strong>restorations</strong> were replaced inmales, 16% in femalesJokstad, 1992 29 468/ Class II amalgam 10 Survival associated with patient age, but sampleincluded caries-susceptible childrenSmales and Gerke, 1992 166 700/ Class III and V composite 4 Significantly more failures occurred among theelderly patientsRetrospectiveHawthorne and Smales, /100 Amalgam and composite 1–40 Lowest survival rates in the 0–20 and 61+ year1997 150 age groupsPelka et al., 1996 162 520/56 Cast 1–40 No association between patient age and outcomes<strong>of</strong> quality and survivalMahmood and Smales, /1588 All types 1–15 Restoration survival was superior in female1994 152 patientsDunne and Millar, 1993 141 315/ Porcelain (veneer) 5 Patient age and gender could not be associatedwith failureBentley and Drake, 1986 168 1207/70 All types 1–20 Survival less favourable for patients >60 yrs oldcompared to the younger patients. No differencesbetween males and femalesCross-sectionalMjör et al., 2000 36 6761/ Amalgam and composite 1–30 Minor differences noted in longevity betweenmale and female patientsGlantz et al., 1993 180 /77 Fixed prosthesis 1–15 No differences regarding fracture, loss <strong>of</strong>retention and/or <strong>dental</strong> caries between agesubgroupsDawson and Smales, 1992 181 1918/100 Amalgam and composite 1–16 Survival lower in the oldest <strong>of</strong> three age groups.A small gender difference also notedSmales, 1991 153 950/ Anterior composite 1–16 Median survival 7 yrs in age group 60 yrs, 12 yrs in group 21–60 yrs. More failuresseen in the oldest groupSmales, 1991 154 1476/ Amalgam 10 Patient age had a significant association for one<strong>of</strong> five alloysKerschbaum et al., 1991 179 /1841 Fixed prosthesis 1–15 A lower survival <strong>of</strong> fixed prostheses wasassociated with higher age group (especially ifthe patient was older than 70 yrs)Kroeze et al., 1990 182 /600 All types ns The prevalence <strong>of</strong> unsatisfactory <strong>restorations</strong>tended to be higher with increasing agegated and no differences wereobserved.International Dental Journal (2001) Vol. 51/No.3Patient factorsCollins et al. 175 concluded, after aneight-year longitudinal study <strong>of</strong>posterior composite resin <strong>restorations</strong>,that there was evidence toconfirm the importance <strong>of</strong> theinfluence <strong>of</strong> the patient, since many<strong>of</strong> the observed failures occurredamong few patients. However, nodetails were reported that characterisedthese patients.The findings related to anassociation between restorationperformance and patient factors aresummarised in Tables 2 and 3.The majority <strong>of</strong> the studiesdescribe a minor difference. However,as for many <strong>of</strong> the otheralleged associations to clinical factorsone must be aware <strong>of</strong> publicationbias. Moreover, most clinical studieswere designed to address specificclinical problems. Issues such asinfluence <strong>of</strong> operator, patient factorsand intraoral location were usuallycarried out as secondary analyses.It is impossible to know if no suchrelationships are reported becauseno secondary analyses have beencarried out, or if the relationshipswere negative and thus omitted inthe text (Figures 4).Oral environmentReduced salivation and xerostomiaare associated with older patients,side effects <strong>of</strong> drug therapy andcancer treatment. Consequently,


135abcdeFigure 4 (a–e). Old composite resin <strong>restorations</strong> remaining in situ due to patient satisfaction. Other patients, as well as clinicians, might wellconsider the <strong>restorations</strong> ‘unacceptable’.caries risk increases significantly.Wood et al. 183 studied 54 pairs <strong>of</strong>Class V amalgam and glassionomercement <strong>restorations</strong> overtwo years in 36 xerostomic cancerpatients. Survival times were veryshort (8.5 months) for all <strong>restorations</strong>.Among the individuals usingfluoride, 8 per cent <strong>of</strong> the glassionomercement and 100 per cent<strong>of</strong> the amalgam <strong>restorations</strong>survived after six months. For thesubgroup <strong>of</strong> eight non-fluorideusers survivals were approximately100 per cent and 24 per cent. Thus,the degree <strong>of</strong> fluoride use wasassociated with the rate <strong>of</strong> restorationfailure.Caries activity might be expectedto affect the performance <strong>of</strong><strong>restorations</strong>; however, there are fewdata on this aspect. In a longitudinalstudy <strong>of</strong> 242 tunnel <strong>restorations</strong>in 142 individuals, the cumulativeproportion <strong>of</strong> successful <strong>restorations</strong>was 81 per cent after twoyears and 64 per cent after 3.5 years.<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


136Secondary caries caused replacement<strong>of</strong> 50 per cent <strong>of</strong> the <strong>restorations</strong>,but statistical analyses revealed noassociation between caries activityand replacement 158 .Strand et al. 159 observed in a threeyearstudy <strong>of</strong> 161 glass-ionomercermet cement tunnel <strong>restorations</strong>that there were significantly morefailures among patients with a highcaries activity.Restoration survival was stronglyinfluenced by caries activity in aten-year longitudinal study <strong>of</strong> ClassII amalgam <strong>restorations</strong> 29 . The 210patients in the study were dividedinto high, medium and low cariesactivity, depending on the incidence<strong>of</strong> primary and secondary carieslesions during the first eight years<strong>of</strong> a 10 year study <strong>of</strong> amalgam<strong>restorations</strong>. However, some cautionwas expressed in interpreting thesedata, as the study sample includeda group <strong>of</strong> caries-susceptiblechildren.Bentley and Drake 168 reportedon a study <strong>of</strong> 1,207 <strong>restorations</strong>placed by students in 70 patients. Asubset <strong>of</strong> the population (19 percent) with a disproportionatelyhigher failure rate accounted for56 per cent <strong>of</strong> all failed <strong>restorations</strong>.The authors speculated thatit may have been a reflection <strong>of</strong>higher caries activity, but theprecise nature <strong>of</strong> this groupremained uncertain. A subsequentanalysis <strong>of</strong> the study material identifiedonly minor differences <strong>of</strong>salivary risk markers for cariesbetween patients in the high- andlow-failure groups 184 .Patient attendanceHawthorne and Smales 150 relatedsurvival for amalgam and compositeresin <strong>restorations</strong> with patientattendance in a retrospective study<strong>of</strong> five types <strong>of</strong> <strong>restorations</strong> placedby 20 male dentists in 100 adultpatients. There were no significanteffects on restoration survival fromchange <strong>of</strong> dentist, and generally onlyone or two types <strong>of</strong> <strong>restorations</strong>had their survivals influencedsignificantly by frequency <strong>of</strong>patient attendance or experience <strong>of</strong>dentist. Restoration survival was notinfluenced significantly by whether,or not, any replacements weremade by the dentist who placedthe initial <strong>restorations</strong>.Regularly attending patients areprobably more <strong>dental</strong> healthconscious than irregular attenders.Furthermore, at recalls, dentistscorrect minor discrepancies that arebelieved to put the prognosis <strong>of</strong> arestoration at risk. Therefore,improved restoration longevity ismore likely in regular attenders.However, Jokstad et al. 165 did notdetect this difference in a crosssectionalstudy <strong>of</strong> 8,310 <strong>restorations</strong>,where similar restoration ages wererecorded for the regular andirregular attenders.Mahmood and Smales 152 comparedlongevity <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>in selected patients fromdifferent practice environments intwo countries, private practices inPakistan and a <strong>dental</strong> hospital inAustralia. In both countries, <strong>restorations</strong>urvival was significantlyimproved when patients attendedinfrequently for treatment, andwhen the patient routinely changeddentist.In an examination <strong>of</strong> the survival<strong>of</strong> 1,918 <strong>restorations</strong> in an Australianmilitary population, no differencesin survival as a function <strong>of</strong> frequency<strong>of</strong> attendance or frequency <strong>of</strong>change <strong>of</strong> dentists were reported 181 .Kroeze et al. 182 examined, in anational epidemiological survey, therestoration quality <strong>of</strong> 600 dentateadults. The authors reported thatthe restoration quality could berelated to the frequency <strong>of</strong> visits toa dentist.A re-examination <strong>of</strong> 720 dentateScottish residents who had takenpart in a <strong>dental</strong> health survey fiveyears previously suggested superiorrestoration longevity among thepatients who had not changeddentist frequently 185 .Does technical excellencepredict failure?In several longitudinal studies ingeneral practice settings, it has beenobserved that dentists do notnecessarily replace <strong>restorations</strong> evenwhen one or more features havebeen graded as unacceptableaccording to the USPHS/CDAcriteria 140,186–190 . Thus many dentistspractice a treatment philosophywhere the discrepancy is observedrather than effecting an immediateoperative intervention. Interestingly,the same longitudinal studies alsoreveal that <strong>restorations</strong> will performsatisfactorily for many years in spite<strong>of</strong> ‘unacceptable’ USPHS/CDAscorings.One main conclusion from afifteen-year longitudinal study <strong>of</strong>1,213 Class II amalgam <strong>restorations</strong>was that the short-termmarginal performance was not anindication <strong>of</strong> long-term survival, andthat there was a lack <strong>of</strong> a validpredictive parameter 107–109 .In a longitudinal study <strong>of</strong> ClassV composite resins, the authorsreported that it was apparent thatthe results after two years <strong>of</strong>observation could not be used topredict the three-year results 191 .Smales and Webster 188 attemptedto determine the relationshipbetween the deterioration and thelater failure <strong>of</strong> a very large number<strong>of</strong> amalgam and anterior compositeresin <strong>restorations</strong> examined overperiods <strong>of</strong> up to sixteen years.Assessments were made <strong>of</strong> thedeterioration in various characteristics<strong>of</strong> <strong>restorations</strong> that werethought to predict later failure. Foramalgam, there was a significantassociation found between surfacetarnishing and failure. Marginal fractureand marginal staining were notsignificantly associated with any <strong>of</strong>the three failure modes. For thecomposite resins there were significantassociations between thesurface roughness, marginal fractureand colour mismatch. However,surface and marginal staining werenot associated with any <strong>of</strong> the threefailure modes. Many <strong>restorations</strong>assessed as being unsatisfactorycontinued to function on averagefor a further two to three yearsInternational Dental Journal (2001) Vol. 51/No.3


137before being replaced, <strong>of</strong>ten forunrelated reasons.The median function period <strong>of</strong><strong>restorations</strong> can possibly bepredicted by a Weibull distributionfunction. However, Smales et al. 192reported, after applying such atheoretical model on three restorativematerials, that this is problematicif the model includes slowlydeteriorating restoration features.Specific replacementreasonsAllergyGiven the enormous number <strong>of</strong><strong>dental</strong> <strong>restorations</strong> placed worldwide, the incidence <strong>of</strong> adversereactions seems exceedingly small.Researchers have tried to estimatethe population risk <strong>of</strong> adversereactions to materials used indentistry 193 , but the accurateness <strong>of</strong>such estimates <strong>of</strong> risk remainsuncertain. However, there are in allpopulations a minority <strong>of</strong> individualsthat responds negatively to variousextrinsic and intrinsic substances,including biomaterials found to be<strong>of</strong> acceptable biocompatibilityaccording to International Organisationfor Standardisation (ISO)standards. All <strong>dental</strong> restorativeshave the potential to cause adverseeffects, even when used correctly.Higher risks <strong>of</strong> adverse reactionsare present if the material is notproperly stored or handled, forexample, incorrect proportioning,contamination, inadequate polymerisation,date expiration, incorrectstorage temperature and/or humidity.Several comprehensive reviewarticles 194,195 proceedings and consensusstatements on the subject <strong>of</strong>biocompatibility <strong>of</strong> materials canbe found in the <strong>dental</strong> literature.Endodontic complicationsPostoperative sensitivity is anoutcome that is a complex combination<strong>of</strong> the effect <strong>of</strong> the extentand depth <strong>of</strong> the initial carieslesion, cavity preparation depth,period <strong>of</strong> dentine exposure tosaliva and caries, material-to-cavityadaptation, quality and quantity<strong>of</strong> exogenous bacterial products,restoration sealing and variablepatient pain thresholds 196,197 .Post-operative sensitivity aftercementation may be associated witha deformation <strong>of</strong> the abutmentfollowing high pressure or misalignment<strong>of</strong> the casting during cementation.The hypersensitivity resultsfrom fluid movement within thedentinal tubules 198 .Material factorsThe previous belief that pulpcomplications following restorativetreatment were either the consequence<strong>of</strong> insufficient removal <strong>of</strong>bacteria in the dentine or to toxiceffects from the material was challengedapproximately 10 years ago.In a consensus report from 1992,it was stated that much <strong>of</strong> theprevious work on pulpal reactionsto restorative procedures andmaterials had up to then beenflawed because <strong>of</strong> leakage <strong>of</strong>bacteria and their products aroundfilling materials 199 . The general viewtoday is that most restorativematerials do not per se cause pulpdamage as long as they are properlyhandled, but problems willdevelop if the handling proceduresare not followed to ensure optimaladaptation to the cavity walls 200 . Anexception is perhaps glass-ionomercements when applied in a veryclose proximity with the pulp 195 .Poor adaptation between arestoration and remaining toothtissues increases the risk for endodonticcomplications given thepotential leakage <strong>of</strong> detrimentalsubstances. A number <strong>of</strong> highlysophisticated laboratory techniqueshas been developed to measureadaptation, but the association withclinical significance remains uncertain.Thus, quantitative data fromlaboratory leakage studies do notgive sufficient information topredict clinical performance. Theenvironmental conditions in themicrospace between restoration andcavity walls remains unknown, aswell as the interaction mechanismbetween the potentially detrimentalsubstances in the space and thetooth tissues. Even the characteristics<strong>of</strong> the bacteria in, or adjacentto, the gap have not been firmlyestablished. It is clear that thesurface chemistry <strong>of</strong> the materialsignificantly influences the microecologicalenvironment 201 . However,it is unclear if this is due to a localtoxic effect <strong>of</strong> possible componentsreleased from the restorativematerial, or if it is indirectly due toan effect on the initial bi<strong>of</strong>ilmcomposition that is formed on therestoration surface.Anecdotal sources have reportedthat in some practices, alarmingnumbers <strong>of</strong> endodontic procedureshave become necessary because <strong>of</strong>pulp damage after prosthesis cementation202 . Third-party paymentcompanies report that many teethreceiving crowns require endodontictherapy within five years. It isuncertain if this can be related to agradual shift <strong>of</strong> use from conventionalzinc phosphate cements toalternative cements and/or cementationtechniques.Operator factorsCavity designThe remaining dentine thickness isa critical factor in the development<strong>of</strong> pulp damage given the largesurface area <strong>of</strong> open dentinetubules close to the pulp. Dentinetubules may provide diffusionchannels for noxious substances,which diffuse toward the pulpwhere they can activate the immunesystem, provide chemotacticstimuli, cytokine production, andelicit pulpal inflammation 116 . Postoperativehypersensitivity, on theother hand, seems to occur in someindividuals regardless <strong>of</strong> the depth<strong>of</strong> the prepared cavity 203 .The incidence <strong>of</strong> pulpal complicationsfollowing crown and bridgework was discussed by Valderhauget al. 190 in a report describing theresults <strong>of</strong> a longitudinal clinical<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


138study <strong>of</strong> initially 158 fixed prosthesesmade by senior <strong>dental</strong> students25 years previously. The resultsindicated that the frequency <strong>of</strong> pulpdeterioration in association withbridges tends to be related to thesize <strong>of</strong> the prosthesis. It can bespeculated that this can be theeffect <strong>of</strong> biomechanical complexity,including factors such as acomplex alignment <strong>of</strong> preparationswith possible iatrogenic tissueremoval and overtapered abutments;lack <strong>of</strong> acceptable fit in parts<strong>of</strong> the casting; tendency to acceptsmall discrepancies in large, fixedprostheses compared to singlecrowns, and the complications <strong>of</strong>oral hygiene procedures.Periapical complications andvertical root fractures following theplacement <strong>of</strong> <strong>restorations</strong> orcrowns with pulpal or parapulpalposts may be considered as iatrogenic.Cross-sectional studies indicatethat this is perhaps morecommon than is acceptable. Grieveand McAndrew 204 examined radiographically327 post-retained crownsfor length <strong>of</strong> post, length <strong>of</strong>remaining root filling, periapicalcondition, fit and angulation <strong>of</strong>the post and quality <strong>of</strong> root filling.Most root fillings were judged tobe unsatisfactory, and there wasno radiographic evidence <strong>of</strong> anyroot filling in nearly 10 per cent <strong>of</strong>cases.Investigations published before1992 on the possible relationshipbetween Class I and Class II cavitypreparations for amalgam andadverse effects on the pulp hasbeen reported in a previous paper,and will not be discussed further 113 .Material handling andproceduresThe principles for prevention <strong>of</strong>pulpal damage during preparationwith rotating instruments wereoutlined many years ago, and arestill valid today. Key factors arefrictional heat and adequate cooling,excessive dehydration and airblast, and vibration and highspeed 46 .Concerns over possiblemicroleakage and postoperativesensitivity with amalgam <strong>restorations</strong>have led many practitionersto use various varnishes and resincontaininglining, or ‘adhesive’materials. However, others havequestioned the merit <strong>of</strong> amalgambonding, and the results areconflicting 205 .Cases <strong>of</strong> persistent post-operativesensitivity with composite resin<strong>restorations</strong> following total etchingand application <strong>of</strong> some dentineadhesives have been reported. Thisclinical phenomenon can occurdespite careful isolation prior todirect restorative procedures andthe use <strong>of</strong> an incremental build uptechnique, as well as after cementation<strong>of</strong> indirect <strong>restorations</strong> inconjunction with using a compositeresin cement. <strong>Pro</strong>blems regardingpost-operative sensitivity are hypothesisedto be related to a deficientlight-curing source, or incompleteevaporation <strong>of</strong> the primer solventsprior to application <strong>of</strong> the bondingagent. Alternative suggestions aregaps, cracks in enamel related topolymerisation stresses, fracture <strong>of</strong>tooth substance at the restorationcavity interface and polymerisationshrinkage followed by hydraulicforces induced during masticationon the dentinal tubule fluid followingflexure <strong>of</strong> the restoration 195,206 .Patient factorsIntraoral locationIn a study <strong>of</strong> 88 composite resin<strong>restorations</strong> over two years it wasobserved that the postoperativesymptoms were significantly lowerin premolars compared to molars 101 .Borgmeijer et al. 196 reported postoperativesensitivity after placing244 Class II <strong>restorations</strong> <strong>of</strong>composite resin and amalgam. Postoperativesensitivity occurred more<strong>of</strong>ten in the molars than in thepremolars although the differencewas not statistically significant. Theresearch group reported that thisfinding had also been observed inanother study <strong>of</strong> 240 indirect ClassII composite resin inlays and amalgam<strong>restorations</strong> 207 .Age and genderDuring function, secondary andreparative dentine is deposited inthe pulp. At age 55 years, thevolume <strong>of</strong> the pulp is about 20 percent <strong>of</strong> that at age 25, and containsonly 20 per cent <strong>of</strong> the bloodsupply 208 . This suggests that thepulp’s capacity <strong>of</strong> recovery decreaseswith age. However, there are nodata in the literature reporting theincidence <strong>of</strong> endodontic complicationsas a function <strong>of</strong> patient agefollowing restorative therapy.Oral environmentAnecdotal observations suggest thatbruxism may be associated with anincreased risk <strong>of</strong> pulpal complicationsfollowing flexing <strong>of</strong> the<strong>restorations</strong> and gap formation.The clinical evidence for this is poor.In one early in vitro study it wasdemonstrated that composite resin<strong>restorations</strong> placed in third molarsexhibited increased microleakagewhen an antagonist was presentcompared to none 209 . The authorconcluded that bacterial leakagearound <strong>restorations</strong> in cavitiessurrounded by enamel would most<strong>of</strong>ten be the result <strong>of</strong> stress in therestored tooth during occlusion andarticulation. The study has not beenduplicated using modern dentinebonding systems, so it is uncertainto what extent the conclusion isapplicable to newer composite resinmaterials.Does technical excellencepredict failure?A clear relationship between endodonticcomplications and criteriafor technical excellence <strong>of</strong> <strong>dental</strong><strong>restorations</strong> has not been demonstrated.Several review papers havesuggested that such a relationshipexists, but this is substantiatedmostly with laboratory andmicroleakage studies 210 . In one clinicalstudy the radiographic quality<strong>of</strong> the root filling and the appearance<strong>of</strong> the apical one-third <strong>of</strong>International Dental Journal (2001) Vol. 51/No.3


1391,010 endodontically treated teethwas scored 211 . This was related tothe presence <strong>of</strong> open margins,recurrent decay or overhangsdetected in radiographs. By calculatingthe odds ratios <strong>of</strong> periapicalinflammation as a function <strong>of</strong> theroot-filling and restoration qualities,the authors concluded thatrestoration qualities are moreimportant than root-filling quality.Although this finding may give riseto concerns there are many methodologicalissues that can be raised,and scientifically sound researchdesigns are needed to substantiatethe hypothesis. Indeed, a recentidentical study resulted in contradictoryresults 212 .Periodontal problemsThe research focused on <strong>dental</strong><strong>restorations</strong> and periodontal diseaseis a good example <strong>of</strong> how thecriteria for causation as establishedby Hill have been applied successivelyto clarify the relationshipbetween the two factors 213 . Thereis consistent association for severalepidemiological studies between<strong>restorations</strong> with and withoutdiscrepancies and indicators <strong>of</strong> periodontaldisease 214,215 . Strength <strong>of</strong>association and degree <strong>of</strong> exposurehas also been demonstrated.For example, Lang et al. 215 observeda close relationship between aninflammatory response to poormargins and increasing sizes <strong>of</strong> overhangs.Hill’s criteria <strong>of</strong> temporalityhas been verified in numerous studies,showing that the periodontaltissues around restored teeth havemore gingival inflammation thanthe periodontal tissues around intactteeth in intra- and inter-patientcomparisons. The criteria for interventioneffect was demonstratedin a study by Coxhead 216 , whoreported that following removal<strong>of</strong> restoration overhangs on 50patients, the conditions <strong>of</strong> the periodontaltissues improved significantly.A biological plausibility andcoherence <strong>of</strong> results has beenestablished in experimental studieswhere significant changes in themicrobial ecosystem followingintroduction <strong>of</strong> unfavourable characteristics<strong>of</strong> the <strong>restorations</strong> wereobserved 217 . Finally, experimentalevidence <strong>of</strong> a causal relationshiphas been confirmed in numerousanimal studies.Material factorsMultiple studies have compared theperiodontal response to different<strong>dental</strong> materials, but only smalldifferences have been detectedprovided that the restoration surfaceis smooth 218,219 . No studies havebeen located in the literature linkingperiodontal problems to specificphysical properties <strong>of</strong> materials.A speculative suggestion is that<strong>restorations</strong> made from materialswith high creep values will extrudeout <strong>of</strong> the cavity as a result <strong>of</strong>occlusal stress, and thus cause plaqueretention with periodontal diseaseas a consequence. However, noclinical data substantiate thisconcept.Operator factorsCavity designIn general, the proportion <strong>of</strong> <strong>restorations</strong>with poor margins gingivally,correlates with the gingivoaxiallocation, and thus contributes moreto periodontal disease than <strong>restorations</strong>placed away from thegingival sulcus 215,216 . In one clinicalstudy, no improvement in gingivalstatus was noted following theremoval <strong>of</strong> overhangs. It washypothesised that this was due to acorrelation between, on one side,the axial location <strong>of</strong> the restorationmargin, and, on the other side, thedimensions <strong>of</strong> the overhangs andgaps 220 .Investigations published before1992 on the possible relationshipbetween Class I and Class II cavitypreparations for amalgam andadverse effects on the periodontiumhave been reported in aprevious paper, and will not bediscussed further 113 .Material handling andproceduresIndirect <strong>restorations</strong> may beovercontoured or overextended ifthe impression <strong>of</strong> the preparationis deficient. One paper reportedthat there was little evidence thatestablished guidelines for thepreparation <strong>of</strong> teeth for porcelainlaminate veneers were beingapplied in full in general <strong>dental</strong> practices221 . Other papers conclude thatthe general quality <strong>of</strong> impressionsfor crowns received at commercial<strong>dental</strong> laboratories may be a causefor concern 222,223 . Johnson et al. 224reported a wide variation <strong>of</strong> qualitybetween three-unit bridges madefrom different commercial <strong>dental</strong>laboratories. Lack <strong>of</strong> a satisfactoryprescription, representative <strong>of</strong>which may be used by clinicians,was singled out as an importantexplanatory factor.Patient factorsThe most important aetiologicalfactor in periodontal disease is thepresence <strong>of</strong> microbial plaque.Unless the patient can establishplaque control the risk <strong>of</strong> developingperiodontal disease is high,regardless <strong>of</strong> the technical excellence<strong>of</strong> a restoration. The fact thatoral health maintenance is a majorsignificant factor in avoiding periodontitisand caries has beenestablished repeatedly since themid-1970s 225 . Grasso et al. 214concluded after a cross-sectionalstudy including 291 patients thatplaque control measures were probablymore important in reducingperiodontal disease than improvingthe technical excellence <strong>of</strong> the<strong>restorations</strong>.No studies have been identifiedin the literature linking periodontalproblems to restoration propertiesas indirectly influenced by specificpatient factors.Does technical excellencepredict failure?It is difficult to separate the effects<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


140<strong>of</strong> various local aetiological factors,when assessing the associationbetween periodontal disease and<strong>restorations</strong>. Reported restorationparameters include the axiogingivallocation <strong>of</strong> the restoration margin,the location <strong>of</strong> the contact area andthe axial contour <strong>of</strong> the restoration.Other factors are the plaqueretentive ability, chemical state androughness <strong>of</strong> the restorative material,the occurrence and size <strong>of</strong>overhangs and crevices, and thepossible contributing effects <strong>of</strong> arestoration on an adjacent tooth.Adverse effects on the s<strong>of</strong>ttissues have been attributed either toimproper contact areas with foodimpaction 226 or to details such assurface roughness, contour gaps andoverhangs. Within limits, it appearsthat surface roughness does notlead to gingival changes 218,219,227 .There is general consensus thatall factors that enhance theaccumulation <strong>of</strong> plaque promoteperiodontal disease. Therefore,cavity designs that increase theprevalence <strong>of</strong> restoration discrepanciesindirectly cause supportivetissue breakdown. The prevalence<strong>of</strong> gingival restoration margindiscrepancies varies among differentreports. One major reason is the lack<strong>of</strong> common assessment techniquesand a common terminology.AestheticsThe topic <strong>of</strong> aesthetics includesboth the qualities <strong>of</strong> shape andappearance 228 . Shape depends on theoperator’s ability to contour andfinish the surface, as well as thematerial’s wear resistance. Appearancedepends primarily on materialoptical properties <strong>of</strong> colour andtranslucency. Metamerism, (a differencein colour appearance that varieswith the light source) is also acommon concern in aesthetic dentistry.The same concern exists forfluorescence <strong>of</strong> materials and teeth.Many would argue that a highlypolished, anatomically correct castingor amalgam restoration isaesthetically pleasing. When patientsare asked their opinion <strong>of</strong> aesthetics,it comes as little surprise, that toothcoloured<strong>restorations</strong> are preferred.Numerous clinical studies haveconfirmed strong patient acceptance<strong>of</strong> tooth-coloured inlays. Rimmerand Mellor 229 evaluated patients’perceptions <strong>of</strong> different types <strong>of</strong>fixed anterior <strong>restorations</strong>. Respondentsthought that crowns and fixedprostheses with normal marginswere <strong>of</strong> a higher technical standard,and those <strong>restorations</strong> were preferableto crowns with metalmargins. The shade and colour <strong>of</strong>the <strong>restorations</strong> were the mostimportant factors in the participants’assessments.Abrams et al. 230 compared theassessment <strong>of</strong> quality by 117 patientsafter two dentists had assessed their<strong>restorations</strong>. The authors observedthat when the patients and dentists’perceptions <strong>of</strong> the quality <strong>of</strong> the<strong>dental</strong> <strong>restorations</strong> were compared,no relationship existed. It wasconcluded that patients anddentists employ different criteriaand priorities when judging quality<strong>of</strong> <strong>dental</strong> care. The logical view <strong>of</strong>aesthetics <strong>of</strong> <strong>dental</strong> <strong>restorations</strong> isrelated to the patient perspective,notably to what extent do <strong>dental</strong><strong>restorations</strong> deviate from the appearance<strong>of</strong> sound teeth? (Figure 5)Material factorsSeveral tests have been devised toscreen materials at risk for bulkdiscolouration. These tests employhigh intensity light sources and/orliquids with high chromaticity toevaluate colour stability. No testsexist that correlate laboratory findingswith clinical observations <strong>of</strong>discolouration. The complex eventsproducing restoration discolourationin the oral environment arepoorly understood. Bulk andmarginal discolouration variesconsiderably among different types<strong>of</strong> <strong>dental</strong> materials, and within<strong>dental</strong> material groups such ascomposite resins 189,231,232 and glassionomercements 233 .In general, there is consensusthat amongst the direct restorativematerials, the composite resins havethe best long term clinical performanceregarding aesthetics 234–236 . Somedifferences among the compositeresins have also been reported:• Among conventional chemicallycured composite resins, <strong>restorations</strong>with macr<strong>of</strong>illers discolourmore over time comparedto the same composite resincontaining micr<strong>of</strong>illers 189 .• Chemically cured composite resinsdiscolour more than light-curedover time, probably because <strong>of</strong>different polymerisation initiators232 .• Most restorative materials increaseopacity and lightness after aperiod intraorally followingwater absorption, but this variesconsiderably between variousproducts 237 .Operator factorsMaterial handling andproceduresThe aesthetic limits <strong>of</strong> <strong>restorations</strong>in anterior teeth are determinedmainly by :• The size and nature <strong>of</strong> the lesion• The characteristics <strong>of</strong> the fillingmaterials• The technique <strong>of</strong> application• The age <strong>of</strong> the filling• The oral environment 231 .It is self-evident that an optimalmaterial handling and restorationprocess should be followed toensure a satisfactorily aesthetic result.Numerous papers have detailedtechniques necessary to create aestheticallysatisfactorily <strong>restorations</strong>focusing on variables such as longbevels, polishing, anatomic andsurface sharpening, multi-layeringtechniques, sufficient veneering andapplication <strong>of</strong> subsurface tints. Noclinical studies, however, have testedthe relative importance <strong>of</strong> these variousprocedures on the aesthetic outcomesin a long-term perspective.Patient factorsOral environmentIn a clinical study over 18 monthsInternational Dental Journal (2001) Vol. 51/No.3


141abcdeFigure 5 (a–e). ‘<strong>Quality</strong>’ has a different meaning for different individuals, patients and dentists alike.<strong>of</strong> composite veneers and artificialresin teeth the effects <strong>of</strong> consumptionbehaviour, such as c<strong>of</strong>fee, teaand smoking, and cleaning habitson discolouration was evaluated.No significant influence on discolourationcould be detected 238 .Qvist and Strom 88 observed 52pairs <strong>of</strong> Class III composite resinsover 11 years. Surface discolourationwas most <strong>of</strong>ten recordedamong smokers. Significant correlationswere also found betweenthe patients’ consumption <strong>of</strong> alcoholicbeverages and bulk andsurface discolouration <strong>of</strong> the <strong>restorations</strong>.On the basis <strong>of</strong> a clinical studyit has been suggested that local oralhygiene may play a role in theextent <strong>of</strong> surface staining 239 .Does technical excellencepredict failure?Although the initial aesthetics <strong>of</strong><strong>restorations</strong> <strong>of</strong> tooth-colouredmaterials can be outstanding, a lastingoutcome requires a material thathas a high proportion <strong>of</strong> polymer-<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


142ised matrix. Restorations with aninsufficiently polymerised matrixwill discolour faster than those witha well-polymerised matrix, and forlight-cured materials this depends,among other things, on an acceptablelight intensity. There are a largenumber <strong>of</strong> variables (instrumentation,manipulative, restorative) thatalso influence the degree <strong>of</strong> conversion<strong>of</strong> monomer to polymer.Rarely are <strong>restorations</strong> well cured,which is partly because manydentists use light-curing units withlow light intensities 240 . The intensityoutput from a lamp in a lightcuringunit deteriorates over time,and the minimum acceptable lightintensity level is 300mW/cm 2 .Unfortunately, there seems to be alow awareness among dentists <strong>of</strong>the need for maintenance and regularchecking <strong>of</strong> light intensity aspart <strong>of</strong> a quality managementprogramme 241 .Material deteriorationMaterial deterioration includes bulkand marginal fracture, as well asexcessive wear or dissolution <strong>of</strong>the material. Excessive dissolutionis seldom seen when materialswhich comply with ADA or ISOstandards are correctly used. Inparticular, the presence <strong>of</strong> salivaduring material placement may havea strong negative effect on theresistance to deterioration, forexample glass-ionomer cement 242 .Excessive wear <strong>of</strong> luting cementsmay occur if the cement margin iswide. Modern composite resincements seem to resist wear betterthan conventional cements 243 . Cementsmay also begin to disintegrateunder luted <strong>restorations</strong> duringdeformation <strong>of</strong> the restoration,initiating and propagating cracksleading to cement fracture.Bulk fracture is a common reasonfor restoration replacement, and itis <strong>of</strong>ten associated with caries.Caries may either have precededthe fracture, or have developedrapidly after fracture if a remnant<strong>of</strong> a broken restoration remains inthe cavity 244 . No reference in theliterature has been found as to estimates<strong>of</strong> the incidence <strong>of</strong> the twooccurrences.Material factorsMjör 245 compared reasons for replacement<strong>of</strong> <strong>restorations</strong> with thosefrom another study recorded 16years previously. The proportion<strong>of</strong> replacements <strong>of</strong> amalgam <strong>restorations</strong>due to bulk fracture hadremained much the same over theperiod, which suggest little effect<strong>of</strong> the improvements in amalgamalloy compositions. On the otherhand, a significant relationshipbetween the zinc and coppercontents <strong>of</strong> amalgam alloys andbulk fractures was reported byLetzel et al. 69 on the basis <strong>of</strong> alongitudinal study over 15 years.The authors attributed this to thesuperior corrosion resistance <strong>of</strong> thenon-gamma-2 (high-copper) amalgamscompared to conventionalalloy compositions.For composite resin <strong>restorations</strong>there has been a notable decreasein the relative frequency <strong>of</strong> replacementsas a result <strong>of</strong> degradationand wear, and an increase in thereplacements following bulk andmarginal fractures, which is attributedto changes <strong>of</strong> materialcomposition 245 .Tyas 246 reported three-year observations<strong>of</strong> 102 Class IV <strong>restorations</strong><strong>of</strong> four composite resins.Significant correlations (P < 0.01)were found between surface chipping/bulkfracture and fracturetoughness, elastic modulus andtensile strength. Moreover, therewas a trend towards an associationbetween incisal wear and both elasticmodulus and inherent flaw size.Operator factorsIt is difficult to assess the influence<strong>of</strong> important operator factors suchas poor cavity adaptation, extent<strong>of</strong> porosities and extent <strong>of</strong> contamination<strong>of</strong> the material duringhandling on the restoration. It canbe assumed this occurs at leastoccasionally with a possible detrimentaleffect on restoration strength.To what extent this acceleratesmaterial deterioration remainsunknown.Of the 15 studies inclu-ded ina systematic review <strong>of</strong> CEREC<strong>restorations</strong>, comparable results interms <strong>of</strong> fracture rates were seen instudies undertaken in general practiceand university environments 15 .Malament and Socransky 247observed 1,444 <strong>restorations</strong> madefrom Dicor glass-ceramic over 14years, and found no significantdifference between the bulk fracture<strong>of</strong> inlay and onlay <strong>restorations</strong>.The fracture incidence improvedsignificantly when <strong>restorations</strong> wereacid-etched before luting. Therewas no significant differencebetween acid-etched Dicor <strong>restorations</strong>that were placed onshoulder or chamfer preparations,nor did the thickness <strong>of</strong> the restorationmeasured at the mid-axialpoint <strong>of</strong> each surface relate to fractureincidence.Letzel et al. 69 reported a significantassociation between amalgamalloys and bulk fractures over anobservation period <strong>of</strong> five to 15years. Although not much wascommented on in their report, atable showed a marked associationbetween operator and incidence <strong>of</strong>bulk fracture.Lemmens et al. 248 reported ananalysis <strong>of</strong> 176 fractured amalgam<strong>restorations</strong> and concluded thatthere was a statistically significantinfluence <strong>of</strong> the dentist on the incidence<strong>of</strong> bulk fracture.Cavity designWilson et al. 146 reported on a fivecentrestudy for the five-yearoutcome <strong>of</strong> 232 <strong>restorations</strong>. Large<strong>restorations</strong> consistently deterioratedmore than moderate-sizedones, with respect to class <strong>of</strong>restoration or type <strong>of</strong> toothrestored. In a cross-sectional study<strong>of</strong> approximately 2,500 amalgam<strong>restorations</strong> bulk fractures wereInternational Dental Journal (2001) Vol. 51/No.3


143most prevalent in fillings with threeor four surfaces 163 .In a three-year study <strong>of</strong> 438Class I <strong>restorations</strong> <strong>of</strong> glassionomercement, composite resinand amalgam, loss <strong>of</strong> materialand surface cracking or crazingappeared to a greater extent in largeconventional preparations, andespecially among glass-ionomercement <strong>restorations</strong> 249 .The possible relationship betweenbulk fracture risk and Class I andClass II cavity preparations foramalgam has been reported in aprevious paper, and will not bediscussed further 113 .Patient factorsIntraoral locationMalament and Socransky 245 foundfewer fractures <strong>of</strong> Dicor <strong>restorations</strong>in female than in malepatients. The highest fracture levelwas observed in second molars.Mjör and Jokstad 187 examinedthe clinical performance <strong>of</strong> 274amalgam, glass-ionomer cermetcement and composite resin <strong>restorations</strong>over five years in small ClassII cavities. The majority <strong>of</strong> the bulkfractures, which were mostly inglass-ionomer cermet cement<strong>restorations</strong>, were located in theupper molars.In a 10 year longitudinal study<strong>of</strong> 468 amalgam <strong>restorations</strong>,Jokstad 29 reported findings thatcontrasted with the results byLemmens et al. 243 . Only one <strong>of</strong> 27fractured <strong>restorations</strong> was locatedin lower premolars. No effect <strong>of</strong>the intra-oral location on bulkfracture was observed.On the basis <strong>of</strong> longitudinalstudies over seven years, including176 fractured amalgam <strong>restorations</strong>,it was suggested that <strong>restorations</strong>in the mandibular teeth and especiallyin the premolars were verysusceptible to bulk fracture 248 .Oral environmentWilson et al. 146 reported that thepresence <strong>of</strong> occlusal contacts had asignificant effect on deterioration<strong>of</strong> occlusal marginal adaptationover five years. This appearedgreatest in the large Class I andboth small and large Class II amalgam<strong>restorations</strong> in molars. Thisfinding led the authors to suggestthat future longitudinal studiesshould include assessment <strong>of</strong>occlusal function, diet and chewingpatterns.In a longitudinal clinical study<strong>of</strong> ceramic inlays over three yearsÅberg et al. 251 reported that <strong>of</strong>the fractured inlays, two-thirdsoccurred in patients with signs <strong>of</strong>bruxism.In a retrospective study after10–15 years on the quality <strong>of</strong> 793<strong>restorations</strong>, bulk fractures in amalgam<strong>restorations</strong> were recordedprimarily in patients with severebruxism 170 . The levels <strong>of</strong> oral healthand smoking were also included inthe analyses, but no influence <strong>of</strong>either was found.Klausner et al. 252 recorded thereasons for replacements <strong>of</strong> <strong>restorations</strong>.For bulk fracture, 43 percent <strong>of</strong> <strong>restorations</strong> were 10 years<strong>of</strong> age or older, while 80 per centwere older than four years. Theauthors commented that if faultyocclusion or thin pulpal-occlusalsections <strong>of</strong> amalgam were the principalreasons for isthmus fracture,then these fractures should havebecome evident at an earlier time.Does technical excellencepredict failure?Few clinical studies have addressedthe correlation between materialdeterioration and duration <strong>of</strong>clinical service. Early occlusalmarginal fractures, may 69,106,248 ormay not 188,253,254 correlate withfurther material deterioration.Jokstad 29 observed 468 amalgam<strong>restorations</strong> <strong>of</strong> five alloys placed in210 patients for 10 years. Marginalfractures after relatively short clinicalservice were associated with laterbulk fracture.Wear is not linear over time. Ithas therefore been suggested thatit is misleading to calculate anddescribe wear levels in terms <strong>of</strong>micrometre/year 255 . It is probablethat wear decreases over time,because the more wear-resistantadjacent enamel surface protects theremaining material surface to anincreasing extent 256 .CariesThere is no reason to considersecondary caries as any differentfrom primary caries 257 . It is a localiseddisease caused by a localaccumulation <strong>of</strong> mechanicallyundisturbed bacterial biomass.Several facts should be examinedin this regard. First, even whenthere is a very close adaptationbetween a restoration and tooth,there is still more than enough spacefor bacterial ingrowth. Second, thereis little evidence <strong>of</strong> ‘undetectablemicroleakage’ causing secondarycaries. Third, most papers havereported only weak evidence <strong>of</strong> arelationship between marginaldiscrepancies and secondary caries.Fourth, in spite <strong>of</strong> hundreds <strong>of</strong>laboratory microleakage studies, nocorrelation with secondary carieshas been established. Fifth, groundsections <strong>of</strong> restored teeth withsecondary caries <strong>of</strong>ten revealsubsurface lesions unrelated to thecavity wall. Finally, some clinicaldata suggest that the occurrence <strong>of</strong>secondary caries is a localisedphenomenon related to the conditionsfor evolution <strong>of</strong> cariogenicplaque, rather than a universalattack along the entire interfacebetween tooth and restoration 257 .Thus, secondary caries maydevelop in the presence <strong>of</strong> cariogenicplaque, but will neverdevelop if cariogenic plaque isabsent regardless <strong>of</strong> the technicalexcellence <strong>of</strong> the restoration. Adiscussion <strong>of</strong> which restorationdetail constitutes a major or aminor ‘focus’ <strong>of</strong> plaque retentionappears from this aspect to be anacademic discussion. It is thepatient’s oral hygiene habits that willdetermine if caries develops, not<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


144whether the restoration can beconsidered as ‘excellent’, ‘adequate’or ‘deteriorated’.Material factorsThe increased popularity <strong>of</strong> restorativematerials that release fluorideis in part explained by the beliefthat secondary caries can somehowbe prevented by incorporating thiscomponent. The anticariogenicproperties <strong>of</strong> glass-ionomer cement<strong>restorations</strong> are not strongly substantiatedby clinical investigations 13 .On the other hand, the lack <strong>of</strong>strong evidence <strong>of</strong> an anticariogenicpotential may stem fromclinical studies conducted in academicenvironments on selected patientswith minimal caries risk instead <strong>of</strong>in ‘real-life’ general practices.In a longitudinal study <strong>of</strong> 274large Class II open-sandwich resinmodifiedglass-ionomer cement<strong>restorations</strong> over three years, nosecondary caries was noted, despitea large number <strong>of</strong> participatingpatients with high caries risk 258 . Inthese patients, a far higher cariesfrequency around the other restorativeswas recorded, leading theauthors to suggest a possibleanticariogenic effect <strong>of</strong> this material.Operator factorsSeveral textbooks advise thatbecause <strong>of</strong> polymerisation shrinkage,the location <strong>of</strong> the gingivalmargin for posterior compositeresin <strong>restorations</strong> should be placedat least one millimetre, when possible,from the enamel-cementmargin. Several cross-sectionalstudies imply that this rule is notfollowed by general practitioners.However, it has not been possibleto identify in the literature any clinicalstudies that have associated thischaracteristic <strong>of</strong> the cavity designwith the development <strong>of</strong> secondarycaries.The possible relationship betweensecondary caries risk and Class Iand Class II cavity preparations foramalgam has been reported in aInternational Dental Journal (2001) Vol. 51/No.3previous paper, and will not bediscussed further 113 .Patient factorsIn a clinical evaluation <strong>of</strong> 63 ClassII composite resin <strong>restorations</strong> overfive years, 8 <strong>of</strong> the 11 patients with<strong>restorations</strong> that failed because <strong>of</strong>caries and marginal defects hadhigher counts <strong>of</strong> mutans streptococciat baseline compared to theremaining patients. This led theauthors to suggest that the cariesactivity should be managed to avoidfuture secondary caries 87 .In a study <strong>of</strong> 4,294 childrenobserved over three years, the incidence<strong>of</strong> secondary caries be associatedwith some oral hygieneparameters 259 . Water rinsing afterbrushing and the use <strong>of</strong> a beakerfor rinsing was associated withsecondary caries. Also, subjects whobrushed less <strong>of</strong>ten than twice a daydeveloped more secondary cariesthan the others did.Clarkson and Worthington 260reported that in a group <strong>of</strong> 270adults an association was seenbetween caries, most commonlysecondary caries, and attendance.The irregular attenders experiencedmore caries than did the regularattenders.van Dijken 261 has carried outseveral longitudinal clinical studieswhere part <strong>of</strong> the study design hasbeen to classify the patient accordingto few or many caries riskfactors, based on the net effect <strong>of</strong>microbial counts, oral hygiene, salivaryflow levels, buffer values, andfermentable carbohydrate intake 32 .However, the difference betweenthese two groups has only beenreported in two studies. In a sixyear longitudinal study <strong>of</strong> 150tooth-coloured <strong>restorations</strong>, all<strong>restorations</strong> which subsequentlydeveloped secondary caries (n=7),except for one, were from the highcaries-riskgroup 261 . In another sixyear study <strong>of</strong> seven anteriorcomposite resins, a markedly higherincrement <strong>of</strong> caries was recordedamong the patients with manycaries risk factors 32 .In a 10 year longitudinal study<strong>of</strong> 468 amalgam <strong>restorations</strong>, themost important factor that couldbe associated with the development<strong>of</strong> secondary caries was thepatient’s yearly DFT increment 29 .Does technical excellencepredict failure?Roulet 104 stated categorically thatmarginal integrity is an importantparameter for restoration longevity,since recurrent caries and pulpaldisease is associated with marginalgaps. However, the gap size per semay not play any role in secondarycaries initiation. What is <strong>of</strong> importanceis whether it promotesformation <strong>of</strong> cariogenic plaque,which encompasses additional variablesbesides just gap size. It istheoretically possible that the quantityand quality <strong>of</strong> the plaqueformation around <strong>restorations</strong> maybe a better prognostic marker <strong>of</strong>restoration longevity than variouscriteria <strong>of</strong> technical excellence 201 .The alleged correlation betweenmarginal fractures and recurrentcaries is controversial. Two factorsshould be considered in this context.What is the association between thelocation <strong>of</strong> the defect and location<strong>of</strong> secondary caries? What is therelationship between the size <strong>of</strong> thedefects and secondary caries? Noreports have been identified thatdemonstrate a correlation betweenocclusal discrepancies and marginaladaptation on the approximalsurfaces, which are the areas wheresecondary caries lesions prevail.Therefore, it is difficult to understandhow marginal fractures onthe occlusal surface can be relatedto a higher risk <strong>of</strong> secondary caries.While some authors reportassociations between marginal fractureand secondary caries 262 , othersdo not 263,264 . Other laboratoryexperiments suggest that a correlationdoes not seem to exist betweenthe size <strong>of</strong> the crevice and secondarycaries 265 , or describe only acorrelation in extremely cariogenicenvironments 266 .


145Staining around both amalgam 267and tooth-coloured <strong>restorations</strong> 268is considered unreliable in the diagnosis<strong>of</strong> active recurrent caries.Hewlett et al. 269 reported that 86per cent <strong>of</strong> 822 <strong>restorations</strong> withmarginal defects revealed nosecondary caries on radiographs.This led the author to suggest thatthe replacement <strong>of</strong> all <strong>restorations</strong>with defects related to a perceivedrisk <strong>of</strong> secondary caries wouldconstitute overtreatment.After observing 468 amalgam<strong>restorations</strong> <strong>of</strong> five alloys placed in210 patients for ten years, Jokstad 29reported that those <strong>restorations</strong>with early marginal fractures couldnot be correlated with later development<strong>of</strong> secondary caries.In an observational longitudinalstudy, an increased prevalence <strong>of</strong>secondary caries was recorded inthe <strong>restorations</strong> with the poorestmargin fracture scores 270 .A longitudinal clinical study over10 years demonstrated no differencesin secondary caries levelsbetween a spherical amalgam alloyand a non-gamma-2 alloy, despitedifferences in marginal deteriorationduring the first years 271 .Goldberg et al. 272 examined1,556 <strong>restorations</strong> in 87 patients ina cross-sectional study. The prevalence<strong>of</strong> secondary caries wascorrelated with the marginal fracturescores and indices <strong>of</strong> thepatients’ oral health. Using loglinearanalyses, the investigatorssuggested that there was a significantrelationship between thesethree factors.Tooth fractureTooth fractures include cusp fractures,dentine cracks, incompletedentine fractures, crack lines, andcracked tooth syndrome. Some usethe term ‘infraction’ when the crackline is limited only to the enamel.There is general consensus that arestored tooth is stronger than anon-restored tooth with caries, butthat a tooth with an intracoronalrestoration is weaker than an intacttooth regardless <strong>of</strong> material. Theprocesses involved regarding theway in which tooth strength isassociated with choice <strong>of</strong> restorativematerial, adaptation and themicrostructural relationship at thetooth-material interface are controversial.This can be explained bythe fact that the incidence <strong>of</strong> toothfractures is relatively low, whichimpedes the execution <strong>of</strong> clinicalstudies. Our understanding <strong>of</strong> therelationship between clinical factorsand tooth fractures is therefore toa large extent based on extrapolation<strong>of</strong> case descriptions andlaboratory findings 273–275 .Material factorsThermal dimensional stability, hygroscopicexpansion and setting/polymerisation shrinkage <strong>of</strong> restorativematerials, as well as excessiveloading, have been related tostress build-up in tooth tissues 1,65,73 .The stress will be best tolerated bydentine due to its elasticity, whileinfractions may develop in theenamel. The material-caused stressapplies to restorative materials thatdo not exceed the limits accordingto the existing material test standards.An in vitro investigation hasdemonstrated fractures <strong>of</strong> ceramiccrowns with cores made fromresin-modified glass-ionomer cementmaterials. It was hypothesised thatthe fractures were caused by hygroscopicexpansion 276 . Several manufacturersdo not recommend thisclass <strong>of</strong> material for core builds oras a luting cement for full-ceramic<strong>restorations</strong>.Several papers postulate thatinfractions in enamel and cuspfractures in teeth restored withamalgam are caused by an expansionassociated with heat or withchemical reactions in the alloy, orcorrosion <strong>of</strong> the amalgam. However,there is no clinical documentation<strong>of</strong> such relationship.Furthermore, no standardisationtests have been devised to screenmaterials for this alleged expansion.Operator factorsCast <strong>restorations</strong>, especially inlaysor dowels with improper fit, causestress that increases the risk <strong>of</strong> toothfracture. Strain can also be introducedif high pressure is used ormisalignment <strong>of</strong> the casting occursduring cementation. Also, it hasbeen suggested that tooth preparationusing eccentric or worn bursincrease the risk <strong>of</strong> crack propagationin the tooth.The possible relationships betweentooth fracture risk and Class I andClass II cavity preparations foramalgam have been reported in aprevious paper, and will not bediscussed further 113 .Patient factorsMost tooth fractures originate inthe 30–50 year age group and inteeth with large intracoronal <strong>restorations</strong>or caries. Fractures <strong>of</strong>contralateral teeth are common 275 .All factors that cause high strain ontooth tissue increase the risk <strong>of</strong>crack-line development and cuspfractures. Examples are bruxism,lack <strong>of</strong> occlusal support followingloss <strong>of</strong> teeth, malocclusion, supracontactsor frequent intake <strong>of</strong>coarse foods. Cusp fractures andcrack lines in the posterior teethare most frequently observed onbalancing cusps, which are subjectedto lateral chewing forces (thatis, the lingual cusps in the mandibleand buccal cusps in the maxilla).Ellis et al. 274 reported a metaanalysis<strong>of</strong> the influence <strong>of</strong> patientage on tooth fracture and concludedthat incomplete tooth fracturesare uncommon in students attendingan emergency clinic. It wasnoted that complete fractures mightoccur at any age, while it appearsthat incomplete fractures are associatedwith older age groups.Does technical excellencepredict failure?No papers were found that reportany association between clinicalobservations <strong>of</strong> technical excellence<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


146<strong>of</strong> <strong>restorations</strong> versus the incidence<strong>of</strong> later tooth fractures.Loss <strong>of</strong> restorationLoss <strong>of</strong> entire restoration is limitedmostly to Class V <strong>restorations</strong>,especially when placed withoutcavity preparation. Loss <strong>of</strong> othertypes <strong>of</strong> <strong>restorations</strong> would onlyoccur due to inappropriate selection<strong>of</strong> material, improper preparation<strong>of</strong> the tooth before restoration,violation <strong>of</strong> biomechanical principlesfor designing the restoration,mishandling the material, or acombination <strong>of</strong> these factors.Material factorsDeviations from the manufacturer’sdirection for use may lead todecreased clinical performance. Thishas been reported in a three-yearstudy <strong>of</strong> resin-modified glassionomercements placed in cervicalcavities, where retention wasassociated with surface wettingfollowing variations in the powder/liquid ratio at the time <strong>of</strong> placement277 .Much discussion has taken placeon the importance <strong>of</strong> the modulus<strong>of</strong> elasticity <strong>of</strong> polymeric materials278 . There are diverging views onits relevance, and this confusion typifiesour poor understanding <strong>of</strong>tooth biomechanics. There is alsoan ongoing discussion about whichmaterial is most appropriate fornon-carious cervical lesions. Iscomposite resin, glass-ionomercement, polyacid-modified compositeor resin-modified glass-ionomercement the optimal material <strong>of</strong>choice 279,280 ?Operator factorsAlthough presented as anecdotalevidence, Friedman 281 recounts hisclinical experiences after placing asubstantial number <strong>of</strong> porcelainveneers (n=3,500) over 15 years.The author attributes the three mainfailure reasons (debonding, marginaldiscolouration and adhesive fractures)to poor bonding, which isan indirect indication <strong>of</strong> impropertooth preparation or materialhandling. It has been suggested thatthe effectiveness <strong>of</strong> dentine bonding,and thus the retention <strong>of</strong><strong>restorations</strong>, is influenced byoperator factors such as cavitypreparation relative to the activecaries process, technique for materialapplication, and procedures forpolishing 282 .Cavity designVital dentine is continuously remodellingits microstructure to respondto physiological and pathologicalchanges. Therefore, bonding mayencounter differences such assclerotic dentine, hypersensitivedentine (with open tubules), cariesaffectedareas, superficial dentinewith few tubules, or deep dentinelayers close to the pulp 283 .IsolationCervical composite resin <strong>restorations</strong>placed with isolation eitherwith rubber dam or with cottonrolls suggested no statisticallysignificant differences in retentionafter two years <strong>of</strong> service. Thus,loss <strong>of</strong> retention was not differentwhen a careful cotton roll techniquewas used as a moisturecontrol method, as long as salivacontamination was avoided 174 .Patient factorsGender and ageIt has been suggested that bondingto sclerotic dentine is less reliablethan to young dentine, at least witholder restorative materials 284 . Tworecent clinical studies have refutedthe hypothesis 285,286 . In these studies,lower failure frequencies wereseen in the oldest age groups andthe <strong>restorations</strong> placed in scleroticdentine had an almost equal failurerate compared to the ones placedin non-sclerotic lesions.McCoy et al. 191 suggested, aftera three-year longitudinal study <strong>of</strong>Class V composite resin <strong>restorations</strong>,that a high proportion <strong>of</strong> oldpatients in the study contributed torelatively high proportion <strong>of</strong> loss,as a result <strong>of</strong> changes in the character<strong>of</strong> the dentinal surface.Heymann et al. 287 reported onthe determinants <strong>of</strong> failure <strong>of</strong> bondedcomposite resin <strong>restorations</strong> in noncariouscervical lesions. A strongassociation between patient age andrestoration loss was identified,which was attributed to the greatertooth flexure in older patients, andthe smooth, sclerotic nature <strong>of</strong> ‘old’dentine.Oral environmentIn a recount <strong>of</strong> a clinical experienceafter placing 3,500 porcelainveneers over 15 years, Friedman 281distinguishes between three fracturetypes, and suggests the aetiologicalmechanism for each type. Theauthor ascribed static fracture linesand cohesive fractures to excessiveloading.McCoy et al. 191 noted in a threeyearlongitudinal study <strong>of</strong> Class Vcomposite resin <strong>restorations</strong> thatsome <strong>restorations</strong> were lost fromteeth with marked signs <strong>of</strong> occlusalwear, supporting earlier reports <strong>of</strong>higher loss <strong>of</strong> <strong>restorations</strong> amongbruxers compared to non-bruxers 288 .Does technical excellencepredict failure?There are no known papers thathave reported any associationbetween clinical observations <strong>of</strong>technical excellence <strong>of</strong> <strong>restorations</strong>versus later loss <strong>of</strong> <strong>restorations</strong>.Orthodontic changes andtemporomandibular jointproblemsComposite resins placed during1970–1980 had relatively poorresistance to wear when used in theposterior occlusal segments. In spite<strong>of</strong> this, products were sold in largequantities and used for thispurpose by many general practitioners.Annual wear <strong>of</strong> 30–60mmwas reported, raising concerns forpossible hypereruption <strong>of</strong> opposingteeth and mesial migration <strong>of</strong>teeth distal to those undergoing lossInternational Dental Journal (2001) Vol. 51/No.3


147<strong>of</strong> approximal contact 289 . Despiteconcerns that composite resins wereunsuitable for surfaces exposed toheavy masticatory forces, therewere no reports that extensive useled to tooth misalignments ortemporomandibular joint problemscaused by occlusal changes.The resistance to wear <strong>of</strong>correctly handled modern materialshas now been markedlyimproved. Thus, the risk <strong>of</strong> largegeneralised wear and potentialpathological joint changes andtemporomandibular problems areminor. However, the potentialexists if materials are handledincorrectly, following eating disordersor harmful occupationalenvironments, or if new restorativematerials are applied that havenot been adequately tested forclinical wear.Treatment decisions andtechnical excellenceThe planning <strong>of</strong> restorative treatmentconsists <strong>of</strong> a series <strong>of</strong> interactiveexchanges <strong>of</strong> information betweenthe dentist and the patient. Lack <strong>of</strong>awareness <strong>of</strong> a patient’s complaintsand expectations can lead tounnecessary conflicts, a situationcherished by sensational newsmedia. A common principle inmarketing is ‘do/say the right thingto the right people at the righttime’. This quotation can beimproved for our purposes byrefining the words into the <strong>dental</strong>strategy. Optimal <strong>dental</strong> treatment<strong>of</strong> patients consists <strong>of</strong>: (1) At theright time, (2) <strong>of</strong>fer the right treatment,(3) to the right patient, (4) inthe right way, (5) with the rightresults.The interpretation into <strong>dental</strong>terms is (1) make the correct diagnosisat the outset, (2) explain tothe patient the options and suggestthe best alternative, (3) take intoconsideration the patient’s prioritiesand preferences, (4) carry outthe restorative intervention accordingto correct procedures andmaterial handling to ensure thehighest possible technical excellence,and (5) aim to accomplishthe pre-set objectives that you as adentist and patient have agreed on.The most relevant issue is the evaluation<strong>of</strong> the technical excellence<strong>of</strong> <strong>dental</strong> <strong>restorations</strong> as part <strong>of</strong>the first step: making a correctdiagnosis. Main objectives are torecognise a pathological condition,understand the patient’s problemand identify risk markers <strong>of</strong>progressive oral disease. The technicalexcellence <strong>of</strong> <strong>restorations</strong> is,at this stage, <strong>of</strong> secondary importance.Patient dissatisfactionPatient dissatisfaction with a particularrestoration is easy to detect. Itincludes complaints about pain,aesthetics, surface roughness, contour,detectable margins, supraocclusionor food retention. As long as thecontributing factor is identified andcorrected, there is no problem.However, if the patient’s dissatisfactionis with a restoration thatmeets all criteria for technicalexcellence the situation is morecomplicated. Then patient informationand ethical considerationsmust prevail, tempered by existinglocal or national legislative precedentsor regulations. The patientmust always be advised <strong>of</strong> thepotentially iatrogenic damage associatedwith restoration replacement,such as risk <strong>of</strong> pulp deteriorationand increased cavity dimensions,as well as the possibility <strong>of</strong> noimprovement in spite <strong>of</strong> restorationreplacement.Some studies have attempted tocategorise or distinguish patientcharacteristics versus expectations<strong>of</strong> <strong>dental</strong> treatment. Several taxonomieshave been presented. Håkestamet al. 290 suggests there are threegroups: the aesthetic, the costconscious and the longevityfocussed.Lutz and Krejci 291categorised the patients as in orally‘functional’, ‘presentable’, ‘healthy’,‘beautiful’ and ‘metal-free’ groups.Other studies have applied furtherpatient classifications. Althoughsuch systems tend to appear simplistic,they highlight the fact thatpatients’ personalities, values andpriorities differ considerably. Modernhealth care places great emphasison patient satisfaction as a criterionfor quality. However, thetheoretical principles <strong>of</strong> patient satisfactionare complex and correlatepoorly with criteria for technicalexcellence commonly used bydentists 292 .Estimating risk for oraldiseaseAssessment <strong>of</strong> risk markers <strong>of</strong> oraldisease is detailed in at least oneexcellent textbook on risk and oraldiseases 293 , and within textbookson cariology 46,267,294,295 and periodontology296,297 . Briefly, risk assessmentassociated with susceptibility to oraldisease consists <strong>of</strong> determiningfactors related to the patient’ssocial and medical history, plaquecontrol, saliva and clinical signs <strong>of</strong>disease. Estimating the risk for oraldisease progression can be assessedat the patient, tooth and site levels.At the patient level, the key oraldisease risk markers are the presence<strong>of</strong> a systemic disease, irregular<strong>dental</strong> attendance, prior carieshistory, periodontal problems, plaqueand/or bleeding scores, medicationside effects and saliva quantityand quality. Other risk markers mayrefine decisions about interventions.For example, information on socialdeprivation, active oral disease insiblings or low <strong>dental</strong> IQ andhistory <strong>of</strong> repeated interventionsmay be relevant factors.For periodontal disease, presence<strong>of</strong> residual pockets and cigarettesmoking are additional factors tobe considered when assessing riskwhile additional factors for addressingrisk for future caries are dietaryhabits, frequency <strong>of</strong> sugar intake,availability <strong>of</strong> snacks and use <strong>of</strong>fluorides.At the tooth and site levels, riskfactors include residual periodontalsupport, inflammatory parameters<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


148Table 4 Intervention strategies based on combining a risk assessment for oral disease and the technical excellence <strong>of</strong>the patient’s <strong>restorations</strong>. The codes used in the CDA evaluation system criteria 31 are included to enhance the interpretation<strong>of</strong> the table (see Table 5 for abbreviations)1. Consider consequences <strong>of</strong> monitoring, correcting, removing or replacing restoration in case <strong>of</strong>:CariesCaries along restoration margin (VCAR)Radiographic evidence <strong>of</strong> caries/voidsMarginRetained excess cement (TCEM)Restoration overhang/surplus (TOV)OtherTooth structure is fractured (VTF)Mobile restoration (VMO)Superficial or penetrating fracture line (VFR)Restoration is partially or in toto missing (VMIS)Evoked pain during clinical examination2a. If markers <strong>of</strong> high-risk caries present: 2b. If markers <strong>of</strong> high-risk periodontitis present:1. Assess if these criteria possibly are associated with 1. assess if these criteria possibly are associated withor can contribute to diseaseor can contribute to disease2. Consider consequences <strong>of</strong> monitoring, correcting or 2. Consider consequences <strong>of</strong> monitoring, correcting orreplacing restorationreplacing restorationSurfaceSurfaceFractured, rough or pitted or irregular, flaking or hasFractured, rough or pitted or irregular, flaking or hasgross porositiesgross porosities(SRO)(TGI)(TPIT)(VSF)(VFK)(VGP)(SRO)(TGI)(TPIT)(VSF)(VFK)(VGP)ContourContourExposed dentine or base (TDE)(TBA)Contact slightly open or faulty (SCO)(TCO)Undercontoured cervical area approximally (SPX)(TPX)Undercontoured cervical area approximally (SPX)(TPX)MarginMarginDitch or gap along the margin (SCR)(TMD)Ditch or gap along the margin (SCR)(TMD)Discoloured margin (SDIS)(TPEN)OtherTraumatic occlusion (VTO)3. Limit intervention to monitoring – unless the patient is dissatisfied.ContourUndercontoured or overcontoured restoration (SUCO)(TUCO)(SOCO)(SOC)(TOCO)(VUO)The occlusion is affected (SOH)(TET)(TOC)Under-contoured marginal ridges (SMR)Flattening present facially or lingually (SFA)(SLG)and their persistence, presence <strong>of</strong>ecological niches with difficult toaccess sites such as furcations, andthe presence <strong>of</strong> iatrogenic factorssuch as restoration discrepancies.Information gathered by clinicalmonitoring and continued multilevelrisk assessment produces anestimate <strong>of</strong> the oral health status <strong>of</strong>an individual, and risk <strong>of</strong> oraldisease progression at a particulartooth or site. It is not until thisstage that concern about the technicalexcellence <strong>of</strong> one or moreparticular <strong>restorations</strong> should beaddressed. Thus, the risk level fororal diseases must in a systematicway first be recognised, and thencoupled with treatment optionsthat are consistent with the potentialfuture caries increment orperiodontal disease. It has beensuggested that a decision-treeInternational Dental Journal (2001) Vol. 51/No.3approach and/or treatment-indexconcept should be applied tospecific clinical conditions andpreventive-restorative options toestimate the probable outcomes 298 .In the introduction, a definition<strong>of</strong> the quality <strong>of</strong> <strong>restorations</strong>emphasised the risk <strong>of</strong> jeopardisingthe integrity <strong>of</strong> remaining<strong>dental</strong>-related tissues. Patients withindications <strong>of</strong> severe caries or periodontaldisease require moreattention to possible detrimentalcharacteristics <strong>of</strong> <strong>restorations</strong>compared to patients with no signs<strong>of</strong> disease. The concept <strong>of</strong> such anapproach is consistent with thetreatment decision philosophypractised by many clinicians. Theclinician always has three options indeciding on a strategy for intervention.Either to ignore the currentstatus, to adjust or repair, or restorationreplacement.Table 4 suggests how the technicalexcellence <strong>of</strong> <strong>restorations</strong> shouldbe appraised clinically in light <strong>of</strong>risks <strong>of</strong> oral disease. The wording<strong>of</strong> the criteria parallels descriptionsused in the CDA evaluationsystem 31 and textbooks cited at thestart <strong>of</strong> this section. Table 5 showsthe CDA quality evaluation ratingsystem and USPHS criteria forevaluation.It must be emphasised that theconsiderations <strong>of</strong> the consequences<strong>of</strong> monitoring, correcting, removingor replacing an existing restorationmust be but one component <strong>of</strong> themanagement <strong>of</strong> oral disease. Furtherrequirements to justify operativeintervention should include patientunderstanding <strong>of</strong> risks and prognosis,assessment <strong>of</strong> aetiology andthe instruction <strong>of</strong> preventive proce-


149Table 5 Relationship between the CDA and the USPHS criteria for evaluation <strong>of</strong> <strong>dental</strong> restoration systems 30,31 .Surface and color Anatomic form Margin IntegrityR: Range <strong>of</strong> excellenceRestoration is <strong>of</strong> satisfactory quality and is expected to protect tooth and surrounding tissueCDA USPHS CDA USPHS CDA USPHSSurface <strong>of</strong> restoration is The restoration appears to Restoration contour is in The restoration is a No visible evidence <strong>of</strong> There is no visual evidence <strong>of</strong> dark,smooth match the shade and functional harmony with continuation <strong>of</strong> existing crevice on margin into deep discoloration adjacent to theNo irritation <strong>of</strong> adjacent translucency <strong>of</strong> adjacent adjacent teeth and s<strong>of</strong>t anatomic form or is slightly which explorer will penetrate. restorationtissue is occurring tooth tissues tissues with good individual flattened. It may be over- Satisfies principles <strong>of</strong> marginThere is no mismatch in anatomic form contoured. When the side <strong>of</strong> placement wherever possible. The explorer does not catch whencolor or translucency (The restoration must be the explorer is placed No discoloration on margin drawn across the surface <strong>of</strong> thebetween <strong>restorations</strong> and examined without using a tangentially across the between restoration and restoration toward the tooth, or, if theadjacent teeth* mouth mirror) restoration, it does not touch tooth structure* explorer does catch there is no visibletwo opposing cavosurface crevice along the periphery <strong>of</strong> theline angles at the same time restorationThere is no visual evidence <strong>of</strong>marginal discoloration difference fromthe color <strong>of</strong> the restorative materialand from the color <strong>of</strong> the adjacenttooth structureS: Range <strong>of</strong> acceptabilityRestoration is <strong>of</strong> acceptable quality but exhibits one or more features that deviate from idealCDA USPHS CDA USPHS CDA USPHSSRO Surface <strong>of</strong> restoration The restoration does not SOCO Restoration over- A surface concavity is SCR Visible evidence <strong>of</strong> There is visual evidence <strong>of</strong> marginalis slightly rough or pitted; match the shade and contoured slightly (but evident. When the side <strong>of</strong> an slight marginal discrepancy discoloration at the junction <strong>of</strong> thecan be polished translucency <strong>of</strong> adjacent excess material can be explorer is placed tangentially with no evidence <strong>of</strong> decay; tooth structure and the restoration, butSMM Slight mismatch tooth tissues, but the removed) across the restoration, the repair can be made or is the discoloration has not penetratedbetween shade <strong>of</strong> mismatch is within the normal SUCO Restoration slightly explorer touches two unnecessary (visible ditching along the restoration in a pulpalrestoration(s) and adjacent range <strong>of</strong> tooth shades undercontoured opposing cavosurface line along the margin not directiontooth or teeth* SOH Occlusion is not totally angles at the save time, but extending to the DE junction)(… and tooth structure within functional (or height reduced the dentin or base is not The explorer catches and there isthe normal range <strong>of</strong> tooth locally(not in toto)) exposed SDIS Discoloration on margin visible evidence <strong>of</strong> a crevice, into whichcolor, shade and/or SMR Marginal ridges slightly between restoration and the explorer penetrates, indicating thattranslucency) undercontoured tooth structure* the edge <strong>of</strong> the restoration does notSCO Contact slightly open adapt closely to the tooth structure.SFA Facial flattening present The dentin and/or the base is notSLG Lingual flattening present exposed, and the restoration is notSAF Anatomic form <strong>of</strong> pontic mobilemay cause food retention; noirritation <strong>of</strong> s<strong>of</strong>t tissueSOC Occlusal contour notcontinous with that <strong>of</strong> cuspsand planesSPX Interproximal cervicalarea slightly undercontoured<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


150T: Replace or correct for preventionRestoration is not <strong>of</strong> acceptable quality. Future damage to tooth or surrounding tissue is likely to occurCDA USPHS CDA USPHS CDA USPHSTGI Surface grossly irregular, The restoration does not TUC0 Restorations grossly There is a loss <strong>of</strong> restorative TFAM Faulty margins that There is visual evidence <strong>of</strong> dark, deepnot related to anatomy and match the shade and undercontoured substance so that a surface cannot be properly repaired discoloration adjacent to the restorationnot subject to correction translucency <strong>of</strong> the adjacent TOCO Restorations grossly concavity is evident and the TPEN Penetrating discoloration (but not directly associated withTMM Mismatch between tooth structure, and the overcontoured base and/or dentin is along margin <strong>of</strong> restoration in cavosurface margins)restoration(s) and adjacent mismatch is outside the TET Occlusion affected exposed pulpal direction*tooth or teeth outside normal normal range <strong>of</strong> tooth shades TCO Contact is faulty TCEM Retained excess The explorer penetrates a crevicerange <strong>of</strong> color, shade, or and translucency TOV There is marginal cement defect that extends to the dentoenameltranslucency* overhang junctionTAF Anatomic form <strong>of</strong> pontic TMD visible ditching alongTPIT Surface deeply pitted, likely to result in food the margin extending to the There is visual evidence <strong>of</strong> marginalirregular grooves (not related retention, causing irritation to DE junction discoloration at the junction <strong>of</strong> the toothto anatomy) cannot be s<strong>of</strong>t tissue or caries in TMB ditching along the structure and the restoration that hasrefinished abutments margin extending to the penetrated along the restoration in acenter base pulpal directionTDE Dention is exposedTBA base is exposedTOC A occlusion is affectedTPX contact is faulty –self-correction unlikelyV: Replace statimRestoration is not <strong>of</strong> acceptable quality. Damage to tooth or its surrounding tissues is now occurringCDA USPHS CDA USPHS CDA USPHSVSF Surface is fractured VTO Traumatic occlusion VMO Mobile restorationVGP There are gross VUO Gross underocclusion or VFR Fractured restorationporosities in crown material restoration VCAR Caries continuous withVSD Shade in gross VPN Restoration causes margin <strong>of</strong> restorationdisharmony with adjacent unremitting pain in tooth or VTF Tooth structure isteeth* adjacent tissue fracturedVDM Damage is now occurringVFK surface is flaking to tooth, s<strong>of</strong>t tissue, orVUN Esthetically displeasing supporting bonecolor, shade and/ortranslucency VMIS restoration is missing*Criteria apply only to anterior teethInternational Dental Journal (2001) Vol. 51/No.3


151dures, including dietary advice,counselling and plaque controlinstruction.References1. Anusavice K J. Criteria for selection<strong>of</strong> restorative materials: propertiesversus technique sensitivity. In:Anusavice, K J (ed.) <strong>Quality</strong> Evaluation<strong>of</strong> Dental Restorations. pp 15–56.Chicago: Quintessence PublishingCompany, 1989.2. Mjör I A. Repair versus replacement<strong>of</strong> failed <strong>restorations</strong>. Int Dent J 199343: 466–472.3. Maryniuk G A. Replacement <strong>of</strong> amalgam<strong>restorations</strong> that have marginaldefects: variation and cost implications.Quintessence Int 1990 21: 311–319.4. Bronkhorst E. Interne gegevensonderzoeklijn gezondheidszorg.Vakgroep Cariologie en Endodontologie.Nijmegen: Katolieke Universiteit,1988.5. NHS Centre for Reviews andDissemination. Sheldon T, TreasureE (eds). Dental Restoration: WhatType <strong>of</strong> Filling? Effective Health Care1999 5: 1–12.6. Elderton R J. The causes <strong>of</strong> failure <strong>of</strong><strong>restorations</strong>: a literature review. JDent 1976 4: 257–262.7. Elderton R J. The quality <strong>of</strong> amalgam<strong>restorations</strong>. In: Allred H (ed).Assessment <strong>of</strong> the <strong>Quality</strong> <strong>of</strong> Dental Care.pp 45–81. London: Dorriston, 1977.8. Downer M C, O’Brien G J. Evaluatinghealth gains from restorative<strong>dental</strong> treatment. Community DentOral Epidemiol 1994 22: 209–213.9. Shugars D A, Bader J D. Appropriateness<strong>of</strong> care. Appropriateness <strong>of</strong>restorative treatment recommendations:a case for practice based outcomesresearch. J Am Coll Dentist1992 59: 7–13.10. Bader J D, Shugars D A. Variation,treatment outcomes, and practiceguidelines in <strong>dental</strong> practice. J DentEduc 1995 59: 61–95.11. Friedman J W, Atchison K A. Thestandard <strong>of</strong> care: an ethical responsibility<strong>of</strong> public health dentistry. JPublic Health Dent 1993 53: 165–169.12. Knibbs P J. Methods <strong>of</strong> clinicalevaluation <strong>of</strong> <strong>dental</strong> restorativematerials. J Oral Rehabil 1997 24:109–123.13. Randall R C, Wilson N H F. Glassionomerrestoratives: a systematicreview <strong>of</strong> a secondary caries treatmenteffect. J Dent Res 1999 78:628–637.14. Bergman M A. The clinical performance<strong>of</strong> ceramic inlays: a review. AustDent J 1999 44: 157–168.15. Martin N, Jedynakiewicz N M. Clinicalperformance <strong>of</strong> CEREC ceramicinlays: a systematic review. DentMater 1999 15: 54–61.16. Chadwick B L, Dummer P M,Dunstan F D, et al. What type <strong>of</strong>filling? Best practice in <strong>dental</strong> <strong>restorations</strong>.Qual Health Care 1999 8:202–207.17. Downer M C, Azli N A, Bedi R, et al.How long do routine <strong>dental</strong> <strong>restorations</strong>last? A systematic review. BrDent J 1999 187: 432–439.18. Merriam-Webster’s Collegiate DictionaryOnline. URL (1.12.2000):http://www.m-w.com/19. Shugars D A, Bader J D. Cost implications<strong>of</strong> differences in dentists’restorative treatment decisions. JPublic Health Dent 1996 56: 219–222.20. Osborne J W, Phillips R W, Gale EN, et al. Three-year clinical comparison<strong>of</strong> three amalgam alloy typesemphasizing an appraisal <strong>of</strong> theevaluation methods used. J Am DentAssoc 1976 93: 784–789.21. Mahler D B, Marantz R. The effect<strong>of</strong> the operator on the clinical performance<strong>of</strong> amalgam. J Am DentAssoc 1979 99: 38–41.22. Borgmeijer P J, Advokaat J G,Akerboom H B. <strong>Quality</strong> control <strong>of</strong>amalgam <strong>restorations</strong>. Part II. Acomparison <strong>of</strong> four non-clinicalmethods <strong>of</strong> evaluation. Ned TijdschrTandheelkd 1983 90 (Suppl 22): 24–43.23. Roulet J F, Reich T, Blunck U, et al.Quantitative margin analysis in thescanning electron microscope. ScanningMicrosc 1989 3: 147–158.24. Grossman E S, Matejka J M. Amalgammarginal quality assessment: acomparison <strong>of</strong> seven methods. J OralRehabil 1997 24: 496–505.25. van Amerongen W E, Eggink C O.The cervical margin <strong>of</strong> amalgam <strong>restorations</strong>:a radiographic and clinicalassessment. ASDC J Dent Child 198653: 177–183.26. Leinfelder K F, Taylor D F,Barkmeier W W, et al. Quantitativewear measurement <strong>of</strong> posterior compositeresins. Dent Mater 1986 2:198–201.27. Bryant R W. Comparison <strong>of</strong> threestandards for quantifying occlusalloss <strong>of</strong> composite <strong>restorations</strong>. DentMater 1990 6: 60–62.28. Smales R J, Creaven P J. Evaluation<strong>of</strong> clinical methods for assessing thesurface roughness <strong>of</strong> <strong>restorations</strong>. J<strong>Pro</strong>sthet Dent 1979 42: 45–52.29. Jokstad A. Class 2 cavity preparationsand restoration performance.(Dr. thesis) Oslo: University <strong>of</strong> Oslo,Norway, 1992.30. Cvar J F, Ryge G. Criteria for theclinical evaluation <strong>of</strong> <strong>dental</strong> restorativematerials. San Francisco: UnitedStates Dental Health Center 1971,publication no. 7902244.31. California Dental Association. Guidelinesfor the assessment <strong>of</strong> clinical qualityand pr<strong>of</strong>essional performance. 3rd ed.Sacramento, CA: California DentalAssociation, 1995.32. van Dijken J W. A clinical evaluation<strong>of</strong> anterior conventional,micr<strong>of</strong>iller, and hybrid compositeresin fillings. A 6-year follow-upstudy. Acta Odontol Scand 1986 44:357–367.33. Allander L, Birkhed D, Bratthall D.<strong>Quality</strong> evaluation <strong>of</strong> anterior <strong>restorations</strong>in private practice. Swed DentJ 1989 13: 141–150.34. Lutz F U, Krejci I, Besek M. Operativedentistry: the missing clinicalstandards. Pract Periodontics AesthetDent 1997 9: 541–548.35. Carpay J J, Nieman F H, Konig K G,et al. <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>and <strong>dental</strong> treatment in Dutch schoolchildren. Community Dent Health 19907: 43–51.36. Mjör I A, Dahl J E, Moorhead J E.Age <strong>of</strong> <strong>restorations</strong> at replacement inpermanent teeth in general <strong>dental</strong>practice. Acta Odontol Scand 2000 58:97–10137. Sano H, Kanemura N, Burrow M F,et al. Effect <strong>of</strong> operator variabilityon dentin adhesion: students vs. dentists.Dent Mater J 1998 17: 51–58.38. Ciucchi B, Bouillaguet S, Holz J, etal. The battle <strong>of</strong> the bonds 1995.Schweiz Mschr Zahnmed 1997 107: 37–39.39. Lussi A, Brunner M, Portmann P, etal. Condensation pressure duringamalgam placement in patients. EurJ Oral Sci 1995 103: 388–393.40. Wasson E A, Nicholson J W. Effect<strong>of</strong> operator skill in determining thephysical properties <strong>of</strong> glass-ionomercements. Clin Mater 1994 15: 169–172.41. Billington R W, Williams J A,Pearson G J. Variation in powder/liquid ratio <strong>of</strong> a restorative glassionomercement used in <strong>dental</strong> practice.Br Dent J 1990 169: 164–167.42. Gjerdet N R, Hegdahl T. Porosity,strength, and mercury content <strong>of</strong>amalgam made by different dentists<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


152in their own practice. Dent Mater1985 1: 150–153.43. Dermann K. Untersuchungen zurReproduzierbarkeit einiger Werkst<strong>of</strong>fprüfungenbei Amalgamen.Dtsch Zahnärztl Z 1989 44: 764–765.44. McCabe J F, Watts D C, Wilson H J,et al. An investigation <strong>of</strong> test-housevariability in the mechanical testing<strong>of</strong> <strong>dental</strong> materials and the statisticaltreatment <strong>of</strong> results. J Dent 1990 18:90–97.45. Ferracane J L, Mitchem J C. <strong>Pro</strong>perties<strong>of</strong> posterior composite: results<strong>of</strong> round robin testing for a specification.Dent Mater 1994 10: 92–99.46. Sturdevant, C M, Roberson T M,Heymann H O, et al. The Art andScience <strong>of</strong> Operative Dentistry. 3rd ed.pp 184–205. St. Louis: Mosby, 1995.47. Poorterman J H, Aartman I H,Kalsbeek H. Underestimation <strong>of</strong> theprevalence <strong>of</strong> approximal caries andinadequate <strong>restorations</strong> in a clinicalepidemiological study. CommunityDent Oral Epidemiol 1999 27: 331–337.48. Tobi H, Groen H J, Kreulen C M, etal. Observer variation in the assessment<strong>of</strong> resin composite. Dent Mater1998 14: 1–5.49. Oleinisky J C, Baratieri L N, RitterA V, et al. Influence <strong>of</strong> finishing andpolishing procedures on the decisionto replace old amalgam <strong>restorations</strong>:an in vitro study. Quintessence Int1996 27: 833–840.50. Dunne S M. The limitation <strong>of</strong> visualperception in restorative dentistry.Dent Update 1993 20: 198–205.51. Dunninger P, Einwag J, Sitter H.Reproduzierbarkeit von Messungenzur Ergebnisqualität zahnärztlicherFüllungen. Dtsch Zahnärztl Z 199146: 212–214.52. Kidd E A, Ricketts D N, Pitts N B.Occlusal caries diagnosis: a changingchallenge for clinicians and epidemiologists.J Dent 1993 21: 323–33153. Pitts N B. Diagnostic tools andmeasurements – impact on appropriatecare. Community Dent OralEpidemiol 1997 25: 24–35.54. Horn D J, Bulan-Brady J, Hicks M L.Sphere spectrophotometer versushuman evaluation <strong>of</strong> tooth shade. JEndod 1998 24: 786–790.55. Kay E J, Blinkhorn A S. A qualitativeinvestigation <strong>of</strong> factors governingdentists’ treatment philosophies.Br Dent J 1996 180: 171–176.56. Hawthorne W S, Smales R J. Factorsaffecting the amount <strong>of</strong> long-termrestorative <strong>dental</strong> treatment providedto 100 patients by 20 dentists in 3Adelaide private practices. Aust DentJ 1996 41: 256–259.57. Paterson F M, Paterson R C, WattsA, et al. Initial stages in the development<strong>of</strong> valid criteria for the replacement<strong>of</strong> amalgam <strong>restorations</strong>. J Dent1995 23: 137–143.58. el-Mowafy O M, Lewis D W.Restorative decision making byOntario dentists. J Can Dent Assoc1994 60: 305–306.59. Bader J D, Shugars D A. Agreementamong dentists’ recommendationsfor restorative treatment. J Dent Res1993 72: 891–896.60. Bader J D, Shugars D A. Understandingdentists’ restorative treatmentdecisions. J Public Health Dent 199252: 102–110.61. Drake C W, Maryniuk G A, BentleyC. Reasons for restoration replacement:differences in practice patterns.Quintessence Int 1990 21: 125–130.62. Worthington H, Holloway P,Clarkson J, et al. Predicting whichadult patients will need treatmentover the next year. Community DentOral Epidemiol 1997 25: 273–277.63. Jacobsen P H. Design and analysis <strong>of</strong>clinical trials. J Dent 1988 16: 215–218.64. Fédération Dentaire Internationale.Recommendations for clinicalresearch protocols for <strong>dental</strong> materials.Int Dent J 1982 32: 403–423.65. National Institutes <strong>of</strong> Health, ADAand FDA. Wych<strong>of</strong>f H D, Weigl D G.(eds). Clinical Evaluation <strong>of</strong> DentalMaterials. Bethesda: National Institutes<strong>of</strong> Health Publication 1980,publication no. NIH 80–117.66. ADA Council on Dental Materialsand Devices. Recommended standardpractices for clinical researchprotocol for <strong>dental</strong> materials anddevices. J Am Dent Assoc 1972 84:388–390.67. ADA Council on Dental Materials,Instrument, and Equipment. AmericanDental Association Guidelines forSubmission <strong>of</strong> Composite Resin Materialsfor Occlusal Class I and Class IIRestorations. Chicago: American DentalAssociation, 1989.68. ADA Council on Dental Materials,Instruments, and Equipment. AmericanDental Association Acceptance <strong>Pro</strong>gramfor Dentin Adhesive Materials andEnamel and Dentin Adhesive Materials.Chicago: American Dental Association,1994.69. Letzel H, van ‘t H<strong>of</strong> M A, MarshallG W, et al. The influence <strong>of</strong> theamalgam alloy on the survival <strong>of</strong>amalgam <strong>restorations</strong>: a secondaryanalysis <strong>of</strong> multiple controlled clinicaltrials. J Dent Res 1997 76: 1787–1798.70. Tyas M J. Dental materials science –the maintenance <strong>of</strong> standards. J OralRehabil 1992 18: 105–110.71. NIDR, National Institute <strong>of</strong> DentalResearch. Broadening the scope – longrange research plan for the nineties.Bethesda: US Department <strong>of</strong> Healthand Human Services 1990, publicationno. NIH 90–1188.72. Letzel H. Survival rates and reasonsfor failure <strong>of</strong> posterior composite<strong>restorations</strong> in multicentre clinicaltrial. J Dent 1989 17 (Suppl 1): S10–S17.73. Wilson N H F. The evaluation <strong>of</strong>materials: relationship between laboratoryinvestigations and clinicalstudies. Oper Dent 1990 15: 149–155.74. Altman D G. Practical Statistics forMedical Research. pp 455–460.London: Chapman & Hall, 1991.75. Mair L H. Understanding wear indentistry. Compend Contin Educ Dent1999 20: 19–26.76. Dahl B L, Carlsson G E, Ekfeldt A.Occlusal wear <strong>of</strong> teeth and restorativematerials. A review <strong>of</strong> classification,etiology, mechanisms <strong>of</strong> wear,and some aspects <strong>of</strong> restorative procedures.Acta Odontol Scand 1993 51:299–311.77. Freilich M A, Goldberg A J,Gilpatrick R O, et al. Direct andindirect evaluation <strong>of</strong> posterior composite<strong>restorations</strong> at three years.Dent Mater 1992 8: 60–64.78. Wilson N H F, Wilson M A, WastellD G, et al. Performance <strong>of</strong> Occlusinin butt-joint and bevel-edged preparations:Five year results. Dent Mater1991 7: 92–98.79. Lutz F, Phillips R W, Roulet J F, etal. In vivo and in vitro wear <strong>of</strong>potential posterior composites. JDent Res 1984 63: 914–920.80. Mair L H, Vowles R W, CunninghamJ, et al. The clinical wear <strong>of</strong> threeposterior composites. Br Dent J 1990169: 355–360.81. Ratanapridakul K, Leinfelder K F,Thomas J. Effects <strong>of</strong> finishing on thein vivo wear rate <strong>of</strong> a posterior compositeresin. J Am Dent Assoc 1989118: 333–335.82. Lambrechts P, Braem M, Vuylsteke-Wauters M, et al. Quantitative invivo wear <strong>of</strong> human enamel. J DentRes 1989 68: 1752–1754.83. Johnson G H, Bales D J, Gordon GE, et al. Clinical performance <strong>of</strong>posterior composite resin restora-International Dental Journal (2001) Vol. 51/No.3


153tions. Quintessence Int 1992 23: 705–711.84. Wendt Jr S L, Ziemiecki T L,Leinfelder K F. <strong>Pro</strong>ximal wear ratesby tooth position <strong>of</strong> resin composite<strong>restorations</strong>. J Dent 1996 24: 33–39.85. Wendell J J, Vann Jr W F. Wear <strong>of</strong>composite resin <strong>restorations</strong> in primaryversus permanent molar teeth.J Dent Res 1988 67: 71–74.86. Rasmusson C G, Köhler B, ÖdmanP. A 3-year clinical evaluation <strong>of</strong> twocomposite resins in Class-II cavities.Acta Odontol Scand 1998 56: 70–75.87. Köhler B, Rasmusson C G, ÖdmanP. A five-year clinical evaluation <strong>of</strong>Class II composite resin <strong>restorations</strong>.J Dent 2000 28: 111–116.88. Qvist V, Strøm C. 11-year assessment<strong>of</strong> Class-III resin <strong>restorations</strong>completed with two restorative procedures.Acta Odontol Scand 1993 51:253–262.89. Yap A U, Tan K B, Bhole S. Comparison<strong>of</strong> aesthetic properties <strong>of</strong>tooth-colored restorative materials.Oper Dent 1997 22: 167–172.90. Cook W D. Optical properties <strong>of</strong>esthetic restorative materials andnatural dentition. J Biomed Mater Res1985 19: 469-488.91. Mjör I A, T<strong>of</strong>fenetti F. Secondarycaries: a literature review with casereports. Quintessence Int 2000 31:165–179.92. Kreulen C M, van Amerongen W E,Akerboom H B, et al. Two-yearresults with box-only resin composite<strong>restorations</strong>. ASDC J Dent Child1995 62: 395–400.93. Nordbø H, Leirskar J, von der FehrF R. Saucer-shaped cavity preparationsfor posterior approximal resincomposite <strong>restorations</strong>: observationsup to 10 years. Quintessence Int 199829: 5–11.94. Lutz F, Krejci I, Luescher B, et al.Improved proximal margin adaptation<strong>of</strong> Class II composite resin <strong>restorations</strong>by use <strong>of</strong> light-reflectingwedges. Quintessence Int 1986 17:659–664.95. van Dijken J W, Horstedt P, WaernR. Directed polymerization shrinkageversus a horizontal incrementalfilling technique: interfacial adaptationin vivo in Class II cavities. Am JDent 1998 11: 165–172.96. Koran P, Kurschner R. Effect <strong>of</strong>sequential versus continuous irradiation<strong>of</strong> a light-cured resin compositeon shrinkage, viscosity, adhesion, anddegree <strong>of</strong> polymerization. Am J Dent1998 11: 17–2297. King P A. Adhesive techniques. BrDent J 1999 186: 321–326.98. Van Meerbeek B, Inokoshi S, BraemM, et al. Morphological aspects <strong>of</strong>the resin-dentin interdiffusion zonewith different dentin adhesive systems.J Dent Res 1992 71: 1530–1540.99. Qvist V. Resin <strong>restorations</strong>: leakage,bacteria, pulp. Endod Dent Traumatol1993 9: 127–152.100. Manhart J, Neuerer P, Scheibenbogen-FuchsbrunnerA, et al. Threeyearclinical evaluation <strong>of</strong> direct andindirect composite <strong>restorations</strong> inposterior teeth. J <strong>Pro</strong>sthet Dent 200084: 289–296.101. Scheibenbogen-Fuchsbrunner A,Manhart J, Kremers L, et al. Twoyearclinical evaluation <strong>of</strong> direct andindirect composite <strong>restorations</strong> inposterior teeth. J <strong>Pro</strong>sthet Dent 199982: 391–397.102. Peumans M, Van Meerbeek B,Lambrechts P, et al. The 5-year clinicalperformance <strong>of</strong> direct compositeadditions to correct tooth form andposition. II. Marginal qualities. ClinOral Investig 1997 1: 19–26.103. Qualtrough A J, Piddock V. Fittingaccuracy <strong>of</strong> indirect <strong>restorations</strong>: areview <strong>of</strong> methods <strong>of</strong> assessment. EurJ <strong>Pro</strong>sthodont Restor Dent 1992 1: 57–61.104. Roulet J F. Marginal integrityrs: clinicalsignificance. J Dent 1994 22(Suppl 1): S9–S12.105. Mahler D B. The high-copper <strong>dental</strong>amalgam alloys. J Dent Res 1997 76:537–541.106. Osborne J W, Norman R D, Gale EN. A 14-year clinical assessment <strong>of</strong>14 amalgam alloys. Quintessence Int1991 22: 857–864.107. Akerboom H B, Advokaat J G, VanAmerongen W E, et al. Long-termevaluation and rerestoration <strong>of</strong> amalgam<strong>restorations</strong>. Community DentOral Epidemiol 1993 21: 45–48.108. Gruythuysen R J, Kreulen, C M, TobiH, et al. 15-year evaluation <strong>of</strong> ClassII amalgam <strong>restorations</strong>. CommunityDent Oral Epidemiol 1996 24: 307–310.109. Kreulen C M, Tobi H, GruythuysenR J, et al. Replacement risk <strong>of</strong> amalgamtreatment modalities: 15-yearresults. J Dent 1998 26: 627–632.110. Bryant R W, Marzbani N, Hodge KV. Occlusal margin defects arounddifferent types <strong>of</strong> composite resin<strong>restorations</strong> in posterior teeth. OperDent 1992 17: 215–221.111. Stratis S, Bryant R W. The influence<strong>of</strong> modified cavity design and finishingtechniques on the clinical performance<strong>of</strong> amalgam <strong>restorations</strong>: a2-year clinical study. J Oral Rehabil1998 25: 269–278.112. Kreulen C M, van Amerongen W E,Akerboom H B, et al. Evaluation <strong>of</strong>occlusal marginal adaptation <strong>of</strong> ClassII resin-composite <strong>restorations</strong>.ASDC J Dent Child 1993 60: 310–314.113. Söderholm K J, Tyas M J, Jokstad A.Determinants <strong>of</strong> quality in operativedentistry. Crit Rev Oral Biol Med 19989: 464–479.114. Fukushima M, Setcos J C, Phillips RW. Marginal fracture <strong>of</strong> posteriorcomposite resins. J Am Dent Assoc1988 117: 577–583.115. Van Meerbeek B, Vanherle G,Lambrechts P, et al. Dentin- andenamel-bonding agents. Current OpinionsDent 1992 2: 117–127.116. Pashley D H. Dynamics <strong>of</strong> the pulpodentincomplex. Crit Rev Oral BiolMed 1996 7: 104–133.117. Van Meerbeek B, Yoshida Y,Lambrechts P, et al. A TEM study <strong>of</strong>two water-based adhesive systemsbonded to dry and wet dentin. J DentRes 1998 77: 50–59.118. Tay F R, Gwinnett A J, Pang K M,et al. Resin permeation into acid-conditioned,moist, and dry dentin: aparadigm using water-free adhesiveprimers. J Dent Res 1996 75: 1034–1044.119. Van Meerbeek B, Yoshida Y,Snauwaert J, et al. Hybridizationeffectiveness <strong>of</strong> a two-step versus athree-step smear layer removingadhesive system examined correlativelyby TEM and AFM. J AdhesiveDent 1999 1: 7–23.120. Uno S, Finger W J. Effect <strong>of</strong> acidetchant composition and etch durationon enamel loss and resin compositebonding. Am J Dent 1995 8:165–169.121. Pioch T, Stotz S, Staehle H J, et al.Applications <strong>of</strong> confocal laser scanningmicroscopy to <strong>dental</strong> bonding.Adv Dent Res 1997 11: 453–461.122. Gwinnett A J. Moist versus dry dentin:its effect on shear bond strength.Am J Dent 1992 5: 127–129.123. Kanca J. Resin bonding to wetsubstrate. I. Bonding to dentin. QuintessenceInt 1992 23: 39–41.124. Kanca J. Wet bonding: effect <strong>of</strong> dryingtime and distance. Am J Dent1996 9: 273–276.125. Tay F R, Gwinnett A J, Wei S H.Relation between water content inacetone/alcohol-based primer andinterfacial ultrastructure. J Dent 199826: 147–156.<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


154126. Powers J M, Finger W J, Xie J. Bonding<strong>of</strong> composite resin to contaminatedhuman enamel and dentin. IntJ <strong>Pro</strong>sthodont 1995 4: 28–32.127. Barghi N, Knight G T, Berry T G.Comparing two methods <strong>of</strong> moisturecontrol in bonding to enamel: a clinicalstudy. Oper Dent 1991 16: 130–135.128. Knight G T, Berry T G, Barghi N, etal. Effects <strong>of</strong> two methods <strong>of</strong> moisturecontrol on marginal microleakagebetween resin composite andetched enamel: a clinical study. Int J<strong>Pro</strong>sthodont 1993 6: 475–479.129. Berg N G, Derand T. A 5-year evaluation<strong>of</strong> ceramic inlays (CEREC).Swed Dent J 1997 21: 121–127.130. Osborne J W, Gale E N. Relationship<strong>of</strong> restoration width, toothposition, and alloy to fracture at themargins <strong>of</strong> 13- to 14-year-old amalgams.J Dent Res 1990 69: 1599–1601.131. Osborne J W, Gale E N. Failure atthe margin <strong>of</strong> amalgams as affectedby cavity width, tooth position andalloys selection. J Dent Res 1981 60:681–685.132. Goldberg J, Munster E, Rydinge E,et al. Experimental design in theclinical evaluation <strong>of</strong> amalgam <strong>restorations</strong>.J Biomed Mater Res 1980 14:777–788.133. Derand T. Marginal failure <strong>of</strong> amalgam.Effect <strong>of</strong> alloy selection andbite forces. Swed Dent J 1983 7: 65–68.134. Harris R K. Dental amalgam: successor failure? Oper Dent 1992 17: 243–252.135. Owens B M. Replacement and initialplacement <strong>of</strong> tooth colored <strong>restorations</strong>:a review and discussion. J TennDent Assoc 1998 78: 26–29.136. Maryniuk G A, Kaplan S H. Longevity<strong>of</strong> <strong>restorations</strong>: survey results <strong>of</strong>dentists’ estimates and attitudes. JAm Dent Assoc 1986 112: 39–45.137. Gladys S, Van Meerbeek B, InokoshiS, et al. Clinical and semiquantitativemarginal analysis <strong>of</strong> four toothcolouredinlay systems at 3 years. JDent 1995 23: 329–338.138. Kramer N, Frankenberger R, PelkaM, et al. IPS Empress inlays andonlays after four years – a clinicalstudy. J Dent 1999 27: 325–331.139. Lundin S A, Andersson B, Koch G,et al. Class II composite resin <strong>restorations</strong>:a three-year clinical study <strong>of</strong>six different posterior composites.Swed Dent J 1990 14: 105–114.140. Rasmusson C G, Lundin S A. Class II<strong>restorations</strong> in six different posteriorcomposite resins: five-yearresults. Swed Dent J 1995 19: 173–182.141. Dunne S M, Millar B J. A longitudinalstudy <strong>of</strong> the clinical performance<strong>of</strong> porcelain veneers. Br Dent J 1993175: 317–321.142. van Noort R, Davis L G. A prospectivestudy <strong>of</strong> the survival <strong>of</strong> chemicallyactivated anterior resin composite<strong>restorations</strong> in general <strong>dental</strong>practice: 5-year results. J Dent 199321: 209–215.143. Pilebro C E, van Dijken J W,Stenberg R. Durability <strong>of</strong> tunnel <strong>restorations</strong>in general practice: a threeyearmulticenter study. Acta OdontolScand 1999 57: 35–39.144. Meijering A C, Creugers N H, RoetersF J, et al. Survival <strong>of</strong> three types <strong>of</strong>veneer <strong>restorations</strong> in a clinical trial:a 2.5-year interim evaluation. J Dent1998 26: 563–568.145. Plasmans P J, Creugers N H, MulderJ. Long-term survival <strong>of</strong> extensiveamalgam <strong>restorations</strong>. J Dent Res1998 77: 453–460.146. Wilson N H, Wastell D G, NormanR D. Five-year performance <strong>of</strong> highcoppercontent amalgam <strong>restorations</strong>in a multiclinical trial <strong>of</strong> a posteriorcomposite. J Dent 1996 24: 203–210.147. Wilson N H F, Norman R D. Fiveyearfindings <strong>of</strong> a multiclinical trialfor a posterior composite. J Dent1991 19: 153–159.148. Phantumvanit P, Songpaisan Y,Pilot T, et al. Atraumatic restorativetreatment (ART): a three-year communityfield trial in Thailand –survival <strong>of</strong> one-surface <strong>restorations</strong>in the permanent dentition. J PublicHealth Dent 1996 56 (Spec iss 3):141–145149. Wendt L K, Koch G, Birkhed D.Replacements <strong>of</strong> <strong>restorations</strong> in theprimary and young permanent dentition.Swed Dent J 1998 22: 149–155.150. Hawthorne W S, Smales R J. Factorsinfluencing long-term <strong>restorations</strong>urvival in three private <strong>dental</strong> practicesin Adelaide. Aust Dent J 199742: 59–63.151. Shaini F J, Shortall A C, Marquis PM. Clinical performance <strong>of</strong> porcelainlaminate veneers. A retrospectiveevaluation over a period <strong>of</strong> 6.5 years.J Oral Rehabil 1997 24: 553–559.152. Mahmood S, Smales R J. Longevity<strong>of</strong> <strong>dental</strong> <strong>restorations</strong> in selectedpatients from different practiceenvironments. Aust Dent J 1994 39:15–17.153. Smales R J. Effects <strong>of</strong> enamel-bonding,type <strong>of</strong> restoration, patient ageand operator on the longevity <strong>of</strong> ananterior composite resin. Am J Dent1991 4: 130–133.154. Smales R J. Longevity <strong>of</strong> low- andhigh-copper amalgams analyzed bypreparation class, tooth site, patientage, and operator. Oper Dent 199116: 162–168.155. Lundin S A, Koch G. Class I and IIposterior composite resin <strong>restorations</strong>after 5 and 10 years. Swed DentJ 1999 23: 165–171.156. Burke F J, Cheung S W, Mjör I A, etal. Restoration longevity and analysis<strong>of</strong> reasons for the placement andreplacement <strong>of</strong> <strong>restorations</strong> providedby vocational <strong>dental</strong> practitionersand their trainers in the United Kingdom.Quintessence Int 1999 30: 234–242.157. Raskin A, Michotte-Theall B, VrevenJ, et al. Clinical evaluation <strong>of</strong> a posteriorcomposite 10-year report. JDent 1999 27: 13–19.158. Pyk N, Mejare I. Tunnel <strong>restorations</strong>in general practice. Influence <strong>of</strong> someclinical variables on the success rate.Acta Odontol Scand 1999 57: 195–200.159. Strand G V, Nordbø H, Tveit A B,et al. A 3-year clinical study <strong>of</strong> tunnel<strong>restorations</strong>. Eur J Oral Sci 1996 104:384–389.160. Stoll R, Sieweke M, Pieper K, et al.Longevity <strong>of</strong> cast gold inlays andpartial crowns – a retrospective studyat a <strong>dental</strong> school clinic. Clin OralInvest 1999 3: 100–104.161. Prati C, Chersoni S, Cretti L, et al.Retention and marginal adaptation<strong>of</strong> a compomer placed in non-stressbearingareas used with the totaletchtechnique: a 3-year retrospectivestudy. Clin Oral Investig 1998 2:168–173.162. Pelka M, Schmidt G, Petschelt A.Klinische Qualitätsbeurteilung vongegossenen Metallinlays und -onlays.Dtsch Zahnärztl Z 1996 51: 268–272.163. Friedl K H, Hiller K A, Schmalz G.Placement and replacement <strong>of</strong> amalgam<strong>restorations</strong> in Germany. OperDent 1994 19: 228–232.164. Friedl K H, Hiller K A, Schmalz G.Placement and replacement <strong>of</strong> composite<strong>restorations</strong> in Germany. OperDent 1995 20: 34–38.165. Jokstad A, Mjör I A, Qvist V. Theage <strong>of</strong> <strong>restorations</strong> in situ. ActaOdontol Scand 1994 52: 234–242.166. Smales R J, Gerke D C. Clinicalevaluation <strong>of</strong> light-cured anteriorresin composites over periods <strong>of</strong> upto 4 years. Am J Dent 1992 5: 208–212.International Dental Journal (2001) Vol. 51/No.3


155167. Fritz U, Fischbach H, Harke I.Langzeitverweildauer von Goldgussfüllungen.Dtsch Zahnärztl Z 1992 47:714–716.168. Bentley C D, Drake C W. Longevity<strong>of</strong> <strong>restorations</strong> in a <strong>dental</strong> schoolclinic. J Dent Educ 1986 50: 594–600.169. Smales R J. Effect <strong>of</strong> rubber damisolation on restoration deterioration.Am J Dent. 1992 5: 277–279.170. Smales R J. Rubber dam usagerelated to restoration quality andsurvival. Br Dent J 1993 174: 330–333.171. Letzel H, van ‘t H<strong>of</strong> M A, VrijhoefM M. The influence <strong>of</strong> the condensationinstrument <strong>of</strong> the clinicalbehaviour <strong>of</strong> amalgam <strong>restorations</strong>. JOral Rehabil 1987 14: 133–138.172. Jeffrey I W, Pitts N B. Finishing <strong>of</strong>amalgam <strong>restorations</strong>: to whatdegree is it necessary? J Dent 198917: 55–60.173. Knibbs P J, Plant C G. An evaluation<strong>of</strong> a rapid setting glass ionomercement used by general <strong>dental</strong> practitionersto restore deciduous teeth.J Oral Rehabil 1990 17: 1–7.174. van Dijken J W, Horstedt P. Effect<strong>of</strong> the use <strong>of</strong> rubber dam versuscotton rolls on marginal adaptation<strong>of</strong> composite resin fillings to acidetchedenamel. Acta Odontol Scand1987 45: 303–308.175. Collins C J, Bryant R W, Hodge K L.A clinical evaluation <strong>of</strong> posteriorcomposite resin <strong>restorations</strong>: 8-yearfindings. J Dent 1998 26: 311–317.176. Donly K J, Jensen M E, Triolo P, etal. A clinical comparison <strong>of</strong> resincomposite inlay and onlay posterior<strong>restorations</strong> and cast-gold <strong>restorations</strong>at 7 years. Quintessence Int 199930: 163–168.177. Geurtsen W, Schoeler U. A 4-yearretrospective clinical study <strong>of</strong> Class Iand Class II composite <strong>restorations</strong>.J Dent 1997 25: 229–232.178. McDaniel R J, Davis R D, MurchisonD F, et al. Causes <strong>of</strong> failure amongcuspal-coverage amalgam <strong>restorations</strong>:a clinical survey. J Am DentAssoc 2000 131: 173–177.179. Kerschbaum T, Paszyna C, Klapp S,et al. Verweilzeit- und Risik<strong>of</strong>aktoranalysevon festsitzendem Zahnersatz.Dtsch Zahnärztl Z 1991 41: 20–24.180. Glantz P O, Nilner K, Jendresen MD, et al. <strong>Quality</strong> <strong>of</strong> fixed prosthodonticsafter 15 years. Acta OdontolScand 1993 51: 247–252.181. Dawson A S, Smales R J. Restorationlongevity in an Australiandefence force population. Aust DentJ 1992 37: 196–200.182. Kroeze H J, Plasschaert A J, van ‘tH<strong>of</strong> M A, et al. Prevalence and needfor replacement <strong>of</strong> amalgam andcomposite <strong>restorations</strong> in Dutchadults. J Dent Res 1990 69: 1270–1274.183. Wood R E, Maxymiw W G, McCombD. A clinical comparison <strong>of</strong> glassionomer (polyalkenoate) and silveramalgam <strong>restorations</strong> in the treatment<strong>of</strong> Class 5 caries in xerostomichead and neck cancer patients. OperDent 1993 18: 94–102.184. Bentley C D, Broderius C A, DrakeC W, et al. Relationship betweensalivary levels <strong>of</strong> Mutans streptococciand restoration longevity.Caries Res 1990 24: 298–300.185. Elderton R J. Longitudinal study <strong>of</strong><strong>dental</strong> treatment in the GeneralDental Service in Scotland. Br DentJ 1983 155: 91–96.186. Jokstad A, Mjör I A. Analyses <strong>of</strong> longtermclinical behaviour <strong>of</strong> Class-IIamalgam <strong>restorations</strong>. Acta OdontolScand 1991 49: 47–63.187. Mjör I A, Jokstad A. Five- year study<strong>of</strong> Class II <strong>restorations</strong> in permanentteeth using amalgam, glass polyalkenoate(ionomer) cermet and resinbasedcomposite materials. J Dent1993 21: 338–343.188. Smales R J, Webster D A. Restorationdeterioration related to laterfailure. Oper Dent 1993 18: 130–137.189. Jokstad A, Mjör I A, Nilner K, et al.Clinical performance <strong>of</strong> three anteriorrestorative materials over 10years. Quintessence Int 1994 25: 101–108.190. Valderhaug J, Jokstad A, AmbjørnsenE, et al. Assessment <strong>of</strong> the periapicaland clinical status <strong>of</strong> crowned teethover 25 years. J Dent 1997 25: 97–105.191. McCoy R B, Anderson M H, Lepe X,et al. Clinical success <strong>of</strong> Class V compositeresin <strong>restorations</strong> withoutmechanical retention. J Am DentAssoc 1998 129: 593–599.192. Smales R J, Webster D A, Leppard PI. Predictions <strong>of</strong> restoration deterioration.J Dent 1992 20: 215–220.193. Kallus T, Mjör I A. Incidence <strong>of</strong>adverse effects <strong>of</strong> <strong>dental</strong> materials.Scand J Dent Res 1991 99: 236–240.194. Mjör I A. Biological side effects tomaterials used in dentistry. J R CollSurg Edinb 1999 44: 146–149.195. Stanley H R. Dental iatrogenesis. IntDent J 1994 44: 3–18.196. Borgmeijer P J, Kreulen C M, vanAmerongen W E, et al. The prevalence<strong>of</strong> postoperative sensitivity inteeth restored with Class II compositeresin <strong>restorations</strong>. ASDC J DentChild 1991 58: 378–383.197. Browning W D. Incidence and severity<strong>of</strong> postoperative pain followingroutine placement <strong>of</strong> amalgam <strong>restorations</strong>.Quintessence Int 1999 30:484–489.198. Brännström M. Reducing the risk <strong>of</strong>sensitivity and pulpal complicationsafter the placement <strong>of</strong> crowns andfixed partial dentures. QuintessenceInt 1996 27: 673–678.199. Pashley D, Kim S, Hanks C T, et al.Pathobiology <strong>of</strong> the Dentin/PulpComplex. International conference.Charlotte, North Carolina, May 25–29, 1991. <strong>Pro</strong>c Finn Dent Soc 1992 88(Suppl 1): 571–581.200. Schmalz G. The biocompatibility <strong>of</strong>non-amalgam <strong>dental</strong> filling materials.Eur J Oral Sci 1998 106: 696–706.201. Svanberg M, Mjör I A, Ørstavik D.Mutans streptococci in plaque frommargins <strong>of</strong> amalgam, composite, andglass-ionomer <strong>restorations</strong>. J Dent Res1990 69: 861-864.202. Anonymous. Fixed prosthodontics:avoiding pulp death. Oral Health1995 85: 47–49.203. Gordan V V, Mjör I A, Moorhead JE. Amalgam <strong>restorations</strong>: postoperativesensitivity as a function <strong>of</strong> linertreatment and cavity depth. OperDent 1999 24: 377–383.204. Grieve A R, McAndrew R. A radiographicstudy <strong>of</strong> post-retainedcrowns in patients attending a <strong>dental</strong>hospital. Br Dent J 1993 174: 197–201.205. Mahler D B, Engle JH. Clinical evaluation<strong>of</strong> amalgam bonding in Class Iand II <strong>restorations</strong>. J Am Dent Assoc2000 131: 43–50.206. Bergenholtz G. Evidence for bacterialcausation <strong>of</strong> adverse pulpalresponses in resin-based <strong>dental</strong> <strong>restorations</strong>.Crit Rev Oral Biol Med2000 11: 467–480.207. Kreulen C M, van Amerongen W E,Gruythuysen R J, et al. Prevalence <strong>of</strong>postoperative sensitivity with indirectClass II resin composite inlays.ASDC J Dent Child 1993 60: 95–98.208. Mjör I A. Dentin and pulp. In: Mjör,I A. (ed.). Reaction Patterns in HumanTeeth. pp 63–74. Boca Raton: CRCPress: 1983.209. Qvist V. The effect <strong>of</strong> masticationon marginal adaptation <strong>of</strong> composite<strong>restorations</strong> in vivo. J Dent Res 198362: 904–906.210. Smith C T, Schuman N. Restoration<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


156<strong>of</strong> endodontically treated teeth: aguide for the restorative dentist.Quintessence Int 1997 28: 457–462.211. Ray H A, Trope M. Periapical status<strong>of</strong> endodontically treated teeth inrelation to the technical quality <strong>of</strong>the root filling and coronal restoration.Int Endod J 1995 28: 12–18.212. Tronstad L, Asbjørnsen K, DøvingL, et al. Influence <strong>of</strong> coronal <strong>restorations</strong>on the periapical health <strong>of</strong>endodontically treated teeth. EndodDent Traumatol 2000 16: 218–221.213. Hill A B. The environment and disease:association and causation? <strong>Pro</strong>cRoy Soc Med 1965 58: 295–300.214. Grasso J E, Nalbandian J, Sanford C,et al. Effect <strong>of</strong> restoration quality onperiodontal health. J <strong>Pro</strong>sthet Dent1985 53: 14–19.215. Lang N P, Kaarup-Hansen D, Joss A,et al. The significance <strong>of</strong> overhangingfilling margins for the health status<strong>of</strong> inter<strong>dental</strong> periodontal tissues <strong>of</strong>young adults. Schweiz MonatsschrZahnmed 1988 98: 725–730.216. Coxhead L J. The role <strong>of</strong> the general<strong>dental</strong> practitioner in the treatment<strong>of</strong> periodontal disease. N Z Dent J1985 81: 81–85217. Lang N P, Kiel R A, Anderhalden K.Clinical and microbiological effects<strong>of</strong> subgingival <strong>restorations</strong> withoverhanging or clinically perfectmargins. J Clin Periodontol 1983 10:563–578.218. van Dijken J W, Sjöström S, Wing K.Development <strong>of</strong> gingivitis arounddifferent types <strong>of</strong> composite resin. JClin Periodontol 1987 14: 257–260.219. van Dijken J W, Sjöström S. Theeffect <strong>of</strong> glass ionomer cement andcomposite resin fillings on marginalgingiva. J Clin Periodontol 1991 18:200–203.220. Jenkins W M, Cross L J, Kinane D F.Cervical marginal fit <strong>of</strong> proximalamalgam <strong>restorations</strong>. J Dent 199422: 108–111.221. Brunton P A, Wilson N H. Preparationsfor porcelain laminate veneersin general <strong>dental</strong> practice. Br Dent J1998 184: 553–556.222. Carrotte P V, Winstanley R B, GreenJ R. A study <strong>of</strong> the quality <strong>of</strong> impressionsfor anterior crowns received ata commercial laboratory. Br Dent J1993 174: 235–240.223. Winstanley R B, Carrotte P V,Johnson A. The quality <strong>of</strong> impressionsfor crowns and bridges receivedat commercial <strong>dental</strong> laboratories. BrDent J 1997 183: 209–213.224. Johnson A, Winstanley R B, NortheastS E, et al. Variations in bridgeconstruction in commercial laboratories.Restorative Dent 1991 7: 74–77.225. Axelsson P, Lindhe J. Effect <strong>of</strong> controlledoral hygiene procedures oncaries and periodontal disease inadults. J Clin Periodontol 1978 5: 133–151.226. Hancock E B, Mayo C V, Schwab RR, et al. Influence <strong>of</strong> inter<strong>dental</strong>contacts on periodontal status. JPeriodontol 1980 51: 445–449.227. van Dijken J W, Sjostrom S. Development<strong>of</strong> gingivitis around aged <strong>restorations</strong><strong>of</strong> resin-modified glassionomer cement, polyacid-modifiedresin composite (compomer) andresin composite. Clin Oral Investig1998 2: 180–183.228. Qualtrough A J, Burke F J. A look at<strong>dental</strong> esthetics. Quintessence Int 199425: 7–14.229. Rimmer S E, Mellor A C. Patients’perceptions <strong>of</strong> esthetics and technicalquality in crowns and fixed partialdentures. Quintessence Int 199627: 155–162.230. Abrams R A, Ayers C S, VogtPetterson M. <strong>Quality</strong> assessment <strong>of</strong><strong>dental</strong> <strong>restorations</strong>: a comparison bydentists and patients. Community DentOral Epidemiol 1986 14: 317–319.231. Lambrechts P, Willems G, VanherleG, et al. Aesthetic limits <strong>of</strong> lightcuredcomposite resins in anteriorteeth. Int Dent J 1990 40: 149–158.232. Tyas M J. Colour stability <strong>of</strong> compositeresins: a clinical comparison.Aust Dent J 1992 37: 88–90.233. Verdonschot E H, Oortwijn J C,Roeters F J. Aesthetic properties <strong>of</strong>three type II glass polyalkenoate(ionomer) cements. J Dent 1991 19:357–361.234. Yap A U, Tan K B, Bhole S. Comparison<strong>of</strong> aesthetic properties <strong>of</strong>tooth-colored restorative materials.Oper Dent 1997 22: 167–712.235. Hickel R, Dasch W, Janda R, et al.New direct restorative materials. IntDent J 1998 48: 3–16.236. Gladys S, Van Meerbeek B,Lambrechts P, et al. Evaluation <strong>of</strong>esthetic parameters <strong>of</strong> resin-modifiedglass-ionomer materials and a polyacid-modifiedresin composite inClass V cervical lesions. QuintessenceInt 1999 30: 607–614.237. Johnston W M, Reisbick M H. Colorand translucency changes during andafter curing <strong>of</strong> esthetic restorativematerials. Dent Mater 1997 13: 89–97.238. Rosentritt M, Esch J, Behr M, et al.In vivo color stability <strong>of</strong> resinveneers and acrylic resin teeth inremovable partial dentures. QuintessenceInt 1998 29: 517–522.239. Asmussen E, Hansen E K. Surfacediscoloration <strong>of</strong> restorative resins inrelation to surface s<strong>of</strong>tening and oralhygiene. Scand J Dent Res 1986 94:174–177.240. Martin F E. A survey <strong>of</strong> the efficiency<strong>of</strong> visible light curing units. JDent 1998 26: 239–243.241. Poulos J G, Styner D L. Curing lights:changes in intensity output with useover time. Gen Dent 1997 45: 70–73.242. Forsten L, Mount G J, Knight G.Observations in Australia <strong>of</strong> the use<strong>of</strong> glass ionomer cement restorativematerial. Aust Dent J 1994 39: 339–343.243. Zuellig-Singer R, Bryant R W. Threeyearevaluation <strong>of</strong> computermachinedceramic inlays: influence<strong>of</strong> luting agent. Quintessence Int 199829: 573–582.244. Foster L V. Validity <strong>of</strong> clinical judgementsfor the presence <strong>of</strong> secondarycaries associated with defective amalgam<strong>restorations</strong>. Br Dent J 1994 177:89–93.245. Mjör I A. The reasons for replacementand the age <strong>of</strong> failed <strong>restorations</strong>in general <strong>dental</strong> practice. ActaOdontol Scand 1997 55: 58–63.246. Tyas M J. Correlation between fractureproperties and clinical performance<strong>of</strong> composite resins in Class IVcavities. Aust Dent J 1990 35: 46–49.247. Malament K A, Socransky S S. Survival<strong>of</strong> Dicor glass-ceramic <strong>dental</strong><strong>restorations</strong> over 14 years. Part II:Effect <strong>of</strong> thickness <strong>of</strong> Dicor materialand design <strong>of</strong> tooth preparation. J<strong>Pro</strong>sthet Dent 1999 81: 662–667.248. Lemmens P, Peters M C, van ‘t H<strong>of</strong>M A, et al. Influences on the bulkfracture incidence <strong>of</strong> amalgam <strong>restorations</strong>:a 7-year controlled clinicaltrial. Dent Mater 1987 3: 90–93.249. Smales R J, Gerke D C, White I L.Clinical evaluation <strong>of</strong> occlusal glassionomer, resin, and amalgam <strong>restorations</strong>.J Dent 1990 18: 243–249.250. Malament K A, Socransky S S. Survival<strong>of</strong> Dicor glass-ceramic <strong>dental</strong><strong>restorations</strong> over 14 years. Part I:Survival <strong>of</strong> Dicor complete coverage<strong>restorations</strong> and effect <strong>of</strong> internalsurface acid etching, tooth position,gender, and age. J <strong>Pro</strong>sthet Dent 199981: 23–32.251. Åberg C H, van Dijken J W, Ol<strong>of</strong>ssonA L. Three-year comparison <strong>of</strong> firedceramic inlays cemented with compositeresin or glass ionomer cement.Acta Odontol Scand 1994 52: 140–International Dental Journal (2001) Vol. 51/No.3


157149.252. Klausner L H, Charbeneau G T.Amalgam <strong>restorations</strong>: a crosssectionalsurvey <strong>of</strong> placement andreplacement. J Mich Dent Assoc 198567: 249–252.253. Osborne J W, Norman R D. 13-yearclinical assessment <strong>of</strong> 10 amalgamalloys. Dent Mater 1990 6: 189–194.254. Smales R J, Webster D A, Leppard PI, et al. Prediction <strong>of</strong> amalgam restorationlongevity. J Dent 1991 19: 18–23.255. Wilder A D, May K N, Bayne S C, etal. Seventeen-year clinical study <strong>of</strong>ultraviolet-cured posterior compositeClass I and II <strong>restorations</strong>. EstheticDent 1999 11: 135–142.256. Bayne S C, Taylor D F, Heymann HO. <strong>Pro</strong>tection hypothesis for compositewear. Dent Mater 1992 8: 305–309.257. Özer L, Thylstrup A. What is knownabout caries in relation to <strong>restorations</strong>as a reason for replacement? Areview. Adv Dent Res 1995 9: 394–402.258. van Dijken J W, Kieri C, Carlen M.Longevity <strong>of</strong> extensive Class II opensandwich<strong>restorations</strong> with a resinmodifiedglass-ionomer cement. JDent Res 1999 78: 1319–1325.259. Chestnutt I G, Jones P R, JacobsonA P, et al. Prevalence <strong>of</strong> clinicallyapparent recurrent caries in Scottishadolescents, and the influence <strong>of</strong> oralhygiene practices. Caries Res 199529: 266–271.260. Clarkson J E, Worthington H V.Association between untreatedcaries and age, gender and <strong>dental</strong>attendance in adults. Community DentOral Epidemiol 1993 21: 126–128.261. van Dijken J W. A 6-year evaluation<strong>of</strong> a direct composite resin inlay/onlay system and glass ionomercement-composite resin sandwich<strong>restorations</strong>. Acta Odontol Scand 199452: 368–376.262. Hodges D J, Mangum F I, Ward M T.Relationship between gap width andrecurrent <strong>dental</strong> caries beneathocclusal margins <strong>of</strong> amalgam <strong>restorations</strong>.Community Dent Oral Epidemiol1995 23: 200–204.263. Merrett M C, Elderton R J. An invitro study <strong>of</strong> restorative <strong>dental</strong>treatment decisions and <strong>dental</strong>caries. Br Dent J 1984 157: 128–133.264. Kidd E A, O’Hara J W. The cariesstatus <strong>of</strong> occlusal amalgam <strong>restorations</strong>with marginal defects. J DentRes 1990 69: 1275–1277.265. Söderholm K J, Antonsson D,Fischlschweiger W. Correlationbetween marginal discrepancies atthe amalgam/tooth interface andrecurrent caries. In: Anusavice, K J(ed). <strong>Quality</strong> Evaluation <strong>of</strong> Dental Restorations.pp 95–110. Chicago: QuintessencePublishing Company, 1989.266. Derand T, Birkhed D, EdwardssonS. Secondary caries related to variousmarginal gaps around amalgam<strong>restorations</strong> in vitro. Swed Dent J1991 15: 133–138.267. Kidd E A, Joyston-Bechal S. Essentials<strong>of</strong> Dental Caries. Oxford: OxfordUniversity Press, 1997.268. Kidd E A, Beighton D. Prediction <strong>of</strong>secondary caries around toothcolored<strong>restorations</strong>: a clinical andmicrobiological study. J Dent Res1996 75: 1942–1946.269. Hewlett E R, Atchison K A, WhiteS C, et al. Radiographic secondarycaries prevalence in teeth with clinicallydefective <strong>restorations</strong>. J DentRes 1993 72: 1604–1608.270. Eriksen H M, Bjertness E, Hansen BF. Cross-sectional clinical study <strong>of</strong>quality <strong>of</strong> amalgam <strong>restorations</strong>, oralhealth and prevalence <strong>of</strong> recurrentcaries. Community Dent Oral Epidemiol1986 14: 15–18.271. Hamilton J C, M<strong>of</strong>fa J P, Ellison J A,et al. Marginal fracture not a predictor<strong>of</strong> longevity for two <strong>dental</strong> amalgamalloys. A ten year study. J <strong>Pro</strong>sthetDent 1983 50: 200–202.272. Goldberg G J, Tanzer J, Munster E,et al. Cross-sectional clinical evaluation<strong>of</strong> recurrent caries. Restoration<strong>of</strong> marginal integrity and oral hygienestatus. J Am Dent Assoc 1981 102:635–641.273. Bader J D, Martin J A, Shugars D A.Preliminary estimates <strong>of</strong> the incidenceand consequences <strong>of</strong> toothfracture. J Am Dent Assoc 1995 126:1650–1654.274. Ellis S G, Macfarlane T V, McCordJ F. Influence <strong>of</strong> patient age on thenature <strong>of</strong> tooth fracture. J <strong>Pro</strong>sthetDent 1999 82: 226–230.275. Geurtsen W. The cracked-tooth syndrome:clinical features and casereports. Int J Periodontics RestorativeDent 1992 12: 395–405.276. Sindel J, Frankenberger R, KrämerN, et al. Crack formation <strong>of</strong> allceramiccrowns dependent on differentcore build-up and luting materials.J Dent 1999 27: 175–181.277. Wilder A D, Boghosian A A, BayneS C, et al. Effect <strong>of</strong> powder/liquidratio on the clinical and laboratoryperformance <strong>of</strong> resin-modified glassionomers.J Dent 1998 26: 369–377.278. Kemp-Scholte C M, Davidson C L.Complete marginal seal <strong>of</strong> Class Vresin composite <strong>restorations</strong> effectedby increased flexibility. J Dent Res1990 69: 1240–1243.279. Lang H, Schwan R, Nolden R. DasVerhalten von Klasse-V-Restaurationenunter Belastung. DtschZahnärztl Z 1996 51: 613–616.280. Browning W D, Brackett W W,Gilpatrick R O. Retention <strong>of</strong> micr<strong>of</strong>illedand hybrid resin-based compositein noncarious Class 5 lesions:a double-blind, randomized clinicaltrial. Oper Dent 1999 24: 26–30.281. Friedman M J. A 15-year review <strong>of</strong>porcelain veneer failure – a clinician’sobservations. Compend ContinEduc Dent 1998 19: 625–636.282. Söderholm K J. Does resin baseddentine bonding work? Int Dent J1995 45: 371–381.283. Bayne, S C, Heymann H O,Sturdevant J R, et al. Contributingco-variables in clinical trials. Am JDent 1991 4: 247–250.284. Duke E S, Robbins J W, Snyder D S.Clinical evaluation <strong>of</strong> a dentinaladhesive system: three-year results.Quintessence Int 1991 22: 889–895.285. van Dijken J W. Clinical evaluation<strong>of</strong> four dentin bonding agents in ClassV abrasion lesions: a four-yearfollow-up. Dent Mater 1994 10: 319–324.286. van Dijken J W. Clinical evaluation<strong>of</strong> three adhesive systems in Class Vnon-carious lesions. Dent Mater 200016: 285–291.287. Heymann H O, Sturdevant J R, BayneS, et al. Examining tooth flexureeffects on cervical <strong>restorations</strong>: atwo-year clinical study. J Am DentAssoc 1991 122: 41–47.288. Lambrechts P, Braem M, VanherleG. Evaluation <strong>of</strong> clinical performancefor posterior composite resinsand dentin adhesives. Oper Dent 198712: 53–78.289. Leinfelder K F. Composite resin inposterior teeth. Dent Clin North Am1981 25: 357–364.290. Håkestam U, Söderfeldt B, RydenO, et al. Personality factors versusexpectations and self-reported symptomsamong patients awaitingadvanced prosthodontic treatment.Eur J <strong>Pro</strong>sthodont Restor Dent 1997 5:105–110.291. Lutz F, Krejci I. Direct posteriorfilling materials. In: Vanherle G,Degrange M, Willems G (eds). State<strong>of</strong> the Art on Posterior Filling Materialsand Dentine Bonding. <strong>Pro</strong>ceedings <strong>of</strong> anInternational Symposium. pp 15–28.Leuven: Van der Poorten, 1993.<strong>FDI</strong> <strong>Commission</strong>, Jokstad et al.: <strong>Quality</strong> <strong>of</strong> <strong>dental</strong> <strong>restorations</strong>


158292. Newsome P R, Wright G H. Areview <strong>of</strong> patient satisfaction: 1.Concepts <strong>of</strong> satisfaction. Br Dent J1999 186: 161–165.293. Johnson N W. Risk Markers for OralDisease. Volume 1 Dental Caries.Markers <strong>of</strong> High and Low Risk Groupsand Individuals. Cambridge: CambridgeUniversity Press, 1991.294. Schwartz R S, Summitt J B, RobbinsJ W. Fundamentals <strong>of</strong> Operative Dentistry:A Contemporary Approach. pp33–46. Chicago: Quintessence PublishingCompany, 1996.295. Thylstrup A, Fejerskov O. Textbook<strong>of</strong> Clinical Cariology. 3rd ed. pp 333–353, 393–411. Copenhagen:Munksgaard, 1999.296. Carranza F Jr, Newman M G. ClinicalPeriodontology. 8th ed. Philadelphia:Saunders, 1996.297. Hall W B. Decision Making in Periodontology.3d ed. St. Louis: Mosby,1998.298. Anusavice K J. Decision analysis inrestorative dentistry. J Dent Educ1992 56: 812–822.International Dental Journal (2001) Vol. 51/No.3

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!