You are on page 1of 264
ADAPTATION OF DENTAL AMALGAM A Thesis submitted to the University of Sydney in support of my candidature for the Degree of Noster of Dentol Surgery (1981) dune. bs Ayers Anne L. Symons, B.D.S. Chapter Chapter Chopter Chopter Chopter Chopter Chopter Chopter Chapter TABLE OF CONTENTS INTRODUCTION REVIEW OF LITERATURE The importance of marginal adaption Methods of evaluating marginal adaptation Factors influencing marginal odaptation The relationship of conventional mechanical ‘and physicol properties to marginal edaptation The nature of the interface between tooth ‘and amalgam ORIGINAL INVESTIGATION Scope of the investigation Experimental method Results Conclusions BIBLIOGRAPHY (i) 20. al. 64. 85. 92. 95. 123. 235. ACKNOWLEDGEMENTS I wish to thank Professor George Wing, Head of the Department of Operative Dentistry, for assisting me with this project. I am grateful for the mony valuoble hours he has spent guiding and encouraging me in the completion of this study and making facilities available. I would also like to thank Mr. George Hewitt, my supervisor, Ms. Joonne Aitken for edvising and assisting with the statistical analysis, ond Mr. Denny Sheehy for his work with the illustrations. To the Department of Dentistry, University of Queensland, I om most grateful for the assistance received in the production of this thesis. To my fomily, I am thankful for. their patience. To my husbend, David, I am thankful for his help and encouragement. (iti) ENTRODUCTION Dentol amalgam has been widely researched since the end of the nineteenth century. The investigation of its physical properties lead to the formulation of specifications for this materiol to ensure © restorative material of quality is produced. Research of dental amalgam has been greatly assisted by the advent of the scanning electron microscope, electron probe microanalyserand the improvement in specimen preparation techniques such as polishing and etching. Recent advances have been made regarding the size ond shope of alloy particles and the composition of the original alloy ponder. The introduction of amalgam alloys containing more copper thon the conventionel AgsSn amalgam alloy system has maintained research interest in omalgam as a restorative material. It is generally accepted that the amalgam restoration becomes self sealing after some time of service in the mouth. It is also cccepted that manipulation of the alloy may affect the quolity of the restoration. Investigators using many methods such as dye penetrotion, air pressure and radioisotope tests have been able to detect leckage of the amalgam restoration which appears to reduce with time. Corrosion at the amolgom-tooth interface may produce o corrosion product which “seals” the restoration against microleckage The introduction of high-copper, low corrosive amalgam clloys caused some concern, as the sealing ability of these high copper restorations may be reduced. These restorations also oppeor to be self sealing. (iv) Microscopic exominction of the adaptation of verious amalgam alloys was undertoken in this study to determine quentitatively the effect of alloy type and condensation technique on the adaptation. Very few extensive studies have been undertaken using microscopy following specimen preporation. From the results it may be possible to determine what foctors significontly affect the adaptation of omalgom to tooth structure ond ultimately the duration of the early microleakage of o freshly placed omelgam restoration. REVIEW OF THE LITERATURE CHAPTER 1 THE IMPORTANCE OF MARGINAL ADAPTATION Introduction Restorative Dentistry involves the restoration of lost or diseased tooth structure so that normal form and function are reinstated ond the health of the tooth maintained (Black, 1908). Restorations should not harm the tooth or surrounding structures. A restoration may maintain the health of the tooth by crecting o good marginal seal. The advantages of good marginal adaptetion ore that i aids in the prevention of recurrent caries and may prevent the ingress of bacteria or materials on which bacteria may survive. An ideal restorative meteriol which would completely seal prepared tooth structure has eluded clinicians and dentel reseorchers. As early as 1895 Black emphasised the need for a restorative moterial to adopt to the walls of covity preparations and to be free of dimensional chonge ofter placement. The properties of an idecl restorative material ore: It should be insoluble in oral fluids. It should adapt to the walls of the cavity forming o seal between the tooth structure and the restorative material. It should be free from dimensional change following placement (no dimensional change resulting from the setting process or due to thermal changes in the mouth). It should resist wear. It should be strong enough to resist the forces of mastication. 6 It should be aesthetic in colour and appearence. It should have low thermal conductivity. 8. It should be easily and satisfactorily manipulated for use in the mouth. % It should resist ternish ond corrosion ond be able to be polished (Gilmore and Lund, 1973). Investigators have assessed the lack of adaptation of various restorative materials by examining activity in the microscopic space that exists between the tooth ond restorative moterial. Kidd (1976(a)) defined microleckage as "the clinically undetectable passage of bacteria, fluids, molecules or ions between cavity wall and the restorctive material applied to it". Microleckage has been implicated in the breakdown ond foilure of restorations including recurrent caries, tooth discoloration under the restorative material, hypersensitivity of restored teeth, pulpal damage, accumulation of plaque, percolation and the ultimate breakdown of the filling material. The extent to which restorations using different materials are susceptible to microleakage has been examined by « number of investigators (Fraser, 1929; Armstrong and Simon, 1951; Nelsen et al 1952; Sousen et al, 1953; Seltzer, 1955; Crawford and Larson, 1956 Hompel, 1959; Wainwright et al, 1959; Going et al, 1960 (a), 1960 (b) Phillips et ol, 1961, Swartz ond Phillips, 1961; Mortensen et al, 1965 Roydhouse et al, 1967; Roydhouse, 1968; and Kidd, 1976 (a) ond 1976 (b)). Many techniques have been employed to test the sealing properties of restorations both in vitro and in vivo. The use of dyes, redioactive isotopes, air pressure, bacteric, neutron activation analysis, artificial caries ond microscopic technique: have all been used to test marginal adaptation. Prevention of microleakage could only come about by the development of a restorative material that would truly adhere to tooth structure with a dimensional change comparable to enamel and dentine (Cornell, 1961; Alter and Fookson, 1969; Buonocore, 1973). McLean end Wilson (1977 (a), 1977 (b)) have developed o glass ionomer cement which, like polycarboxylate cement, bonds to enamel and dentine by means of polar and ionic attractions bringing about physicochemical adhesion. The use of this material is limited by its other properties which foll short of ideol. (See properties of an ideal restorative moterial page 1). 1.1 Foctors Affecting Microleakage The extent of microleakage is dependent on the space between the restorative material and tooth structure, presence of defects in the restorative material, state of enomel ond dentine and the power of the driving forces which allow exchange of material along the interface between the tooth and restoretion. Roydhouse (1968) and Elderton (1976 (a), 1976 (b)) found the operator variable influenced the width of the space between tooth and restorative material. An operator who poorly handled a restorative material usually produced lerger marginal gaps. Some materials ore oble to adapt better to tooth structure than others. Amalgam adapts better thon acrylic resins, for exomple. The difference in osmotic, gaseous or vapour pressure between the pulp chamber and oral cavity allows leckage to occur more readily (Roydhouse, 1968). A difference in electrical potenticl moy drive solutes through seemingly intact dental tissues and restoration-tooth interfaces (Roydhouse, 1968). The collection of bacteria and the concentration of substances at the tooth-restoration interface will cause penetration pulpally or release into the oral cavity. There may be more resistance to pulpal penetration in vivo os the vitel pulp could decrease the permeability of dentine and so restrict the extent of penetronts. The rate and extent of microleakage is increased when defects are present in the tooth structure, allowing ready penetration of ions molecules and bacteric (Roydhouse, 1968). By subjecting specimens to thermal cycling a greater exchange of material at the tooth restoration interface occurs as most restorative materials have a different coefficient of thermal expansion from that of enomel and dentine (Nelsen et ol, 1952; Parris and Kapsimalis, 1960) 1.2 Secondary Caries Microorganisms hove been implicated as on oeitiological factor in the production of dental caries. Dento-bacterial plaque forms and is firmly attached to tooth structure. This plaque conteins many acidogenic bacteria. Acids are produced from carbohydrates by the becteria and are of sufficient potential to decaleify tooth structure (Stephan, 1948; Thoma ond Robinson, 1955). It hes been noted that some sites ore more cories susceptible than others (Stephon, 1948) for exomple those creas where cleansing is difficult ollows o buildup ond eging of ploque. Mortensen et al (1965) ond Phillips (1973) recognised that microlekage may be a factor involved in secondary dental decoy. It seems possible that secondary caries may be initiated by the actual penetration of microorganisms between the tooth ond restorative moterial. The penetration of oral fluids containing acids and enzymes between the tooth and restoration may alone also induce recurrent cories or may combine with bacteria to produce a carious lesion. Phillips (1965 (a)) states that the possibility of recurrent cories is increased in the presence of both microorganisms and a suitoble substrate. An effort must be made to inhibit microleckage as success or failure of a restoration may partly rely on the magnitude of the marginal penetration. Recurrent caries due to improper excavation or filling of a carious defect coupled with continual marginal leakage can cause pulpal degeneration and infection (Thoma and Robinson, 1955; Massler, 1965). Kidd (1976 (b), 1978) artificiallyinduced caries along the margins of amalgam, composite and glass ionomer cement restorations. This study linked microleakage with its presumed consequence, demineralisation of the cavity wall. By employing an acidified gel technique lesions were created in human enamel, in vitro, which oppeared indistinguishable from natural caries. The lesions formed consisted of two regions, an ‘outer’ lesion showing features of primary attack ond a ‘cavity wall’ lesion where microleckage oppecred to have taken place between the restoration and the cavity wall ollowing the attack to proceed at right angles to the cavity wall See Figure 1.1. Different restorative materials were assessed for their seal. The depth of penetration of the wall lesion was used os an indicotor of the degree of microleakage for each material. 1.3 Pulpal Effects Pulpal damage can be brought about by cavity preparation and may vary from mild aspiration of odontoblast nuclei to complete necrosis of the pulpal contents (Edwards, 1978). Intratubular contents of the dentine are always affected whenever the dentine is injured and its protoplasmic contents exposed to the oral cavity See Toble 1. . Care must be exercised in the selection of a restorative material, for this moterial will come in close approximation to freshly cut dentine containing many severed odontoblasts with vital protoplasm (Massler, 1965; Cotton ond Siegel, 1978). 1.3.1 Healing and Injury The ability of the pulp to recover from operative injury will rely on the status of the pulpal tissue, extent of injury, the restorotive material employed and the effective seal of the restoration. In cases of a mild reaction increased permeability of cut dentine tubules results. If an inert restorction is placed in the cavity preparation which seals the margin from microleckage repair of dentine and pulpal tissues occurs (Messier, 1965). If the injury is prolonged or severe, effects will spread beyond the subodontoblostic layer calling upon deeper pulpal components to overcome this attack The cell-free layer of Weil is invaded by fibroblasts, mesenchymal cells, smoll copillories and inflammatory cells. These changes may occur under leoking restorations, pain and hypersensitivity moy "plogue” the "restored" tooth. Severe sustained injury, as under a leaking restoration where @ material has been used which may itself be harmful to pulpal tissue, *((9) 926t “PPTA Wosy) SUOTSeT [TOM ,AFTADD, puo ,193n0, 243 skoTdstp ‘unoquos uoFseT sy, || euNBTy eujjoung uj auoz: WOWOA[OAY OUZUEP- quednjsuen Arewyid uno — J97eM UJ UoIsoE uopount yo Apoq eulyne — eupjuep- jeueue- oujoung wy evox 0p 2H uojeoy 110 euoz yueonjsuesy Aseuid euoz yep Wea Ayaeo, uojse1 50 Apog ‘euoz e3ejnS eoapine jousue- Wwasna —> <— WoIAHso Table 1.1 Reaction of Dentin to Injury Mild reaction 4, Increased permeability of dent ‘nal tubules b, Palpo-dentinal membrane dis. «. Odontablastio layer disturbed x Yaruolization bebwean eal r Beginning atrophy of eal ody x Nuclet migrate into tubules ch Complete sicophy of odon- toblastis layer 4, Subodontoblasticcell-trelayer of ‘Weil invaded by 4, Stall round calle 2, Mbroblaste 4, Capillaries ¢. Suhodontablastic body of pulp & Hemorzhage ‘Note: Very mild acute injury ina changes only. «, Tubules become sleroted and Tess permeable , Returns to normal «Formation of ropa ondary) dentin Odontoblestic layer tends to ial morphology, bbut not necessarily 0 ive (0 4. Returns to normal ‘Normal Tn eating with charp bur under water) renal Severe acute injury as caused by application of coustic drugs) results ina +bec+d+e chonges. (From Massler, 1965). % will influence the area of the pulp under the covity preparation showing cellular infiltration, chenges in the ground substances, thrombosis of the nearby vessels and haemorrhage. Death of the entire pulpal tissue moy result from the development of o sterile abscess (Mossler, 1965). In most cases the dentine and pulp will recover from injury by coleificction of the injured portion of the intratubular contents forming an impermeable layer of sclerolic dentine which protects the pulp from further insults. The resting odontoblasts are reactivated or replaced by new odontoblasts from the reserve mesenchymal cells in the pulp, ond a new layer of irregular reparotive dentine is formed. See Figure 1.2. The sclerosis and formation of reparctive dentine constitute the healing mechanism by dentine-forming cells and the emount of dentine laid down on the pulpal side of the dentine is directly related to the amount of dentine removed from the enamel side (Massler, 1965). Severe and prolonged insults which may occur under @ leaking margin, in the recurrent penetrating dentinal caries situotion or by insult from a biologically unacceptable restorotive material results in an increased permeability of dentinal tubules and atrophy of the edontoblastic layer. As seen in Figure 1.2, new reparative dentine is not formed (Massler, 1965). The continued insult spreads into the body of the pulp, causing chronic inflommation, engorgement of the pulpal vessels followed by atrophic changes or necrosis of the pulp. “Continued marginal leakage with secondary recurrent caries is probobly the most common couse of pulpal degeneration and infection”. If the leakage of the restoration is discovered while the pulpol Figure 1.2 Living protoplasm within de tubule hypercaleitied Intertabutar matrix Figure 1.3 Collagen + minerats) 10. Reaction of Dentin to Injury Trotetve tare or liner Selrosis of ‘nbelee EF 8 ling moti | Leoking merging 7 == “No reparative dentin Reaction of dentine to mild acute injury and to severe prolonged injury. (From Massler, 1965) Cross Sections of Young Dentin Old Dentin (Permeable) Vex Portal sclerosis of tubular contents Complete occlusion The Dentin Floor Floor of dentine showing lorge proportion of intretubular cellular living protoplasm in young dentine (left). The introtubuler protoplosm tends to be reduced with age and efter mild injury, by deposition of mineral salts (right). The result is sclerosis of the tubules with grectly reduced permeability (From Massler, 1965) i. defences are still active, removal of the lecking restoration and replacement with a sctisfactory restoration. will usually result in healing of the dentine and pulp, and reparative dentine will continue to form (Mosler, 1965). 1.3.2 Chonges with Age Both enomel and dentine change with age and exposure to irritating substonces. Enomel continues to incredse its mineral content. Young enomel is porous and is easily penetrated by dyes and isotopes but with time becomes almost impenetrable and much more resistant to ocid and bacterial attack. Young dentine tubules are patent ond readily penetrated by large porticles or small bacteria when cut. As dentine ages it becomes more highly mineralised constricted and less penetrable. See Figure 1.3. At the some time more dentine is also deposited rendering the dentine even less permeable ond susceptible to microleakoge ond insult from irritants (Massler, 1965). Dentine acts as a protector to the pulp but if the thickness of dentine is reduced to 0.5 mm or less it con no longer odequately protect the underlying pulpal tissue (Phillips, 1965 (a)). If added irritation such as leakage and irritation from the restorative material ‘continue then pulpal hypersensitivity and eventual death will follow if some other form of protection is not undertaken. Varnishes and cavity linings insulate the pulp from leckoge and irritonts (Massler, 1965; Phillips, 1965 (a), 1965 (b); Kidd, 1976 (a); Evans ond Kasloff, 1976; Younis, 1977). Cavity varnishes ore semipermecble membranes behaving 12. differently in relotion to various types of -ions, permitting some to penetrate freely but inhibiting the passage of others (Phillips 1965 (a), 1965 (b)). Base materials may be required to provide thermal and galvanic protection, reduced microleakage and prevent discoloration of the remaining tooth structure and so enable the pulp to recover from irritation. The type of restorative moterial selected can affect the success of the restorative procedure. Some materials such as silicate cement are known to leak and have a continuing low pH, which couses pulpal inflammation and very often pulpal death particularly in, young teeth (Massler, 1965; Phillips, 1965 (a), 1965 (b)). Lining silicate cement restorotions with calcium hydroxide can overcome this problem. Deep emalgom restorations require a cement base to protect the pulpal tissue from thermal and galvonic shock. Varnish reduces the effects of the early marginal leakage of recently placed amalgam. Improved sealing of some materials in the oral covity with time has been noted. Amalgam restorations in particular are known for their self sealing potential resulting in decreased marginol leckage (Going et al, 1960 (0), 1960 (b); Phillips ond Ryge, 1961; Edwards 1978). In most investigations concerning leakage, the nature of the vital tooth is difficult to study. The vicble pulp in o tooth may control the internal-external balance in flow and content of various ions, solutes and water, thereby controlling the content of dentine end enamel (Roydhouse, 1968). Also the degree of penetration into the dentine os a result of marginal leakage may be dependent on the changes which have taken place in the dentine and the pulp (Going et ol 1960 (a)) 13, 1.4 Percolation The differential thermal dimensional change of a restorative material compared with tooth structure is on important property influencing cdaptotion and microleakage of a restoration. Ideally the coefficient of thermal expansion of the restorative material should be the same os thot of tooth structure. Phillips (1973) states the linear coefficients of thermal expansion for tooth structure and selected restorative materials in Tables 1.2 and 1.3. It can be seen from Toble 1.3 thot acrylic resin will shrink or expand approximately seven times as much as tooth structure for every degree change in temperature (Phillips, 1973) Nelsen et al (1952) investigated the problems created by using acrylic resin as o restorative material. The acrylic resin restoration wos subjected to temperature changes from 9°C to 52°C to simulate the drinking of cold ond hot liquids. The acrylic resin showed thermal shrinkage and expansion causing a change in volume of the space between restoration and tooth. This change in volume would couse fluids to flow into and out of the space around a resin filling when it is subjected to temperature changes encountered in the mouth. The fluid itself also expands and contracts with veriations in temperature. “This pumping action of alternately imbibing and exuding fluids has been termed percolation" (Nelsen et al, 1952; Phillips, 1973). Seltzer (1955), Crawford ond Lerson (1956), Going et al (1960 (b)), Parris ond Kapsimalis (1960) ond Mossler (1965) reported that thermol cycling increased microleckage os penetration of dyes and radioactive isotopes was enhonced with tempereture changes. Toble 1.2 Linear Cocficients of Thermal Expansion of Some Important Dental Materials* Tana Comrie ‘or Exranston Marea aman 31) “Tooth (acres erow)| M4 Sine cement as Deveal smaigatn 80 Porelin 41 ental resi (ply(methyl methacrylate sto (From Phillips, 1973) Table 1.3 Differential Thermal Dimensional Change between Tooth Structure and Verna Factor reas fos mati (Siero ‘er vein 9) =e Ara A cia ny + Gat ot ‘9 03 Siete coment ee eee ESE CEEEEE En SeEEEEEEEEY “1 linea eofcien of expansion. (From Phillips, 1973) 15. Mortensen et al (1965) were of the opinion thet variations in the incidence of microleakage, observed between different restorative materials they tested, could be explained partly by their different coefficients of thermal expansion compared to tooth substance. Parris and Kapsimalis (1960) calculoted the crevice width to be 50 micrometers for acrylic resins which is 5 to 20 times as lorge as the size of bacteria commonly found in the mouth. Seltzer (1955) used bacteria as a penetrating agent and his experiments showed thermal cycling permitted the penetration of microorganisms into the margins of restorations. Going et al (1960 (b)) questioned the extent of fluid exchange. Fluid exchange may not be limited to the crevice between tooth and restoration but may extend into the dentine ond even to the pulp chomber. This emphasises the need to seal the dentinal tubules under all restorations to protect the pulp. The pulp may try to overcome the effect of irritants by inducing changes in density ond thickness of dentine which is seen under old restorations. Beside percolation resulting from thermal expansion (Nelsen et al, 1952) other chemical and physical forces might increase or act independently in causing this seepage. Some other forces may be ‘copillarity, dialysis, diffusion and changes in hydraulic or gas pressures (Nelsen et al, 1952) Mossler (1965) was of the opinion that the only way to overcome percolation and subsequent microleakage is by means of an adhesive ond flexible film long the cavity margin which would seal during expansion and contrection. 1.5 Staining Microleckoge may be responsible for staining of tooth structure Bacteria, ions, heavy metals and organic motter diffuse into the restoration - tooth interface and may react with, be absorbed or deposited into tooth structure or the filling moteriol itself ond so cause marginal staining. Kidd (1976 (b)) recognised that stoining of composite resin ond its morgins was a cause of foilure of this restoration. Etching the dental enamel and coating the surfoce with an unfilled resin hes shown a reduction in staining and the omount of microleakage (Kidd, 1976 (b); Buonocore, 1973) Acrylic resin restoration margins undergo colour change and breakdown which may be related to its gross leckage characteristic (Phillips, 1965 (a)) Phillips (1965 (a)) ond Mossler (1965) both recognised the poor sealing property of silicate cement due to shrinkage and slow dissolution of the matericl itself. Crozing and colour change of silicate cement occurs but despite poor marginal adaptation recurrent caries around these restorations is rare. This cannot ‘be explained fon the basis of superior adaptation of the cement to tooth structure as the marginal leakage is severe for this material. Fluoride ions migrate from the silicate cement into the surrounding enamel, slightly staining it, but bringing obout a reduction in the enomel solubility rendering it more resistont to acid attack and consequently secondary caries. 7. Amalgam restorations frequently stain the surrounding tooth structure. The self sealing property of amalgam restorations by the deposition of corrosion products along the tooth-restoration interfoce hes been well documented (Massler and Barber, 1953; Going et ol, 1960 (b); Lyell et al, 1964; Mossler, 1965; Phillips, 1965 (b) Pickord ond Gayford, 1965; Kurosaki and Fusayama, 1973). The collection of corrosion products block margins preventing marginal leckage. Pickard and Gayford (1965) suggested occasional dislodgement of corrosion substances may render the interface susceptible to further microleckage until more corrosion products occlude the margins. Massler and Barber (1953) and Kurosoki and Fusayamo (1973 found a high percentage of teeth restored with omalgan were stained. Mossler and Barber (1953) considered the green-black discoloration under omalgom fillings "probably involved a slow diffusion of soluble metallic ions through the dentine under the influence of intermittent galvanic action arising from within the omalgom filling itself, with subsequent precipitation of the metallic ions as dark, insoluble metallic sulphides. The sulphides probably are derived externally from the saliva". Brown stains under old amalgams were noted ond soid to be due to residual or recurrent caries. Some doubt exists as to the above classificotion following o paper by Kurosaki end Fusayoma (1973) which showed that softened dentine (not norma: dentine) allowed the penetration of tin ond zinc, probably Precipitating in place of calcium. The tin did form @ sulphide ond consequently forma black stain. To reduce or even prevent discoloration of tooth structure around amalgam the removal of all softened dentine is necessary, occompanied by the use of varnishes and if required a bose, such os accelerated zinc oxide and eugeno! preperetion or zine phosphate cements. 1.6 Plaque Accumulation Plaque is able to collect in acquired or natural defects, or may attach itself to smooth surfaces of enamel by the enamel pellicle. Going et ol (1960 (b)) found good correlation between marginal penetration and surface plaque formation and etching. Marginal penetration by dye and radio active isotope showed greater leckage os the enomel approached the cervical area. The reduced thickness of enomel in this area must also be considered. The relevance of grecter leakage cervically is important in cases where margins are finished in this area, An acrylic crown is likely to accumulate plaque which produces toxins and so cause subsequent irritation to gingival tissues Silicate cement restorations erode ond breokdown, but the fluoride component may inhibit the octivity of plaque and the formation of plaque around the restoration. Glass ionomer cements seal the margins and so prevent the accumulation of plaque. Plaque increases the permeability of margins and is implicated in secondary caries. ‘Summar Marginal secling is an important feature of restorative procedures because it reduces microleakage of the tooth-restoration interface. Sealing the restoration margin may reduce the risk of recurrent caries, hypersensitivity, discoloration ond prevent ploque eceumulating in the tooth-restoration space. It will enable the 19. tooth to recover from structural loss and caries, and aid in maintaining the health of the tooth. 20. CHAPTER 2 METHODS OF EVALUATING MARGINAL ADAPTATION Many techniques have been developed to evaluate the morginal adaptation of restorations. These studies demonstrate that the morgins of restorations are not fixed, inert and impenetrable borders but microcrevices which allow exchange of ions and molecules (Going, 1972; Going, 1979). Roydhouse et al (1967) and Broadhurst (1984) recognised that the interaction of several foctors affected the degree of leakage round restorations. Adaptation is affected by performance of the operator, behaviour of the testing penetrant, pressure differences between the external surface and the external face of the dentine ond the restorctive material used. Most tests used to determine microleokage have been performed in vitro. It must therefore be recognised that the development of useful testing methods is only able to show microleckage quantitatively. Testing methods include:~ To Dyes 2. Rodioactive Isotopes Bocteria Air Pressure Marginal Percolation Artificial Caries Measurements using Microscopy and Replication Neutron Activation Analysis Electron Microprobe 2. 10. Ton Beom Milling 11, Conductimetric Technique 2.1 Dyes The use of dyes to study microleakage around restorations is one of the oldest and most commonly used techniques. Many dyes have been used, aniline dye, basic fuchsin, chromotrope 2R, crystal violet dye, eosin, fluorescin (fluorescent dye), methyl violet, methyl blue hematoxylin and mercuric chloride and Frontosil soluble red ore examples (Going, 1972). In general the method of using dyes, consisted of placement of a restoration in an extracted tooth which is then immersed into o dye solution for a certain time interval. The tooth is then removed, sectioned and examined to determine the extent of dye penetration (Going et al, 1960 (b); Going, 1964; Lyell et al, 1964; Pickard and Gayford, 1965; Roydhouse et al, 1967; Roydhouse, 1968; loiselle et al, 1969; Grieve, 1971; Buonocore, 1973; McCurdy et al, 1974; Kidd, 1976 (b); Khera and Chon, 1978; Smith et al, 1978; Martin, 1980). Fluorescent dyes have been shown to be useful tracers ond many researchers have favoured this penetrant (Loiselle et al, 1969 Grieve, 1971; Going, 1972; McCurdy et al, 1974; Smith et al, 1978) These fluorescent dyes are useful as tracers for the demonstration of leakage around dental restorations becouse "they are detectable in dilute concentrations, are sensitive to ultra violet light, ore easy to photograph, permit more reproducible results, ore inexpensive, contrast sharply with the natural fluorescence of teeth, require short immersion periods, permit direct observation of the total marginal interface during evaluation and scoring of marginal 22. leakage, are nontoxic, permit clinical as well os laboratory investigations, ond can be used apically ond systemically as o tracer in adult humans" (Going, 1972). In most dye studies the teeth are sectioned to examine penetration. A system of classifying the amount of leakage is assigned to each restoration. Unfortunately there seems to be no universal system, eoch researcher has determined his own quantitative method of assessing leakage. This makes the comparison of different studies difficult. Comparing and matching results is also prone to variance as each researcher determines subjectively the extent of dye penetration. Pickard and Gayford (1965) pointed out that it is difficult to prevent the dye from coming in contact with areas other than the morgins of the restorations. It hos been shown that etched enamel ond defects in the enamel surface will allow the penetration of dyes (Going et al, 1960 (b)). Also the diffusibility of the dye renders the discoloration of the cavity wall hard to interpret (Pickard ond Gayford, 1965). Pickard and Gayford (1965) found that severe leakage ot one point could completely mosk lesser penetration nearby Khera and Chan (1978) examined microleakage and enamel finish in ‘Clos V restorotions. Dye penetration increased in the mesiol ond distol extremities but a section token through the central region exhibited less leckage. Consistency is therefore necessary in sectioning for results to be relevant. Destruction of the specimen is usually necessary so it is not possible to examine behaviour over @ period. As most are in vitro studies it is hord to determine behaviour under oral conditions (Pickard ond Gayford, 1965). 23. Attempts have been made to simulate oral conditions, such as: thermal cycling, storing in oral fluids or placing restorations in vivo before examination (Lyell et al, 1964; Roydhouse et al, 1967 Roydhouse, 1968; Loiselle et al, 1969; McCurdy et al, 1974; Smith et ol, 1978). Roydhouse (1988) sectioned teeth and pushed out the restorative material to examine leckage patterns on the cavity walls. Four leckage patterns were observed: (a) dye penetration directly to the pulp chamber and no single channel being visible (b) leakage restricted to the enamel-restoration interface due to small deficiencies at the cavo-surface angle (c) enamel leaks; where deficiencies in the enomel allowed penetration of dyes (d) dye penetration by permeation or diffusion throughout the enamel and dentine. The type of penetration was dependent on operator, material used, the penetrating fluid and driving forces. Dyes ore not chemically reactive, they ore inert and so penetrate by simple diffusion (Broadhurst, 1964). Dye concentration length of exposure ond moleculer size will influence penetration (Broadhust, 1964; Going, 1964; Pickard and Gayford, 1965). The results of in vitro experimentation of microleakage remain speculative os in vivo factors such os the oral environment and pulpal hydrostatic pressure have not been taken into account (Roydhouse, 1968). 2.2 Radioactive Isotopes Solutions of radioactive isotopes have been used in @ similor monner to dyes. Radioactive isotopes are able to penetrate more readily than dye tracers. By using the autoradiographic technique minute quantities of the penetrant can be detected on the film that would not otherwise be detected visually. The outoradiographs can also act os @ permanent record (Pickard and Gayford, 1965; Going 1972; Kidd, 1976 (a)). Restorations are placed in teeth, most of the surface is sealed except for the restoration margins, specimens are then exposed to the radioactive isotope penetrant. A most conmonly used radioactive isotope is the 0.l-m Ci/ml solution of calcium chloride - a radioactive colcium Ca*®. Ca! penetrates selectively and deeply into marginal defects enabling the production of clear sharp autoradiographs (Armstrong and Simon, 1951; Sousen et ol, 1953; Crawford ond Larson, 1956; Hampel, 1959; Going et cl, 1960 (a) Phillips et ol, 1961; Swartz ond Phillips, 1961; Swartz, 1962 Lyell et ol, 1964; McCurdy et ol, 1974; Kidd, 1976 (a); Andrews and Hembree, 1978; Boyer and Torney, 1979; Andrews and Hembree, 1980). After exposure to the penetrant, the teeth are washed, sectioned and placed against an x-ray film. The autoradiographs are examined, radioactive isotope penetrotion is then assessed. Experimental results are quantified by assigning numerals to defined depths of penetration. Going et al (1960 (a)) used a variety of radioactive isotopes to examine the marginal penetration of dental restorations. This study tried to determine whether ionic charge and chemical 25. reactivity influenced the degree of marginal penetration. Going et al (1960 (a)) varied the pH from 5 to 7 but no significant differences resulted. Each restoretive material reacted differently with isotopes and sometimes absorption of cn isotope by the material wos noted. The charge on the ion and its chemical affinity greatly influenced its absorption on to tooth and restorative surface as well as its penetrability through the tooth-restoration interface. Wainwright et al (1959) also reported that different isotopes gave different penetration records. This may be due to a difference in the molecular size of the radioactive isotopes. The smaller molecules wil penetrate more readily thon larger molecules. An increase in the concentration of the radioactive isotope will result in more penetration as will on increased length of exposure. The state of the tooth structure influences penetration. Thin, flowed or etched enamel is more susceptible while sclerotic dentine reduces penetration (Going et al, 1960 (a), 1960 (b); Going, 1964). Cavity preparation con also influence microleakage (Hompel, 1959) To simulate orel conditions specimens were subjected to thermal cycling, aging, storing in oral fluids or by restorations placed in vive which were extracted after on interval. Thermal cycling (Pickard ond Gayford, 1965; Andrews and Hembree, 1980] increases the permeability of margins when the coefficients of thermel expansion of tooth ond restorative materials ore mismatched. Lyell et al (1964 reduced the marginal leckage of teeth with amalgom restorations by storing them in salive ond sulphide solutions. A possible explanation is that corrosion products accumulated in the tooth- 26. omalgom interfoce blocking it to microleakage (Pickard and Gayford, 1965). Aging the restorations in vitro ond in vivo followed by examination of microleakoge indicates the character of the tooth- restoration interface has changed. Amalgam improves its seal with time where as the seal of “temporary stopping” material becomes worse (Phillips et ol, 1961; Andrews and Hembree, 1978). The marginal adaptation of restoration ofter time and service in the mouths of humans and dogs was studied by Phillips et ol (1961). NeCurdy et al (1974) compare in vivo and in vitro microleakege. Although both studies were done under oral conditions, the clinical significance of penetration by tracer remains speculative. The radioactive isotope method of assessing microleckage i: widely used but is difficult to stendordise. Results will change according to operator variable, cavity preporation, monipulation of materials (Sausen et al, 1953), restorative material (Armstrong and Simon, 1951) and type of isotope. When sectioning specimens they must be of equal ond standard width (Going et ol, 1960 (a)) Variation of immersion time (Going et al, 1960 (b); Phillips et ol 1961; Pickard and Gayford, 1965; McCurdy et al, 1974), isotope used pH of solution, molecular size, charge on the ion ond its chemical offinity for the filling material ond tooth surface, and differences in techniques will alter isotope penetration (Going et al, .1960 (b)). There ore limitotions to this procedure os penetration of tracer is only assessed at the points where sectioning occurs ond severe penetration con obscure light penetration (Pickard ond Goyford, 1965). Unlike many others McCurdy et al (1974) was able to compare 27. fovourebly in vitro and in vive microleakage using radioactive isotopes. These in vitro tests were limited as no temperature variations were used. In any study the behaviour of dyes ond isotopes penetrating into dentine must be interpreted with caution (Pickard and Gayford 1965). Most methods do not take into account the biological reactions in dentine ond pulp consequent to cerious invesion and the insertion of o restorotion. 2.3 Bacterio Using bacteria to test for the presence of microleckage is clinically orientated as microorganisms have been implicated in the etiology of cories ond recurrent decay. In 1929 Fraser examined o variety of restorative materials concerning their germicidal ond sealing potentials. Some restorative materials inhibited bacterial growth initially owing to the free acid component. The permanent restorative mate Is exhibited better secling then temporary Filling materials. Seltzer (1955) and Mortensen et al (1965) tested the penetration of colour producing microorganisms through the margins of restorations at constant temperature and ofter thermel cycling. Although penetration was not seen in all specimens increased penetration was noted with thermal cycling. With thermal cycling each restorative material showed its own characteristic penetration pottern, this is likely to be associated with the coefficient of thermal expansion. 28. Brown et al (1962) diffused niacin, in vitro, from the pulp chamber through dentine tubules and was able to support bacterial growth on the exposed external dentine surface. This diffusion of nutrients to support bactericl growth could act os @ means of inducing microorganisms to penetrate dentine. The dentine tubules containing the invading test microorganisms showed dentinal changes resembling clinicol caries. Bacterial penetration studies of the tooth-restoration interface ore a more clinically orientated test for leckage, becouse they con be related to the caries process and recurrent decay (Mortensen et al, 1965; Going, 1972). The results are described on o quolitetive basis (Going, 1972; Kidd, 1976 (a)) which is a disadvantage as no quantitative information is provided. Kidd (1976 (a)) argued that the microleakage detected is gross if bacterial size is compared with that of a hydrogen ion. This could explain why Seltzer (1955) found no penetration by bacteria of restorations thot had not undergone thermal cycling. When bacterial penetration is detected it confirms microleakage of porticulate matter or aggregates of mocromolecules such as large enzymes and microorganisms hes token ploce (Mortensen et ol, 1965). "2.4 Air Pressure The air pressure laboratory test wos used os early os 1912 by Horper to examine leakage paths cround restorations. The method involves placement of Class V restorations, as these are more easily stondardised, in freshly extracted teeth. Air is delivered to the bese of the restoration via the pulp chamber, The teeth ore immersed in water, air pressure is increased ot the floor of the restoration 29. ond the emergence of air bubbles at the cavo-surface margin detected ond assessed using a disecting microscope (Seltzer, 1955; Pickard and Gayford, 1965; Granath, 1971; Kidd, 1976 (a)). See Figures 2.1, 2.2 and 2.3 The advantage in the use of compressed air for leakage studies lies in the precision of the method ond the quantitation of the data (Seltzer, 1955; Pickard and Goyford, 1965; Going, 1972). Reproducible results around the periphery of restorations have enabled a comparison of different materials. Microleckoge could be examined over a period of time os destruction of the specimens was not necessitated by this method (Pickard and Gayford, 1965; Going, 1972; Kidd, 1976 (a)). Fiasconaro and Sherman (1952) studied different moterials and were able to compare their sealing properties. Acrylic resins leaked at 6-8 pounds per square inch (p.s.i.), silicote at 38 p.s.i., gold inlays ot 45-50 p.s.i., and ot a limit of 50 p.s.i., amalgam gold foil and zine phosphate showed no leakage (Going et al, 1960 (b) Pickard and Goyford, 1965). Pickard and Gayford (1965) used air pressure to the limit of 900 nm mercury to show points of leakage for amalgom restorations Readings of microleakoge were token twice on the doy ofter insertion, daily for one week ond weekly for ten weeks. A plon of the periphery of the restoration wos drown and the leaks located, each leak wos treoted seperotely. Pickard ond Gayford (1965) then calculated the “calibre” of pathways through which leakage occurred and was oble to study in detoil o single leckage pathway. A lorge leck did not obliterate nearby smaller leaks os a different sized pathways had 30. Maucan nycon Figure 2.1 Section of the prepared tooth, showing detoils of preparation. Note nylon suture extends from base of the restoration along the needle (From Pickard and Gayford, 1965). \ MICROSCOPE AIR PRESSURE Lp Figure 2.2 Specimen under investigation (From Pickord ond Gayford, 1965). Figure 2.3 Apparatus for meosuring criticol pressure (From Pickard ond Gayford, 1965). 32, different critical pressures and by differing the angulation of observation, examination was made more selective. Generally, an initiel sharp drop in total effective leakage after a few days was followed by a slower fall over the next few weeks. This was accredited to corrosion product formation or microbiological growth obstructing pethways, capillary size paths soon become clogged and stagnant Sudden dislodgement of occluding matericl could reopen pathways and 50 cause an increase in the overall leakage of the restoration. Air pressure seeks out actual defects of continuity and air does not diffuse into dental tissue, As it is limited to pothways which connect the cavo-surface angle to the pulp blind end defects would not be demonstrated. Results might seem decisive in the determination of microleakage, but air pressure is a severe test confined to in vitro use and does not simulate conditions present in the tooth or mouth (Seltzer, 1955; Going, 1972). The passage of oir through paths removes the occluding material and reduces the rate of formation of corrosion products (Pickard and Gayford, 1965). Leckage of compressed air does not necessarily indicate thot a filling would fail under normal conditions in the mouth, but this method provides qualitative and continuous measurements of leakage paths (Seltzer, 1955; Pickard and Gayford 1965). 2.5 Marginal Percolation The ideal restorative material should retain its size ond shape once placed, but very few restorotive materials have a coefficient of thermal expansion similar to thot of tooth structure. Mony investigators have shown that microleckage is enhanced when teeth containing restorations are subjected to temperature change (Nelsen et al, 1952; Seltzer, 1955; Parris and Kapsimalis, 1960; Mortensen et al, 1965; Roydhouse et al, 1967; Granath, 1971; Going, 1972; Kidd, 1976 (a), Smith et al, 1978; Andrews and Hembree, 1980). Microleokoge after thermal cycling is studied using some form of penetrant such as bacteria, air pressure, dyes and isotopes. It is thought that thermal changes cause a deterioration of the margins of a restoration by plastic deformation of filling and tooth due to irregular expansion ond contraction (Roydhouse et al, 1967, Granath, 1971). From this it can be seen that the linear coefficient of thermal expansion of a restorative material should match as nearly as possible thot of tooth structure. So far, only silicate cement comes close to accomplishing this balance (Guzman et al, 1969; Going. 1972). Nelsen et ol (1952) exomined amolgom, gold inlay, silicote cement, gold foil, gutta percha and zine oxide eugenol restorations Teeth contoining the restorations were imnersed in iced water, dried and viewed under a binocular microscope. As teeth were warmed in the fingers droplets of fluid were seen to exude from the morgins no matter what restorative material was used. Marginol percolation was caused by a difference in the coefficient of thermel exponsion between dental tissue ond the restorative material and by thermal expansion of fluids occupying the crevice between the tooth ond restoration. Calculation of the size of the tooth-restoration interspace was undertoken by Nelsen et al (1952). They considered that a channel of ten micrometers was a reasonable estimate, it would allow exchange of microorganisms and products increasing the susceptibility of the tooth to secondary caries. Parris ond Kapsimolis (1960) stated that 34. © crevice of ten micrometers is five to twenty times as large as the Size of bacteria commonly found in the mouth. Nelsen et al (1952) placed the limits of temperature tolerance in the mouth between 4°C and 60°C. More recent studies by Plant et al (1974) estimate on upper limit of 58°C for liquids to be taken comfortobly in the mouth. The tempercture extremes at the tooth surface was found to be between 45°C and 15°C. Some form of thermal stressing should be incorporated in microleckage testing as it is desiroble thot the experimental model be subjected to conditions similar to those of the oral environment. 2.6 Artificial Cories The frequency with which recurrent cories is present around the margins of restorations suggests that the enomel-restoration interface is a preferential pathway for ¢ carious lesion. The development of on acidified gel technique to produce caries in vitro ond in vive has been able to link recurrent caries with microleckage (Ellis and Brown, 1967; Rodda, 1970; Hals and Simonsen, 1972 Grieve, 1973; Kidd, 1976 (a), 1976 (b); Kidd, 1978). Lesions produced using this acidified gelotine gel were indistinguishable from notural caries. The artificial lesions consisted of two ports: (a) "outer" surface lesion showing primary attack of the enamel surface (b) "cavity well” lesion which penetrated enamel and dentine adjacent to the restoration. The “cavity wall" lesion is formed by the microleckage of hydrogen ions from the acidified gelatine gel into the restoration-tooth 3s. interface (Mortensen et ol, 1965; Rodda, 1970; Hals ond Simonsen 1972; Kidd, 1976 (0), 1976 (b), 1978). The depth of lesion penetrotion and degree of demineralisation is used to calculate microleakage. Assessment is subjective which is o disadvantage (Kidd, 1976 (o)). Other investigators have been able to show the presence of microorganisms in the tooth-restoration interface Bacteria housed in marginal defects are able to produce acids and so encourage recurrent cories (Thoma and Robinson, 1955; Broadhurst. 1964; Lyell et al, 1964; Mortensen et al, 1965) Oceluding the microspace between restoration and tooth by use of a liner or varnish reduces the degree of demineralisation in the cavity wll. This supports the hypothesis that hydrogen ions aid in the deminerclisation of the cavity wall os sealing the margins reduces secondary caries (Ellis ond Brown, 1967; Grieve, 1973) 2.7 Measurements using Microscopy ond Replication The adaptation of o restorative material to o patterned plate hos been used to assess the quality of adaptation. The ability of the restorative material to copy a pottern is considered to be directly related to its degree of adaptation. By varying line angles from 30° to 90° it hes been shown that cavity preparation can influence odoptotion. Results have been recorded using tracings made of the adopted surface, by measuring surface roughness, measuring the gops at the tooth-restoration interface with microscopy and by diffraction patterns obtcined with a He-Ne loser (Massler and Barber 1953; Hott, 1959; Jérgensen, 1965 (a); Azar et al, 1968; Mitchem and Mahler, 1968; Lee and Swortz, 1970; Grieve, 1971; Conway and Baumhommers, 1972; Jérgensen, 1972; Kurosaki and Fusayama, 1973 36. De Rijk, 1976; Elderton, 1976 (b); Oilo, 1976; Chan et al, 1977) Hatt (1959) using © Telysurf, Surface Analyser found that complete opposition of omolgam to covity wall could not be achieved irrespective of the manipulation technique. Grieve (1971) incorporated dye ond microscopy to exomine adaptation but like Lee ond Swartz (1970) who used isotopes could find no direct correlation between visible marginal defects and demonstrable leakage. Wing and Lyell (1966) placed omalgom restorations in Class I cavity preparations cut in freshly extracted teeth. The teeth were then mounted for metallographic specimen preparation and longitudinal sections were made. Diomond obrasive pastes were used for final polishing to avoid Flow of the amelgom. Polished specimens were later etched. Specimens were exomined using o reflected light microscope, the junction between the tooth ond restoration was examined and the spaces measured using a filar eyepiece. Spaces existed between the amalgam restoration and tooth structure. The spaces were reduced by good condensation but were never completely eliminated. Scanning electron microscopy provides a means of direct visual observotion of morginel adaptation and surface characteristics. The restoration is sacrificed for examination by this method. The potenticl for ortifoct formation created by the vacuum and specimen preparation con influence results. Replication has overcome this problem and provides 0 continuous evaluation of marginal integrity as there is no destruction of the specimen and replicas can be made of restorations in vivo (Charbeneau and Peyton, 1957; Grundy, 1971; Going, 1972; Elderton, 1976 (b); Kidd, 1976 (a); Saltzberg et al, 1976; Schoen et al, 1978; Going, 1979). 37. Scanning electron microscope observation can only measure the tooth-restoration interspace and is not orientated to diffusion and penetration (Going, 1972; Kidd, 1976 (a)) so results are difficult to quantify. As no correlation could be established between dye ond isotope leckage ond the degree of cavity adaptation, improved tests cre needed. 2.8 Neutron Activotion Analysis Neutron activation analysis has been used in vivo and in vitro to determine microleckage. Restored teeth are socked in an aqueous solution of non-radioactive manganese salts, the surfoces ore washed to remove excess salts. The technique when applied in vivo requires the isolation by rubber dom of the teeth to be examined to prevent the solution contacting other oral tissues. For observation it is necessary for teeth to be extracted. Whole teeth are placed in the core of o nuclear reactor, manganese Wn°® is activated to Nn®® ond the x-ray emission of tn’ formed during irradiation meosured (Going 1972; Dennison et al, 1974; Meyer et al, 1974; Kidel, 1976 (a); Going, 1979). The calculated uptake is expressed in micrograms of manganese per tooth. Generally the in vivo uptake is greater than the in vitro uptake. The increased diffusion around in vivo specimens may be more influenced by material differences related to a functional environment than to the testing environment alone. “Temporary stopping" showed reasonable sealing in vitro but mangonese detection was increased by « factor of six for in vivo specimens (Going et al, 1968; Going, 1972). An advontage of this system is that results con be quantified and the in vive exominetion is possible. Going (1972) considered 38. the distribution of manganese in his study suggested considercble penetration of underlying dentine ond possibly pulp, os a result of marginal leckoge. Meyer et al (1974) and Kidd (1976 (a)) recognised the discdventages of this technique. The teeth exomined had to be extracted, the path and depth of tracer was not well defined unless serial sections were made, the method is costly and the dental researcher required the assistance of a nuclear engineer. Mangonese is token up by the tooth ond restorative material which causes a variability of results, a need for on alternative tracer was suggested Meyer et al (1974) tested a variety of neutron activated trecers (Dysprosium, Manganese, Vanadium and Indium) and found Dysprosium proved to be @ most acceptable tracer os it 1, had high selectivity, allowing analysis to be carried out on whole unsectioned teeth, 2. wos ropidly active and possessed a fast counting procedure, 3. was highly reliable and 4, wos highly sensitive Dennison et al (1974) have since used Dysprosium to examine microleakage associated with pit ond fissure sealonts. 2.9 Electron Microprobe This method may be used in the future to study restorations thot have been in service in the mouth for a known period of time Diffusion of o selected element such as manganese through the margins of a restoration in vivo will allow microprobe analysis after the tooth hos been extracted. Microprobe analysis should be able to define the path of diffusion of manganese into tooth structure determine depth of penetration and provide quantitative anclysis ot given locations (Going, 1972). 2.10 Ion Beam Milling Precise serial sectioning of teeth and restorations by ion beam milling could provide the least distorted view of the creo being onalysed. This method eliminotes artifacts of the presently used sectioning systems. Ion beam milling could be used in conjunction with scanning electron microscopy, electron microprobe and neutron activation analysis (Going, 1972) 2.11 Conductimetric Technique Dimensional changes for some anterior restorative materials have been demonstrated in on artificial cavity (Jacobsen et al, 1975). "The cavity woll-restoration interspace was incorporated into an electro-chemical cell ond the chenges in the current passing through this cell reflected chonges in the dimensions of the interspace” The actual dimensions of the interspaces cannot be assessed unless the circuit is colibroted using an interspace of known dimensions which is o difficult procedure. At this stage the conductimetric technique con only be used for non conductive restorative materials. 2.12 IN VITRO Testing Versus IN VIVO Testing ’ Some attempts have been made to age restorations in vivo prior to extraction (Going et al, 1960 (b); Phillips et ol, 1961) most microleckage tests have been carried out in vitro. The value of results from in vitro testing is questionable as mony vital oral foctors are eliminated such as pulpal hydrostatic pressure, plaque, 40. soliva and mosticotory forces. In most cases in vive and in vitro results tend to be conflicting. Fluorescein dye studies by McCurdy et al in 1974 showed good correlation between in vivo and in vitro results while Loiselle’s et al (1969) in vivo and in vitro results contrasted so greatly they considered all microleakage assessments Stuever et al (1971) used should be carried out in viv endodontically treated teeth in vivo and tested with fluorescein dye The leckage score for these teeth was comparable to the leakage score of restored teeth tested in vitro, thus showing that pulpal hydrostatic pressure may influence microleckege. Summary No method of testing marginal adaptation provides information which is conclusive. All methods cre subjective. Even the development of highly technical procedures, such as the neutron activation system, is uncble to determine microleakage data which is absolute. Microleckoge is a series of phenomena and not a single entity. The physicol and chemical nature of the restorative moterial and oral environment and the clinical skills of the operator influence the cheracter of microleakage. 4. CHAPTER 3 FACTORS INFLUENCING MARGINAL ADAPTATION Penetration of fluids around margins of restorations will depend on the skill of the operator, manipulation techniques, solubility of the restorative material ond the material's ability to adapt or adhere to the tooth surface (Phillips et al, 1961; Swartz, 1962; Phillips, 1965 (b); Roydhouse et al, 1967; Roydhouse, 1968). Going (1972) pointed out that marginal leakage was on inherent shortcoming of the commonly used restorative materials and so stressed the use of correct clinicol techniques. Poor packing, condensation or insertion of restorative materials may lead to increased dimensional changes on setting or polymerisation, early dissolution of materials and poor marginel fit. Aging of a restoration in the oral cavity will often change the character of microleckage for the restorative material. It is accepted thot amalgam leaks initially but will ultimately become self sealing. “Temporary stopping” material, gutta percha and acrylic resin restorations become more susceptible to leckage after some time in service (Going et al, 1960 (b)}. It is important to consider the effect of dental ploque and the hydrostatic pulpal pressure on marginal leakoge (Stuever et ol, 1971). Marginal leckege is influenced by cavity preparation and the type of restorctive moteriol. This is shown in terms of recurrent cories, postoperative sensitivity ond marginal breakdown (Phillips 1965 (b)). Morginel failure is the loss of integrity between 42, restoration ond tooth. It permits fluids to seep into the tooth- restoration interspace ond is influenced by the amount of tooth structure, location of the covo-surface angle and the selected restorative material (Mohler and Terkla, 1965). Varnishes ond liners may be used in conjunction with the restorative moterial for a successful reduction in marginal leckoge (Phillips, 1965 (a), 1965 (b)). Gloss ionomer cements and polyccryletes are able to physicochemically adhere to tooth structure ond so offer hope of a truly sealing restoration. Due to other physical characteristics of these materials, their use is limited. 3.1 Cavity Preparation Poor adaptation may be related to cavity preporction, the restorative material or the quality of the remaining tooth structure (Mahler and Terkla, 1965). Irregularities of the cavity wall and the cavo-surface margin of @ cavity preparation have been examined to determine their influence on adaptation of various restorotive matericls. Menegale et ol (1960) tested the adaptation of amalgam, silicate, acrylic resin, gold foil and mat gold to rough and smooth cavity walls. Radioactive isotope (Co“>) penetration was less for rough-walled cavities then for smooth surfaced preparations. It is likely that irregularities provide better retention of the restorative material Charbeneau and Peyton (1959) and Grieve (1971) demonstrated good odoptation of onolgam to rough cavity walls with reduced leckage Grieve (1971) considered the irregularities of the surfeces involved in some way prevented the free passoge of marginel leakage, as smooth 43. walled cavities exhibited more leakage than rough walled covities. Gold inlay restorations require smooth surfaced cavity preparations to facilitate pattern toking ond seating of cast restoration, although very minor irregularities of the cavity wall are used by the cement for interlocking retention (Charbeneau and Peyton, 1959). From the standpoint of cavo-surface marginal strength, smoothness of the cavity wall is desiroble os roughness close to the cavo-surface margin may undermine groups of enomel rods producing weakened cavity morgins (Charbeneau et al, 1957; Charbeneau ond Peyton, 1959). This is a failure of the tooth structure rather then a failure of the restoration itself. Proper use of hand instruments to remove loose enomel rods improves the covo-surface margin (Khera and Chan, 1978). Finish of the restorative material is important os microleokage at the cavity margin may be influenced by the presence of fractured enamel, excess material overlapping the cavity margin or by a frank deficiency of the restorative material (Khera and Chan, 1978). Brittle moterials must be placed in cavities with cavo-surface angles that are formed parallel to the direction of the enamel rods and at 90° to the surface of the tooth (Mahler and Terkla, 1965) Investigators have varied the angulation of the cavo-surface margin and internal ongles of the cavity preparation. Azar et al (1968) using omalgan showed adaptation improved os the internal line angle approached 90°, ‘he used 30° as the smallest tested angle but felt it was not possible to make line angles of this size clinically. Amalgam poorly adapts to edge angles less than 70°, (Elderton, 1976 (b); Khera and Chen, 1978) and is prone to marginal fracture 44, under occlusal load (Forsten ond Kallio, 1976). Enomel finish i important particularly in the cervical region, as the gingival extremities of restorations are difficult to instrument, more prone to bacterial accumulation and subsequent etching (Going, 1972) The covo-surface margins of composite resin restorations may be etched to improve retention. Acceptable finish may be achieved by leaving a slight featheredge of composite flash on the cavo-surface of the tooth (Buonocore, 1975; Going, 1979). A primer may act os on adhesive sealer. Monipulative procedures will also influence adaptation Amalgam restorations are susceptible to handling differences. Alloy selection, mercury content, condensation ond finishing techniques wil vory cavity wall and margin adoptotion (Crawford, 1938; Eoston, 1941 Nadal et al, 1961 (a), 1961 (b); Nedol, 1962; Ryge, 1985; Eomes, 1967 Simon and Welk, 1970; Elderton, 1976 (b); Going, 1979). Poor edaptation of any moterial ond the presence of voids will increase marginal leckoge and the susceptibility of recurrent decay (Eames 1967). As the physical properties of restorative materials and tooth structure differ it is difficult to maintain the integrity of the tooth-restoration interface. Compensation is made for the restorative material by the cavity preparation. Covity preparation should be conservative but adequate. Increased retention minimises the differentiol movement of tooth and restoration. A weak restorative material should be expected to support less load and thereby develop less stress thon a strong restorative material (Mohler and Terkla 45. 1965; Buonocore, 1975). Varnishes and liners may also be used to reduce leakage and improve marginal seal (Going, 1972; Andrews and Hembree, 1978). 3.2 Amalgam 3.2.1 Cavity Preparation Improper cavity preparation can bring about the foilure of an omalgem restoration. Healey and Phillips (1949) attributed improper cavity preparation as the cousative factor in 56% of all foilures of omalgom restorations. Poor treatment of the cavo-surface margin, such as leaving unsupported enamel or feilure to produce cave-surface margins that will provide moximum bulk of the restorative material will result in marginal breakdown (Wing, 1971). By varying the line angles from 30° to 90° Azar et al (1968 demonstrated that amalgam adaptation improved as the line angle increased to 90°. Amolgam condensed into smaller angles contained voids and frank deficiencies (Azar et al, 1968; Khera and Chan, 1978). Jdrgensen ond Polbél (1965) and Forsten and Kallio (1976) examined the strength of the marginal angle. Jdérgensen and Palbél (1965) found the strength of the amalgam decreased as the angle became smaller and the actual dimension of the fracture increased with the “more acute angle, Forsten and Kallio’s (1976) study was more orientated to marginal fracture and olloy type. By using alloys of conventional (Ag,Sn) ond dispersion modified omolgams at a constant cavo-surface angle of 45° they showed more breakdown for the higher creep and more corrosive conventional alloy thon for the dispersion modified alleys. It is difficult to condense omelgom into sherp well defined line angles. This increased difficulty will result in voids 46. and deficiencies with increased marginal leakage (Kher and Chan, 1978). Torney and Noorian (1979) found void areas occurred more at buccogingival or linguogingival line ongles thon at any other single margin of a Class II amalgam restorction. Rounded line ongles are favoured as they facilitate good condensation. Conservative preparations of adequate extension ore permissible with the increased use of fluoride opplications and properties of modern omalgom. There is less morginel breakdown in smoller restorotions per length of margin than there is for more extensive restorations (Fusayoma, 1971) Good retention holds the restoration and prevents differential movement of the tooth structure ond the restoration. In cases of extensive tooth structure loss, pin retention may be required. Pins are placed at least 0.5 to 1.0 mm inside the dentinoenamel junction Properly placed pins provide additional retention and stability of the restoration within the cavity (Going, 1979) 3.2.2 Manipulation 3.2.2.1 Selection of alloy: The type of alloy used may be selected from conventional AgzSn or high copper alloy systems, using lathe cut or spherical particles or a combination of these. Selection may be made according to particle size, required mercury to alloy ratio or rate of set. The selection may be influenced by the dentist's preference and ability to use o perticulor alloy (Wing, 1975) 3.2.2.2 Proportioning: When mercury to alloy ratios of 8:5 and 7:5 are used, mercury is expressed both ofter trituration ond during condensation. Higher residual mercury levels cause greater formation of gamma one ond gamma two phases and result in a weaker restoration Ratios of 1:1 (Eames technique) have been suggested to maintain lower final expansion and high strengths (Schulman and Vaidyanethen 1976). This is ot variance with the findings of Wing ond Hewitt (1965) who found that amalgams with similar strengths could be produced using different initial ratios. The proportioning of alloy ond mercury can be maintained at @ constant ratio by using volumetric dispensers or encapsulated alloy systems, (Wing, 1971, 1975). Holst (1965) found a relationship between prolonged trituration and crushing strength of amalgom. By slightly increasing the trituration time coarse grained omalgom alloys produced a set mass of slightly higher crushing strength. Fine grain alloys were not affected. The importance of proportioning is that sufficient mercury i used so a plastic mix results (Wing, 1975). Granoth (1971) employed the wet technique of condensotion in preparing specimens as he considered better adaptation wos achieved with a plastic mix of high mercury content despite the dimensional, chonge factor. Mitchem and Mahler (1968) and Simon and Welk (1976) also considered adaptation to be influenced by the plasticity of the precondensed mix. Dry mixes (mixes where suboptimal initial mercury is used) are controindicated as they are difficult to condense, are usually porous, poorly adapted ‘and weak (Jérgensen, 1972). 9.2.2.8 Trituration: Triturotion is the mixing of mercury and alloy powder to allow complete alloying. The oxide present on the alloy porticles is removed ond sometimes particles are fractured during trituration. Good triturotion is necessery for a proper interface between silver-mercury ond silver-tin phases (Wing, 1971, 1975 Darvell, 1976; Baron ond O’Brien, 1977) 48. An under triturated amalgam will appear to be granular and sometimes soft and less coherent during condensotion (Eomes, 1976). These amalgams usually are weak, not smooth and have an increased exponsion (Schdlman and Vaidyonathan, 1976) Over triturated amalgams also produce an inferior restoration. The masses ore very fluid and difficult to handle. There is also © reduction in the proper setting expansion (Schulman and Voidyancthon, 1976). Ropid hardening does not permit adequate condensation (Eames, 1976). 3.2.2.4 Condensation: Condensation influences strength, mercury content of the restoration ond will hove a profound effect on the marginal adaptation and corrosion resistence (Wing, 1971). Condensation is affected by direction and magnitude of the condensing force, condenser size, increment size and number, the number of thrusts per increment and the length of time after trituration the omalgam mass is condensed (Phillips, 1953; Wing, 1962, 1965 (b); Jérgensen, 1965 (b); Wing, 1971). Eames (1976) recommends the use of serrated condensers, over ‘smooth condensers which tend to slide across the mass rather than attach to it. The serrated condenser drags the amalgam to the cavity wall providing adaptation to the lateral walls. Leterel end vertical movements are employed during condensction. Small condensers are used to adapt the initial small increments to the retention area, followed by a medium sized condenser ond lestly o larger condenser to pack the excess amalgam. Excess mercury is removed during 49. condensation to reduce residual mercury content (Wing, 1971; Eomes 1967, 1976; Schulman and Vaidyanathan, 1976). Heavy condensation force is recommended. Forces of 36-44 N should be used but 13-27 N ore more likely to be the clinical loods achievable (Nodal, 1962 Wing, 1975). Basker and Wilson (1970) demonstrated a linear increase in tronsverse strength when the packing force was increased from 10 N to 20 N. Increasing the condensation time from 1 to 2 minutes also increosed the transverse strength. This showed there is o need for adequate length of condensation time. An increase in the number of thrusts, also produced an increase in transverse strength. Bosker and Wilson (1971 (a), 1971 (b)) exomined mechanical amalgam packers. Some devices were more effective thon others. The packing force and rate produced by an operator were dependent on the packing instrument used. McHugh (1955) ond Hatt (1959) found mechanical condensers gave better adaptation than hand condensation. Better physical properties of increased tensile and crushing strengths, better corrosion resistance and reduced setting expansion ore found in well condensed amalgams. Adaptation is best’ if condenser shape is selected to fit the area to be condensed relative to the line angles and smoll increments used. Nadal et ol (1961 (c)) exomined the effect of residuel mercury contents of 47%, 58% ond 62%. Though marginal deterioration was noted in all cases the number and severity of marginal feilure increased as residual mercury content increased. Surfoce roughness and general degradation were manifested by many restorotions with average or high mercury content. An upper Limit of 55% wos placed for the residual mercury content (Phillips, 1965 (a)). Phillips (1965 (b}) considered even 54% residual mercury decreased strength 50. properties resulting in corrosion and eorly marginal breakdown Jdérgensen (1965 (c)) also found that increased mercury content reduced the erushing strength of the set omalgam and increosed the relative volume of ¥, and ¥, present in the set omalgom. Wing (1971) pointed out that although the residual mercury content of a restoration may be 50%, the marginal oreas ore mercury rich ond could contain as much as 60% if the condensation was not to 1 mm excess ond mercury removal not facilitated. Investigators claim superior adaptation using different clloys. Eomes (1967) considered the slow setting amalgams adapted better than fast setting amalgoms. Ryge (1965) and Schulman and Vaidyanathen (1976) used smaller sized particles which required more mercury produced a rapid set and were more easily adapted to cavity walls than larger particles. Mitchem ond Mohler (1968) exomined microcut finecut and spherical alloys. Microcut alloys were poorly adapted to cavity walls and showed excessive morginel mercury content. Finecut and spherical alloys were better adapted ond with less marginal mercury. Jérgensen (1965 (a)) preferred pre-omalgomated fine-grained, zinc-containing alloy for maximum adaptation Similar adaptation of amalgam restorations con be secured through the utilization of any sound manipuletive and condensetion procedure {Swartz ond Phillips, 1961]. All omalgam restorations leak initially ond require the application of a cavity varnish and possibly o base to reduce early microleakage (Swartz and Phillips 1961; Gronoth, 1971; Smith et al, 1978; Going, 1979) Contamination of the omolgom by moisture during condensation 51. should be avoided. Proper isolation will maintain a dry field. 3.2.2.5 Carving, burnishing and polishing: Carving with a sharp instrument is performed immediately after condensation (Wing, 1971) to ensure no small excesses exist beyond the prepared cavity margin ‘end to provide for proper form and function. Burnishing has long been held in disfavour because of a tendency to bring excess mercury to the surface but more recently i has been realised that light burnishing opproximetely ten minutes ofter trituration may improve the marginal integrity of the restoretion. Light burnishing results in « better surface when no polishing is employed (Kato et al, 1968; Tidmarsh ond Gavin, 1973. Eames, 1976; Jérgensen and Saito, 1976). Schulman and Vaidyanathen (1976), Chan et al (1977) and Cothren et ol (1978) recommend burnishing as it resulted in decreased marginal leakage, reduced tarnish potential, created a smoother surface, ond a decreased residual mercury content of the burnished layer. Chan et al (1977) examined burnished margins with a scanning ‘electron microscope. Smooth surfaces were produced by burnishing and on omorphous-like bulk of amalgam abutted the cavity margin, this mass might decrease leakage. Burnishing was seen to increase the relotive content of residual alloy grains oround margins and reduce the number of micropores so improving the property of amalgam margins (Kenai, 1966; Tidmarsh and Gavin, 1973). Despite the above reports Cothren et ol (1978) using Co“® demonstrated thot burnishing did not 52. lead to ony significont difference on the incidence of microleakage in either specimens with or without copal resin varnish. These results cre in conflict with Kato et al (1968) who using o dye sprayed on the occlusal surface found less leakage with burnished specimens The dye technique is not as sensitive os Ca*® ond so may not be the idecl experimental method and the use of acrylic dyes may also involidate findings os it would be more clinicolly orientated if tooth structure was used. Polishing removes the surface mercury rich layer resulting in @ more homogeneous layer which will be more resistant to tarnish and corrosion (Wing, 1965 (c), 1971). Excess amalgam is removed and the omalgam-restoration surface rendered flush. The improved marginal areas should be less susceptible to marginal fractures. Too much heat must not be generated by polishing os @ surface tempercture of 65°C will bring mercury to thot orea (Wing, 1971, 1975; Eames, 1976) Tidmarsh and Gavin (1973) question the use of polishing by plug finishing burs, pumice and an oxide paste. Finishing the ename: edge down to the restorative material in places where it is short of the cavity morgin couses ridging and grooving of the enamel surface. They recomend smoothing with a silicon diatomate paste instead of the horsher obrosives such as pumice and zirconium silicate which scour ond scratch both amalgam and enomel. 3.2.3 Corrosion, High Copper Alloys: Corrosion is associcted with the neture of omelgam, which contains several phases of different electrical potential giving rise to galvanic corrosion. It may be this inherent corrosion potential of amalgam restorations that enables 53. them to become self sealing. The introduction of high copper alloys with no gamma-two in the set mass has caused many investigators to study the effect on microleckage. High copper amalgams possess low creep, improved resistance to morginal breakdown end reduced tarnish and corrosion (Eomes, 1976; Schulman ond Veidyanathan, 1976; Torney and Nooricn, 1979). Galen et ol (1973) dynamically. loaded and deformed restorations. High copper alloys resisted fracture and maintained morginel integrity, their low creep properties were thought to be reloted to deformation and/or marginal breakdown. Nahler and Marentz (1979 (a)) showed thet superior resistence to marginal breakdown of high copper alloys was related to the alloy system Operators can influence the quclity of a restoration but morginal fracture is portly influenced by the alloy type. De Rijk (1978) was concerned thot the newer less corrosive amalgams would not be self sealing and so allow continued microleakage of the restoration and subsequent failure. Andrews and Hembree (1978) compored the microleckage of conventional and high copper restorations at 24 hours, 3 months and 6 months. All restorations exhibited moderate to gross marginal leckage at 24 hours and 3 months when cavity vornish was not used. At 6 months no or slight marginal leakage was evident. This study indicates that the corrosion- resistant high-copper amalgams allow the same degree of morginel leckage and subsequent self sealing as conventional (Ag,Sn) alloys 9g dg The gomma-two phase is the most electrochemically active and it is this phase that undergoes preferential and destructive corrosion in dental amalgams (Wing, 1975). The oxygen concentration differences of the surface of the restoration facing the cavity, and the walls in the pores of the amalgam with reletively low oxygen tension make them anodic with respect to other omalgam surfaces. Jdrgensen (1972) is of the opinion that areas of high porosity cause increased corrosion ond marginal breakdown. McCurdy et al (1974) demonstrated corrosion to occur more rapidly in vivo than in vitro. Lyell et al (1964) stored amalgam restored teeth in saliva and sulphide solutions and also found a decrease in marginal leckage. It is thought that metallic ions and corrosive products fill the tooth-restoration interface and reduce microleckage. As marginel percolation, (Nelsen et al, 1952) dye and isotope penetration (Kidd, 1976 (a)) studies demonstrate initial leckege which decreases with age, a cavity varnish should enhance the early seal of the dentine (Oilo, 1976; Going, 1979). Grieve (1973) substantioted the application of cavity varnishes as in vitro development of caries was less for cavities in which varnish was opplied. Cavity liners may block or retard the recurrent cories Process until chonnels ere filled with corrosion products and the “mechanism of fluid exchange is disrupted. 3.3 Composite Resins Kidd (1976 (b)) using on artificial caries technique demonstrated less demineralisation around composite resin restorations than for amalgom. McCurdy et al (1974); Kidd (1976 (a) 1978), Saltzberg et ol (1976) and Leinfelder et al (1980) showed re 55. composite materials tested in vivo and in vitro allowed minima microleakege. In conflict, a study by Khera and Chan (1978) reported greater leakage with composite resin restorations thon with onolgam restorations in both number and degree. The variation in the adoptation of composites is expleined by Lee and Swartz (1970) and Saltzberg et ol (1978). Fillers in the composite resin serve to reduce polymerization shrinkage and thermol dimensional changes. Fillers can provide for o closer marginal odaptation depending on the amount present and size. Composite resins may show vast differences in surfaces, shrinkage and marginal adaptation depending upon the type ond concentration of filler particles and the thermosetting resin used. Lee and Swartz (1970) correlated covity adaptation to the volumetric polymerization shrinkage. The greater the shrinkage the lerger the gop. Asmussen and Jérgensen (1972) used a microscope to study the adaptation of a voriety of “plostic filling materials” to the cavity wall. The materials studied included composite resins ond resins that contained no inorganic filler. The adaptation of the restorative material to dentine ond enomel wos examined shortly after initial set or after vorying periods of immersion in water. Immediately ofter initial set a marginal gop ot both the enamel and dentine levels was observed. Polishing the restorations soon ofter setting resulted in a zone of fractured enamel 20-40 micrometers wide. After storage in water’ the marginal gaps were reduced and for some brands the gaps were completely closed after 32 days. Polishing of the restorations with closed shrinkage gaps resulted in a minimum of fractures of the enamel margin, 5b. Etching-enanel os a means of bonding the composite to tooth structure was hoped to decrease marginal leckage. Many investigators report a reduction .in leakage cround etched composites. Kidd (1978) showed, by on artificial cories technique, no further improvement in sealing the restoration against an etched enamel wall. Merchant et al (1977) and Reeder et al (1978) acid etched dentine ond rendered it more permeable. Acid etching is potentially deleterious in that it increases the ability of fluids to move through the dentine so increasing sensitivity and the ability of bacterial products to diffuse into the pulp. Variation in the finishing and placement techniques may result in a reduction of microleakage.. Finishing to butt joints is not often ochieved clinically. Usuelly @ "featheredge” of composite overlaps the enamel surface resulting in a better clinical finish. After one year service in the mouth greater marginal breakdown is evident by staining, ditching and recurrent caries around butt joint margins. Overlopped restorations maintain a considerably better marginal Integrity although some breakdown was evident. Surprisingly there is no essenticl difference in the degree or pattern of marginal leakage of butt jointed compored with the overlapped restorations. Primer sealing of margin appears to produce @ considerably more effective “ond durable marginal seal than direct seoling. The unfilled sealant is pleced onto the etched enamel to which, in turn, the restorative material is bonded (Buonocore, 1975) The introduction of microfilled resins with their high coefficients of thermal expansion places doubt upon their sealing potential. It hos long been considered that high coefficients of 57. thermal expansion will not permit long term marginal adaptation. Microfilled resins have high water absorption potentials ofter polymerization and expand with water uptake. This expansion presses the restoration moteriol against the cavity wall. The positive force at the cavity margin overcomes the effects of a high coefficient of ‘thermal expansion (Going, 1979). 3.4 Gloss Tonomer Cements Glass ionomer cements adhere to enomel ond dentine by means of polar and ionic attractions - a physicochemical adhesion (McLean and Wilson, 1977 (0), 1977 (b)). If this bonding is of a long-term nature, then the problem of microleckage ond secondory caries around the margins of these restorations could be reduced (Council on Dentol Materials ond Devices, 1979). Short term experimentotion on the marginal integrity have been encouraging. Thermocycled Closs III and Class V restorations exhibit a relotively good seal even where morgins were wholly or partially bordered by cementum or dentine, or both. The bond to enamel is stronger than the bond to dentine (Council on Dentol Materials ond Devices, 1979). Kidd (1978) considered the seal ef glass ionomer cement to be good ond it inhibited demineralisotion along the cavity woll when subjected to on artificial cries technique. There oppeared to be a ‘cariostotic’ effect exerted by the material. Becouse attachment of the cement is vie ionic and polar bonds, the release of fluoride and its uptake by the enamel cre facilitated. The intimate molecular contact focilitates exchange of the fluoride ions with the hydroxyl ions in the apatite of the surrounding enamel (Council on Dental Materials ond Devices, 1979). Other properties of this material restrict its application. 3.5 Silicate Cement Silicate oppeors to allow microleckage between the restoration and cavity wall (Kidd, 1976 (a)). The behaviour of silicate restorations is variable, many investigators found some specimens exhibited gross penetration of the margins and less ingress of penetrant in others. The surface of silicate allows permeation by isotope (Going, 1960; Phillips et al, 1971; Swartz ond Phillips, 1961; McCurdy et ol, 1974). Although silicate cement restorations ore subject to "washing out” of the cavity preparation many report no detected change in the adaptation or on improvement in the marginal seal with time. Swartz ond Phillips (1961) in vitro ond Phillips et ol (1961) in vivo reported no chonge in adaptation after six months. Going et al (1960 (b)) comparing old ond new silicate restorotions, indicated that marginal penetration was reduced with age especially when recurrent caries did not occur. Dimensional change of the setting motericl will affect the adaptation (Going et al, 1960 (b); Lee and Swartz, 1970). The greater the shrinkage the lorger the size of the gop between tooth and restorative material. Swartz (1962) showed increased permeation by isotope in restorations that were finished inmedictely ofter insertion and so recommended a delay of finishing for 24 hours. Smooth cavity preparations will also allow increased leakage of silicate cement restorations (Menegele et al, 1960). The operotor con therefore affect the success of the restoration (Roydhouse et al, 1967). Nelsen et al (1952) and Kidd (1976 (a)) noted that the marginel secl of silicate cement restorations oppears little affected 59. by temperature cycling. The coefficient of thermal expansion of silicate cement is close to that of tooth structure, and hence most of the percolation about these restorations must be caused by thermal expansion of the fluid in the defects between tooth and restorations. Silicate cement restorations when first ploced contain a large amount of free acids which con be hormful to the pulp. The use of a liner or base will protect the pulp from the deleterious effects of the free acid (Mossler, 1965; Phillips, 1965 (a), 1965 (b)). Swartz (1962) wos able to improve the sealing properties of silicate restorations by placing @ varnish before packing the restoration and 20 avoid gross penetration of isotopes. Vornish should be renoved from the margins of o silicate cement restoration. Silicote cement contains 10-15 percent fluoride. The fluoride present in the cement reacts with the edjoining enomel ond dentine to bring obout a substantial reduction in enomel solubility Silicote cement restorations rarely exhibit recurrent caries in spite of the relatively poor marginal seal of this material resulting from shrinkage (Nelsen et ol, 1952; Going, 1960 (b); Massler, 1965 Phillips, 1965 (a), 1965 (b)) 3.6 Acrylic Resins Direct filling acrylic resins shrink os they polymerise. Measurements outside the mouth calculate volumetric shrinkage due to polymerisation to be 6-8 percent according to the brand used Shrinkage inside the mouth is less as some polymerisation occurs before insertion. Polymerisation shrinkage is minimised in the "brush technique” because the material is added in small increments. 60. This technique is superior to the bulk-pack technique (Nelsen et ol 1952; Sausen et al, 1953; Seltzer, 1955; Boyd, 1958; Phillips, 1965 (b)). The high coefficient of thermal expansion, approximately seven times that of tooth structure, could possibly explain the poor secling obility of acrylic resins. Changes in dimensions, of this magnitude could tend to enhance the possibility of penetration of injurious ogents at the marginal area (Nelsen et al, 1952; Sausen et ol, 1953 Seltzer, 1955; Massler, 1965; Phillips, 1965 (b)). Mossler (1965) reports pulpal inflommation begins approximately 21 days ofter insertion with irreversible damage becoming clinically ond histologically evident in young teeth, 40 days after plecement of acrylic resin restorations. This danage moy be due to excessive shrinkage and gross leakage. Monomer may cause dentinel tubules to become more permeable and so penetration to the pulp is focilitoted Slight shrinkage problems may be overcome as sorption of woter by the resin causes an expansion in volume of I-1.5 percent. This is usuclly insufficient to compensate for polymerisation shrinkage (Sousen et al, 1953; Seltzer, 1955). Kidd (1976 (a)) reports acrylic matericls oppear to have a good initial seal which is rapidly destroyed after thermal cycling. The experimental results of Going et al (1960 (a), 1960 (b); Phillip: (1965 (b)) ond Roydhouse et al (1947) showed that acrylic resins leck almost inmediately after placement and this leakage does not diminish with age. 61, The use of o cavity “primer” or “seal” improves adaptation (Swartz and Phillips, 1961; Massler, 1965; Phillips, 1965 (b)) and slightly better secling is evident in rough walled cavity Preparations (Nenegole et al, 1960). Contrary to most research results, the in vivo studies of Phillips et al (1961) and McCurdy et al (1974) found resin materials provided o relotively good seal. Restorations ot 6 months were comparable to restorations recently placed in the mouth. The degree of leckoge of Co“® wos very slight at all times. Most restorations showed slight penetration of isotope ot cervicol ond incisal margins to the depth of the dentinoenamel junction. In the light of the physicgl properties of the matericl the above results cannot be explained. Seltzer (1955) wos able to penetrate the margins with bacteria after thermal cycling and Going (1960 (b)) found dye and isotope penetrated through filling margins to the floor and into the underlying dentine to the pulp chamber. 3.7 Miscelloneous Restorative Materials Going et ol (1960 (b)) compared the sealing ability of a voriety of restorative mgterials, Crystal violet dye and radioactive sodium iodide were employed to test the marginal integrity. Both old ond fresh restorations were used. Figure 3.1 explains his method of assessing vorious depths of penetration by tracer solutions and Figure 3.2 indicates the depths of marginal penetration around different moteriols. All restorations show some degree of penetration by 1131, For gold foil, copper onalgan and red copper cement, the Isotope penetrated to less than half the depth of the cervical and incisal margins. Penetration was to the floor of the cavity for gold 62. inlays. Penetration into the underlying dentine noted for silver amalgam, zinc oxide-eugenol cement and “temporary stopping” material. In silicate cement restorations, dye and isotope penetration into the underlying dentine was recorded. Zinc phosphate cement ond acrylic resins allowed penetration of dye and isotope into the pulp in a manner similar to that observed of unfilled cavities. Temperature changing con affect the adaptation of a restorative moterial. Going et ol (1960 (b)) aged silicate ond amalgam restorations and found their sealing ability improved with time. The use of varnishes con reduce the early leckoge of these materials Swartz (1962) noted on increased leakage of gold inlay restorations after some time. Poorly adapted margins or excessive use of cements, reports Going (1979), predisposes cast gold restorations to marginel leokage, dissolution of the luting cement, recurrent caries and involvement of the pulp. Gold foil restorations require well-planed enomel margins, good retention, disciplined condensation at adequcte pressure ond frequency, ond proper finishing ond burnishing of all margins. Gold foil depends on close edaptation to tooth structure ond cohesion between gold particles for maximum retention, stability and seoling properties. The marginal adaptation of a restorctive material to tooth structure will depend upon the properties of that specific material. Most restorative materials exhibit some degree of marginal leakage no matter how much core is taken in the preporotion and placement of the restorative material. 63. Figure 3.1 se] Cw war Groen ae ec se | Method of assessing the various depths of penetration by tracer solutions (From Going et ol, 1960 (b)). igure 3.2 | See] ——— Depths of marginel penetration around different filling materials (From Going et al, 1960 (b)). 64. CHAPTER 4 JHE RELATIONSHIP OF CONVENTIONAL MECHANICAL AND PHYSICAL PROPERTIES TO MARGINAL ADAPTATION Loboratory studies of the physical properties of dental omalgom aim ct the development of a better alloy for clinical use. By studying the different characteristics under controlled laboratory conditions researchers have tried to relate these properties to the clinical behaviour of amalgam. Such properties os creep, flow, strength ond dimensional change are some properties which may influence the life of the restoration, assuming proper manipulation of the dental omalgam hos been performed. 4.1 Particle Size ond Shape An omalgam alloy may consist of lathe-cut particles, spherical porticles or a combination of both particle types. Lath ut particles: Lathe-cut omalgom alloy particles ore prepared by a number of procedures which render the alloy powder suitcble for clinical use. An ingot formed from the correct proportions of the melted metals is cut on a lathe to produce thin shavings which ore loter reduced in size by ball-milling. During the manufacture the alloys are subjected to two heat treatments. The first takes place before cutting on the lathe and primarily homogenises the alloy. The second heat treatment takes place after cutting on the lathe and subsequent ball-milling, and is an aging process for the finol olloy The oging process produces a more stoble alloy which will hove controllable reactive tendencies and properties which will not change with storage (Wing, 1975). The main difference between different lathe-cut alloys is in the size ond shope of the porticles (Smith et al, 1978). Lathe-cut particles have a rough and irregular surface (Wing, 1975), see Figure 4.1 ond are graded by the manufacturer according to particle size: coarse, medium, fine and microcut. Fine grained alloye are mostly preferred as they ore easily manipulated and produce good orialgom restorotions (Eames, 1976; Smith et ol, 1978). Particle size distribution is controlled by the manufacturer. Extremely large particles ore removed as they would interfere with the carving of the restoration. Very fine particles in the form of dust ond debris must be removed to obtain a moterial with controlled properties (Wing 1975; Smith et ol, 1978). Spherical alloys: Spherical particle omalgam alloys are prepared by an atomisation process (patented by Federol Mogul Division, Ann Arbor Michigan) which produces o mixture of spherical particles of different sizes (Demaree ond Toylor, 1962). The sizes are graded and particles larger than 50 pm are removed as they are unsuitable for clinical use Particles of different sizes ore mixed ond subjected to heat treatment to produce the desired properties. The composition of the spherical alloy porticles differs only slightly from the lathe-cut particles (Wing, 1975). The effect of porticle size ond shape on adaptation Varying the particle size of an amalgam alloy will affect the ‘dimensional chonge that occurs during setting. The smaller the particle size, the less is the exponsion when the same technique of Figure 4.1 Figure 4.2 DIMENSIONAL CHANGE IN MICROMETERSICN Scanning electron microgroph (x 500) of (a) lothe cut and (b) spherical particle amalgam alloy particles (From Wing, 1975) 1500 SR, Effect of the alloy particle size on the dimensional change of an omalgam. Curve R200 omalgam from alloy particles retained on a 200- mesh sieve; curve R325, particles retained on a 825-mesh sieve; curve P325, particles passed through o 325~mesh sieve. (From Jarabak, 1942) 67. monipulation is used (Jorobak, 1942). See Figure 4.2. Jendresen and Ryge (1960) exomined the effects of particle thickness of zinc and non-zine alloys. Thin particle omalgoms resulted in amalgam of high early strength, ond less expansion than for thick particle alloys. The zinc containing omalgoms exhibited more dimensional stability ond higher ultimate compressive strength than non-zine alloys. Demaree and Taylor (1962) examined the properties of dental amalgams made from spherical alloy particles. They showed particle size had a marked effect on dimensional change ‘ond the residual mercury content. The increased particle size resulted in an increased expansion and o decreased mercury content. Bosically large particles showed increased expansion, medium particles a slight expansion ond very small particles o contraction. By suitable blending of particle sizes’ it is possible to control the dimensionol change of amalgam. The important consideration is not the size of the particle in terms of volume, but rather its surface area. The surface area evailoble for amalgamation increases as the particle size decreases Lothe-cut particles which are rough ond irregular have a grecter surface area than spherical particles. The increased surface area favours more solution of the mercury into the alloy particles during trituration, with the result thot @ large initial contraction of the amalgam may occur (Ryge, 1965; Phillips, 1973). Ware (1960), Wilson and Ryge (1963) ond Schulman and Voidyanathan (1976) agreed that the smaller particle sizes require greater amounts of mercury for omalgamation at the some composition but they set rapidly with greater early strength and were more easily odapted to covity walls than 68. large particle omalgans. Many studies have investigated the adaptation quality of each clloy particle type. De Rijk (1974) optically tested groove adeptetion of dental amalgams using a He-Ne Laser. He believed adaptobility was of increased importance for pin-retcined omalgoms where penetration of the omalgam into the pin grooves is required for retention. Also the newer amalgams exhibited less dimensional change and reduced corrosion, and retention was therefore primarily dependent on the adaptation achieved during condensation. From De Rijk's (1976) results which are not conclusive, sphericol amalgam alloys produced surface detoil best and so he concluded that they adapted well to cavity walls than did conventional and dispersion modified alloys. Many investigators do not agree on the most odeptive alloy type. As size ond shape of the alloy particles, the plasticity of the precondensed mix and the condensation force may vary according to each researcher, it is not surprising that the results conflict Mitchem and Mchler (1968, 1969) exomined the adaptation of spherical, fine cut and microcut porticle alloys. At the manufacturers’ recommend mercury to alloy ratios, which produced different degrees of precondensation mix plosticity among olloys, spherical ond finecut alloys showed the same degree of non- adaptation. Microcut alloys showed a significantly higher degree of nonadaptation with a high marginal mercury content. When the precondensed mercury to alloy ratios were adjusted to produce the same plesticity, all alloys demonstroted the same degree of 89. adoptation. Consequently the marginal mercury content of microcut alloy was further increased. The appeal of microcut alloys oppeors to be in their faster setting time and smoother carving characteristics. Since microcut particle alloys have undesirable properties it is best to compare spherical alloys with fine’ cut lothe particle alloys. Spherical particle olloys are less susceptible to dperator variable than lathe-cut particle alloys. Eden ond Waterstret (1976) reported thot spherical alloy amalgams were as strong os or stronger than lathe-cut dental alloys. Basker ond Wilson’s (1971 (b)) results showed that spherical omolgam alloy restorations were stronger than lathe-cut alloy restorations only in the initial setting stages. Due to the actual shope ond blending of the sphericol particles less force is required to condense the olloy. into the cavity preparation to attain reasonable adaptation, low setting exponsion ond strength. Wing (1966 (b)) and Wing and Lyell (1966) reported good adaptation of spherical particle omalgam. Unlike mony other investigators they found the odaptation of fine lathe-cut particles was better than for spherical particles for optimum condensation but the reverse wos true for poorer condensation. Sphericol perticle elloys lock body and tend to flow ahead of the condenser point into chonnels at right ongles to the direction of condensation. Heavy condensation forces are wasted on spherical omalgoms as the particles roll away from a heavy condensing force when using @ small condenser As less force is required to pack a good spherical olloy amelgam these 70. restorations may be superior to lathe-cut cnalgom restorations where access and the maintenance of a high condensation force is difficult (Wing, 1975). Wing (1970) measured the restorotion-tooth interface gop. The best adaptation found was with a well-condensed lathe-cut alloy with a space of 5-10 micrometers. All spherical amalgams with 3 lbs condensing force demonstrated widths of 7-10 micrometers. Low condensation loads greatly influenced the space between lathe-cut omalgam and tooth structure, the mean gap was 25-30 micrometers Spherical amalgams in which the mercury was not fully removed during condensation showed gops of 10-15 micrometers. The blending of spherical particles (30-50 ym) increased the packability although it was not as good as thet of a lathe-cut alloy. The blended spherical omalgams gave better cdaptation than amalgoms prepared from uniformly sized spherical particle amalgams (Wing, 1966 (b)). The surface tension in mercury-rich spherical omalgom may cause “spheroiding" of the omalgam which could have @ greater influence on the lack of edaptation than the subsequent dimensional change of omalgoms. Wing (1971) concluded that the adoptotion of spherical particle omalgams is comparable with well condensed lathe-cut olloys, but better thon the cdoptation of poorly condensed lathe-cut olloys. A dispersed phase alloy containing lothe-cut particles of “silver-tin and spheres of the silver-copper eutectic spheres is known to be reluctant to omalgamate ond requires edditional trituration time to assume adequate omalgamation (Eames, 1976). Other high-copper amalgams may be composed of entirely sphericol particles. Laswell et ol (1980) compared the clinical behaviour of two high-copper omalgams Dispersalloy and Tytin. Loboratory test mn. indicate the physical properties of Tytin to be superior to those of Disperselloy. Clinically Dispersalloy disployed less marginal failure and appeared the better restorative amalgam. It was concluded by Laswell et al (1980) that the difference in composition, menufacture, size-ond shape of particles may affect clinicol behaviour in this case more so thon creep. The advantages of spherical particle amalgams over lathe-cut amalgams may be minimal olthough the spherical particle omelgoms do exhibit increased early strength. This increase in strength is not maintained in the fully set condition at a significant level. The advantage of the spherical particle amalgams is that properties comparable with very well manipulated lathe-cut amalgams moy be produced with much less effort in the spherical particle system. This will be of some advantage in difficult clinical situations where high condensation loads cannot be applied (Wing, 1975). Specific manipulative techniques have however been stressed for spherical omalgams using low condensation pressures There is still disagreement concerning which system is better: lothe or spherical amalgam alloys. Leidal and Dohl (1980) examined the marginal integrity of lathe-cut ond spherical particle omalgams using a replication technique. There was no difference clinically efter four years between lathe-cut end spherical alloys. Some difference wos detected microscopically. Leathe-cut alloys showed better morginel integrity thon sphericel analgam restorations. Leidel ond Dohl (1980) explained the difference in terms of particle size The observations of the cavo-surface margins indicated that the lorger particle size of the lothe-cut alloy is probobly responsible for high 72. resistance to masticatory forces in the marginal area, as creep values did not differ significontly. 4.2 Dimensional change Early investigators of dental onalgam considered it important thot the moterial should exhibit @ very slight expansion during setting. Expansions of the order of 10 pm/cm in the first 24 hours ofter mixing have been considered appropriate (Ware, 1960; Wing 1964, 1975). Because of the uncertointy of the importance of dimensional change it has been considered that omalgams exhibiting even slight contractions may be satisfactory clinically. The Australian Specification for Dental Amalgom specifies amalgam alloys to hove a dimensional change on setting within * 20 pm/em. A slight meosured contraction during setting in the laboratory is now considered acceptable (Wing, 1975). The Federation Dentaire Internationale and American Dental Association also outline a guide to dental materials ond devices. Dimensional change can be oltered by the manufacturer or during monipulation, Spherical alloys exhibit less expansion than lathe-cut elloys ond the blending of spherical particle reduces the 24 hour dimensional change (Wing, 1966 (b)). Increasing the surface area of the alloy particle results in o decrease in the setting expansion (Demaree and Taylor, 1962). Prolonged trituration will cause a reduction in expansion or a contraction (Swartz, 1962; Wolcott et al, 1963). Wing (1965 (b)) increased the condensation pressure ond was able to produce o slightly lowered expansion but in no instence wos o high condensation pressure able to produce a contraction. This elterotion in dimensional change produced by improved condensation wos 73. not clinically significont. See Figure 4.3. In evaluating the adaptability of an amalgom it is important that the omalgom does not pull away from the cavity wall as a result of the setting process, delayed expansion, thermal exponsion or contraction and plastic deformation (Jérgensen, 1965 (a)). Wing (1964 reports that a slightly expanding mass of amalgam will tend to deform the somewhat elastic tooth structure and so ensure intimate contact between the amolgam ond tooth when changes of dimension of both tooth ond restoration occur during variations in temperature High copper amalgams demonstrate acceptable odaptation and superior marginal integrity when compared with AggSn omalgoms. Researchers have claimed the reduced dimensional change, creep and corrosion properties of the high-copper systems produce better restorations (Gronath and Hakensson-Holma, 1961; De Rijk, 1976 Eomes ond MacNomare, 1976; Jordan et al, 1978; Malhotra and Asgor 1978; Sarkar, 1979). The clinical effect of contraction or expansion in amalgam can only be conjectured. Laboratory studies have been unable to detect a correlotion between dimensioncl change during setting ond the “interspace dimension between on amalgom restoration and the cavity woll. Clinicol investigations have not definitely detected differences between restorations made from dental omalgans showing @ high contraction or exponsion as determined according to the standards specifications (Wolcott et al, 1963; Vrijhoef et al, 1974). Swartz (1962) used Co*® to observe marginal leckage. The adaptation of the amalgam restoration was not appreciably altered by expansion or 74, i : | ‘ 1 og a * 4 & ' g 2 : ig, | go | 5” 1 2 0 oer art ae ER 0 a Ba TIME (HUNDRED MINUTES) Figure 4.3. Effect of variation in condensetion pressure on dimensional change. Curve 3, omalgom condensed under a pressure of 70 Kg per sq em (1000 p.s.i.). Curve 4, omolgom condensed under a pressure of 140 Kg per sq cm (2000 p.s.i.). (From Word and Scott, 1932). contraction of the alloy during setting, since the leokage patterns of grossly expanding or controcting elloys were comparable (Boyer and Torney, 1979). ilo (1976) examined the adoptation of amalgoms which: expanded 10 pm/om at 48 hours 2. had no dimensional change at 48 hours 3. contracted 10 pm/cm at 48 hours 4, contracted 20 pm/om at 48 hours Unlike most other researchers he found some correlation between marginal adaptation and the dimensional chonge during setting. There was a gradual decrease in the quality of adaptation from restoration made of expanding omalgom to those made of contracting amolgam. In this study no omalgom exhibited adoptetion to the cavity wall Jérgensen (1972) showed concern over contracting omalgams. He argued there wos an increased risk of marginal fracture as they were less well adapted to the cavity margin and reconmended expanding omolgams Boyer and Torney (1979) report leakage demonstrated by shrinking amalgams when cir pressure tested. It is well known that the early seal of amclgom is not routinely good. The dimensional chenge of omalgom in the cavity of a tooth is much less thon in the laboratory as dimensional stability is influenced by the shape of the mass ond by the retentive features of the cavity preparation (Wolcott et al, 1963). It is even possible that when the omelgom is confined and placed in apposition to a rough wall, the dimensional changes might be quite different to those of the clossic specimen used in physical property tests (Swartz, 1962) It is likely that the restoration will not exhibit a uniform dimensional chonge once set. Vrijhoef et al (1974) reported on the dimensional changes of setting amalgams. An expansion of the material resulted in a deformation of the initial geometric shape of the contacting surface. A shrinkage of dental omalgom may result in @ loss of contact between the restoration and the cavity wall. Mechanical retention of the restoration may allow the restoration to be tronsloted or rotated within the cavity. Controversy still exists over the effects of a contracting or expanding dental amalgam. Amalgams of low dimensional change are occeptable as it is a choracteristic of amalgam restorations to become self sealing with time, ond other manipulative vericbles may have more influence on odaptotion. 4.3 Marginal Percolation Thermal vorictions combined with different coefficients of thermal expansion between the restorative material and tooth structure con result in microleakege ond a loss of marginal integrity (Mchler ond Terkla, 1965). The success of omalgan os a filling material is attributed in part to the relatively small volume changes which occur after it is inserted. If the amalgam expands slightly on setting it may maintain adaptation and so compensate for the differences caused by thermal change in the mouth (Nelsen et al, 1952; Seltzer, 1955). Nelsen ot ol (1952) found marginal percolation was caused by o difference in the coefficient of thermal expansion of the tooth ond 77. restoration and by thermal expansion of the fluid occupying the crevice between the tooth and the restorction. They observed marginal percolation in newly placed omalgam restorations. There wos no evidence of marginal percolation observed in teeth with amalgam restorations that had been in ploce for some time. 4.4 Delayed Expansion All moisture contaminated, zinc-containing amalgams undergo a delayed ond considerable expansion. The expansion commencing 4-5 days after placement of the restoration may continue for months, reaching values greater than 400 micrometers per cm (0.4 percent) (Phillips, 1973). These restorations may cause much pain and discomfort to the patient ond so ore foilures (Schoonover et al, 1942; Jérgensen, 1965 (c); Wing, 1975; Eomes, 1976; Paffenbarger et al, 1979 (a)). 4.5 Mercuroscopic Exponsion Corrosion and contamination of hardened amalgam by mercury reportedly causes expansion. A hardened amalgam will expond in contect with mercury. This may occur if the restoration is repaired or if it comes in contact with some high-mercury-content-slurry. Corrosion of the gomma-two phase in the tin-mercury system, causing mercury to be released, has been shown to cause considerable expansion (Paffenbarger et al, 1979 (a)). Jdrgensen (1965 (c), 1972) referred to this as mercuroscopic expansion. During corrosion metallic mercury is set free. The mercury diffuses into the amalgam from the cavity side and causes o unilateral expansion of the wedge-shaped amalgam margin, which bends away from the supporting cavity wall, see Figure 4.4. This moy lead to ultimate fracture. 78. omalgem tooth Figure 4.4 Mercuroscopic expansion of amalgam resulting in a deflection of the omalgam restoration away from the tooth, at the cavo-surface margin (From Jérgensen, 1965 (c)). 79. 4.6 Flow and Creep Flow relates to deformation, under static load, before the matericl hos completely set. Creep is o.rheologicel property and refers to the time-dependent deformation, produced by o stress, ina completely set solid (Phillips, 1973). Creep may be o more significant property in describing deformation of the clinical restorotion, since masticatory stress is usually opplied following complete hardening of the amalgam (Phillips, 1973). The clinical significance of flow is somewhat uncertain. Some relationship hos been claimed between the property of flow and some undesirable clinical behaviour of amalgams, such as flattened contact points, overhanging margins and protrusion (Wing, 1975). The recently developed high copper alloys show less flow than Ag,Sn alloys and superior marginal integrity (Granath and Hekensson-Holma, 1961; Eames, 1976; Sarkar, 1979). Jérgensen (1965 (¢)) believed flow wos related to the deformation of amalgam under masticotory load. The displacement of amalgam could lead to slits along the filling margins of such a width thot the omalgam margins would breck under load. The mechonism is probably more complex.than suggested since flow produces shortening ond widening of laboratory specimens. Amolgams exhibiting creep values below 1 percent demonstrate better resistence to marginal fracture than amalgams with creep values obove | percent (Wing, 1979). The superior marginal integrity of high copper alloys compared with conventional alloys may be correlated to creep values (Mohler et cl, 1970; Galan et al, 1973; Eames, 1976 Osborne et cl, 1976; Derand, 1977; Jordan et al, 1978; Malhoutra and Asger, 1978; Osborne and Gale, 1979; Wing, 1979). Caution should 80. be exercised when using mechaniéal properties to predict clinical performance (Osborne and Gale, 1979). Jordon et al (1978) questioned the reliability of the property of creep in accurately predicting the marginal integrity of an omalgom alloy. The alloy with the highest creep did not demonstrate the most extensive marginal breakdown. The most extensive marginal breckdown wos observed with @ low static creep alloy. Studies by Forsten and Kallio (1978) ond Loswell et ol (1980) agreed that the mechanical properties of the alloys did not always predict the degree of marginol breakdown of the amalgam restoration. Sarkar (1979) explained the reduction of flow ond creep in high copper alloys as a result of the reduction of gonma-one through the substitution of copper for silver. The gorma-one phose has long been considered to be related to creep. The presence of the strong, hard copper-tin intermetallics dispersed in the motrix moy also reduce the Flow ond creep by providing obstacles to the movement of dislocations. Poffenbarger et ol (1979 (a), 1979 (b)) correlated creep with extrusion, ond extrusion with marginel breakdown. The greater the extrusion the worse the marginel breckdown. These workers suggested that the relationship between the degree of creep ond marginal integrity can be related to the resistence of the amalgam to deformation, 4.7 Modulus of Elasticity The quality of the seal of on amolgom depends on a number of factors such as the elosticity of the material, of the tooth and amalgam, and the adaptability of the amalgam. A restorative moterial al. having a modulus of elasticity different from that of tooth structure will deform under load a different amount then the tooth. This may result in problems such as the opening of the margin between the two materials. If the elostic limit has been reached permanent deformotion or fracture occurs. Amalgam is brittle and fractures instead of deforming (Mahler and Terkla, 1965; Granath, 1971; Phillips, 1973). 4.8 Mercury Content Adaptation of amelgam to the cavity wall con be affected by the plosticity of the trituroted mass. Sufficient mercury must be mixed with the alloy powder to render the mix sufficiently plostic to condense against the cavity woll ond into retentive crecs. Microcut amalgam particles require more mercury per volume of alloy as the evailoble surface area for reaction is increased. Finecut alloys may require 51 percent mercury to produce o workable plasticity whereas microcut alloys may require 54 percent. Microcut alloys produce restorations with « significantly higher fincl merginal residual mercury content thon finecut and spherical amalgam, and have a tendency to poorer marginal adaptation (Mitchem ond Mahler, 1968 1969; Simon and Welk, 1970; Schulman and Voidyonethon, 1976). The best adaptation of amalgam was achieved with o 50-55 percent range of initicl mercury content and this range does not eppear to alter significantly the transverse strength or the dimensional change (Eomes, 1967; Simon ond Welk, 1970). Excess mercury must be eliminated during condensation (Wolcott et al, 1963). These studies were using lothe-cut alloys. There is thought to be a relationship between final mercury 82. content and marginal integrity (Nadal et ol, 1961 (a); Phillips 1965 (b)). These works showed that when the mercury content exceeds 55 percent for lathe-cut amalgams a pronounced decrease in strength occurs. Low compressive, tensile and tronsverse strengths are reflected in acceleroted marginal fracture. Sphericel particle omalgams although they generally require less mercury per mass of alloy, require mercury removal during condensation. Spherical omalgams in which mercury has not been removed during condensation shown an increased tooth-restoration interface width. Gaps at the cavo-surface region are wider than those along the walls of the covity. The reason for this difference is Probably due to the mercury rich amalgam ond the unrestricted expansion of the amalgam thot may take place in the cavo-surface region. Surface tension in mercury rich spherical omalgams causes “spheroiding” of the amalgom and hos a greater influence on the lack of adaptotion thon the subsequent expansion of the amalgam (Wing and Lyell, 1966; Wing, 1970). Reducing the mercury content to o sub optimal concentration will reduce the plosticity of the mix. This reduction in mercury may reduce the dimensional change property of the alloy but it also increases roughness of the amalgam and so reduces adaptation (Jérgensen, 1965 (a); Abdel-Azim, 1977). 4.9 Porosity Porosity of the omalgom margin plays a decisive role in the rate of fracture. A high degree of porosity increases both the rate of corrosion and mercuroscopic deflection of the margin. Increasing 83. the poresity reduces at the some time the margins maximum flexibility. Porosity in particular seems to have a considerable influence on the crushing strength; while variation in the mercury content is less significant in this respect (Jérgensen et al, 1966; Jérgensen, 1972). Burnishing of the amelgom margin expresses both excess mercury and porosities ond improves marginal adaptation (Koto et al, 1968). 4.10 Marginal Fracture Jérgensen (1965 (c)) put forward the theory that marginal frocture was caused by @ "pressure or pull” acting on the onolgam margin. As the amalgam margin ongle became more acute the force necessary to fracture the omelgem wos reduced. A narrow space must however, exist between the cavity wall and the restoration. This spoce may be caused by the corrosion of the omalgom interface Corrosion appeared in the ganma-two phase and is followed by mercuroscopic expansion ond a deflection of the amalgam margin (Brown et al, 1962; Jérgensen and Polbdl; 1965; Jérgensen, 1972 Dérand, 1977). Mahler et al (1973) and Mahler and Marantz (1979 (b) correlated the frequency of marginal fracture with creep of the alloy selected and time. Larson et al (1979) found the high copper omalgoms which were gonma-two free maintained o greater morginal integrity thon conventional alleys. Bryant (1981 (a), 1981 (b)) reviewed the marginal fracture of omalgom restorations and considered a complex sequence of events and a variety of factors influence marginel frocture. a4. Dérand (1977) investigated the stresses in a Class II restoration ond calculated the rate of deformation of the omolgom margin. The magnitude of the deformation can be correloted with the stresses, time and the viscoelastic properties of dental omalgoms ‘As the size of the restoration increased cuspal deflection and stresses increased under load. When o Class II restoration is newly placed and load applied, the creep in the omalgan results in o narrow defect at the margins. Corrosion may occur even in the absence of such a defect. When the contact between amalgam and cavity walls is lost, the stresses ‘and creep in the amalgom are much lower and are of less importance. From thot point, acceleration of the marginal fracture probobly depends fon corrosion and breakdown of amalgam. Dérend (1977) concluded that the correlation between marginal failures ond creep valves is significant in the short-term but in the long-term resistance to corrosion is a more important determining factor. 85. CHAPTER 5 THE NATURE OF THE INTERFACE BETWEEN TOOTH AND AMALGAM 5.1 Corrosion Dental amalgam lacks homogeneity to ensure corrosion resistance. The different phases present in the hardened omalgom exhibit different electrode potentials and produce a corrosion cell with solive as the electrolyte (Phillips; 1973). The gomma-two phase is the most electrochemically active ond it is this phase that undergoes preferential ond destructive corrosion in dental amalgam (Wing, 1975). The golvanic corrosion attacking the gamma-two phase dissolves the tin ond releases metallic mercury into the remaining amalgam. The corrosion feilure of amalgams may occur through the dissolution of the passive oxides (stonnous ond/or stannic oxide) in gorma-two with the formation of tin oxychloride (Sarkar and Greener 1975 (a)). Much of the tin is deposited locally either within the corroded amalgam or on the amalgom surface (Jérgensen, 1972; Phillips 1973). The depth of corrosion of the ganma-two phase in vitro occurs as o function of the amount of the phase ot the surface and is restricted to o depth of approximately 200 pm from surface in well condensed amalgams (Stevenson, 1973; Stevenson and Wing, 1975; Wing 1975). Jérgensen (1972) believes corrosion to be able to progress further into the body of the amalgam. His specimens oppecr to be poorly condensed, containing much ganma-two and porosity. Porosity exposes a greater area of the goma-two phase, increasing the crea of the anode and increasing the susceptibility of the mass to corrosion 86. (Jérgensen, 1967; Jérgensen, 1972). The corrosion of gamma-two reduces the life of the omolgam restorotion by: 1 causing surface roughness and pitting due to galvanic action as a considerable electromotive force is generated between the gamma-one and ganma-two phases (Duperon et al, 1971) marginal breakdown caused by concentration cell currents occurring along the margins. The microleckoge that occurs between the restoration and the tooth provides an electrolyte clong the covity walls thot is different from the electrolyte on the surface of the restoration. The amalgam surface facing the cavity wall acts os an anode and the external surface as the cathode in o concentration cell. Although corrosion products that form, aid in sealing the restoration the reaction may proceed clong porosities and microcracks deep into the restoration and may lead to marginal breakdown (Massler and Barber, 1953; Phillips, 1965 (b), 1973; Duperon et cl, 1971; Brown, 1976). Lyelle et al (1964) immersed conventional amalgam restorations in sodium sulphide and saliva solutions. They demonstrated a marked reduction in marginal penetration by dye and isotope as metallic ions ond corrosive products filled the tooth restoration interface discoloretion which depends on the condition of the cavity when the restoration wos placed. Kurosaki and Fuseyene (1973) found tin and zine penetrated softened dentine remoining after cavity preparation. The tin penetrating the dentine is considered to have been dissolved from the gammo-two phase because of its high ionization tendency. Zinc moy penetrate 87. softened dentine also as it is the most soluble of the omalgom components with the highest ionization tendency. The mercury from the gamma-two phose could penetrate back into the amalgam ond react with the unreacted alloy cores (AggSn). Tin and zine appear to precipitate in the some way and take the place of calcium. If dentine is discoloured it is probably due to sulphide formation with tin, as tin produces a black sulphide. Sarkar ond Greener (1975 (a)) found the corrosion behaviour of CugSn, wos similar to that of ganma-two ond its presence impairs the corrosion resistance of dental amalgom. 5.2 High Copper Alloys The reduction or elimination of the gonma-two phase in high- copper omalgom alloys has rendered them more resistant to corrosion thon conventional alloys (Boyer ond Torney, 1979). Marshall et al (1976) were able to show high copper olloys were more resistant to soline corrosion than conventional clloys. This caused some concern os it questioned the obility of high copper olloys to be self-sealing. Boyer and Torney (1979) and Andrews ond Hembree (1980) showed the increased resistence to corrosion of high-copper amalgams did not “alter the leakage patterns. The leakage around amalgams of high copper alloys is reduced more slowly than that around amalgams of traditionel alloy. By the end of two years restorations of both alloy types had virtually ceased leaking. The conditions in the crevice between the restoration ond tooth are more conductive to corrosion than at the surface and allow high copper omalgoms to corrode sufficiently in the marginel oreas to reduce leakage 88. ‘Sorkar and Greener (1975 (b)) investigated the improved corrosion resistance of Dispersalloy and attributed it to the formation of Cu,Sn os o function of time, due to o solid state reaction between the silver-copper eutectic and gamma-two. The corrosion resistance of Dispersalloy improved with aging Investigators have identified another corrosion factor, the Cu,Sng phose, but it appears to be a problem of little significance for the high-copper amalgoms (Eomes ond MacNarmara, 1976) 5.3 Investigations of the Interface The decreased leakage of amalgam restorations has been attributed to the formation of corrosion products, principally sulphides, in the tooth restoration interface. Hood and Challis (1971) onalysed the interspace debris of extrocted teeth containing restorations for silver, tin, mercury, copper, zinc, sulphur, chloride, phosphate and oxygen, by the use of an electron probe microanalyser. Sulphides, oxides, silver, mercury or copper were not detected in the interspace, but large amounts of tin were present Aggregations of copper and zinc were evident within the restorations ond diffusion of tin from the body of the restoration to the periphery was discerned. , This occurred ot o depth greater thon thot of surface corrosion and Hood and Challis (1971) reported the tin- mercury phase wos likely to be unstable under oral conditions. Wing (1974 (b)) using the electron probe microonalyser examined the amalgam-tooth interface of restorations that had been in service in the mouth for five years. Analysis of the interface indicate the presence of tin, calcium and phosphorous. Wing (1974 (b)) stated it was possible thot the formation of a colcium-phosphorous-tin complex 89. may be responsible for the self sealing property expected of conventional silver-tin omalgams. The direction of migration of tin, colcium and phosphorous to form the Sn-Ca-P complexes was examined by Chapple and Wing (1975). Extracted teoth with amalgam restorctions in situ revealed areas of the Sn-Ca-P complex within the omalgam which indicated the direction of migration had been from tooth to the amalgam. Calcium readily migrated to the tin-mercury areas of dental amalgam. No migration of calcium into the phases containing silver occurred. Chapple and Wing (1975) concluded that in amalgams containing a high proportion of the tin-mercury phase, migration of the calcium and phosphorous may occur into areas throughout the body of the amalgom. Stevenson (1973) and Stevenson and Wing (1975) determined corrosion with the removal of tin-mercury phase occurred as a surface related phenomenon only (Wing, 1974 (a)). For omalgom restorations in vivo the Ca-P-Sn complex formed as early as nine months. In vitro storage of omalgom restorations in saline confirmed that tooth structure was the source of calcium and phosphorous in the formation of the complex (Chapple and Wing, 1976). Marshall et ol (1980) exomined conventional, copper admix and dispersion modified omalgoms for corrosion products ond compositional changes ot the interface. Generally the conventional omalgoms hed lerge omounts of what they believed to be stannous-hydroxy chloride corrosion products throughout the interior of the restoration and were extremely porous. Restorations of copper rich amalgom had corrosion products confined to the surface and a few isoloted oreas. 90. (1) Changes_in conventionel amalgam The conventional amalgam examined by Marsholl et ol (1980) had been in use in the mouth for 10 years ond contained many porous creas. Most of the pores were filled with tin-chloride. Corrosion-induced changes were present throughout the entire restoration. The corrosion process. involving the gamno-two phase resulted in the formation of possibly tin-oxygen- chloride (stonnous-hydroxy-chloride) or tin-oxygen (stannous oxide) products. (ii) Changes ot the exposed surface of copper odmix omalgons These omalgoms contoin little gamno-two. A six-year-old admix amalgam restoration was examined. These omalgoms hove at least four phases: Cug Sng, ve AggSn, %)- Ag Hg ond some vs - Sn Hg. Tin rich deposits were seen ot the exposed surface and small regions of material rich in tin-chloride were detected within 50 pm of the surface. A morphological appearance consistent with small areas the garmo-two phase wos noted of ares rich in tin-chloride. Areas rich in tin were consistent with the appearance of Cu,Sn, areas. It is thought thot copper is preferentiolly removed by the corrosion process ond is released from the restoration. The only change after six years appeared to be restricted to the surface of the omolgan with the presence of deposits rich in tin at the surface exposed to oral fluids “(Marshall et ol, 1980). (444) Chonges in dispersion modified omalgoms: Generally the restoration oppecred unaffected after three years cliniccl use, but severcl areas containing tin and chloride were evident. In creas near the surface tin-rich and stannous-hydroxy-chloride deposits were found around the reaction zone of the residual silver-copper eutectic a. porticle. The Cu,Snz reaction zone wos chemicelly active (Marsholl et al, 1980). (iv) Changes ot the tooth-restoration interface: In all amalgam types examined the interface layer was rich in tin with phosphorous and calcium present. In areas 20 to 30 pm from the margin exposed to the oral cavity sulphur was also detected. Conventional ond high-copper amalgoms undergo similar reactions ot the tooth-restoration interface forming 6 complex layer that is rich in tin and may produce a marginal secl ORIGINAL INVESTIGATION 92. CHAPTER I SCOPE OF THE INVESTIGATION Many investigators have used o number of methods to determine the cdeptation of amalgam to the covity wall, It has been shown by © number of investigators that amalgam, with time, forms seal at the restoration-tooth interface ond marginal leakage is diminished. The adeptation of the recently placed omalgam restoration is sensitive to mony variables such os, condensation technique, mercury to alloy ratio, type of alloy and the skill ond technique of the operator. Tooth morphology and cavity preparation design may also influence the adaptation of the amalgam to tooth structure. Most exomination of the adaptation of amalgam to tooth structure is performed under in vitro conditions. It is not presently possible to observe continually the adaptation of amalgam to the internal cavity wall. In vive exeminetions appear to be limited to the integrity of the cavo-surface margin by the use of clinical observation, photographic recording or replication techniques The placement of omclgom restorations into extracted teeth and their subsequent sectioning and polishing permit thorough examination of vertical and horizontol walls including the cove-surface margin and retentive regions. The use of microscopy allows for the estimation of the interface gap by direct measurement ond o comparison of adaptation of different omalgams prepared using a number o| vorying techniques. 93. Strict guidelines hove been laid down concerning the monufacture and performance of amalgam alloys. All omalgam alloys should fall into these guidelines. Ideolly most types of dental amalgam alloys availeble to the dental practitioner should be capable of producing « satisfactory onalgom restoration. The present investigation aimed to study the relative edaptation of amalgams prepered from o, series of lothe-cut, spherical ond mixed particle alloys of AggSn in composition, as well as dispersion modified and high-copper ternary types. The omalgams studied were placed in naturel teeth and porcelain denture teeth. The variables studied were mercury to alloy ratio, condensation, finishing techniques and type of cavity preparation. Also studied was the effect of different operators using one amalgam alloy at a fixed initiel mercury to alloy ratio Sectioning restored notural teeth ond examination with either light or scanning electron microscopes may cause distortion of the natural tooth structure. Porcelain, being relotively inert is not as susceptible to changes in air pressure and moisture content of the environment as would notural tooth structure. Porcelain teeth containing amalgam restorations could act as good controls in informing the operator if distortion of the notural tooth structure effects the opparent adaptation of amalgam to the internal cavity walls. Sectioning and microscopic exomination will inform the operator of the marginal space quontitatively and ollow for qualitative ronking or comparison of different restorations. The changing of one variable could inform the operator as to which procedures would be most critical for any amalgam alloy ‘to produce a restoration of reasonable standard. Light microscopy techniques ore not os domaging to natural tooth specimens as the processes of scanning electron microscopy Specimens were firstly examined optically and then exomined with o scanning electron microscope, to determine the interspace accurately and compare the measurements with those achieved with a light microscope. The interface of a recently placed amalgam restoration was compared with the interface of on amalgam restoration that had been stored for several months. This examination permitted the estimation ‘of any change in sealing thot may have occurred with time. 95. CHAPTER II EXPERIMENTAL METHOD II-1 Selection of Alloys The amalgam alloys employed in this study were selected to allow the exomination of the behaviour of a wide range of alloy types. The alloys used are shown in Table II.1. The alloys selected represented a variety of alloys of the ‘Ag,Sn” formulation as well as the high-copper alloys. The “Ag, Sn” group included lathe-cut alloys and spherical particle alloys. The high-copper group included dispersion modified and high-copper ternary amalgam alloys. II-2 Selection of Teeth Closs I and Class I! amalgam restorations were placed in natural and porcelain teeth. The natural teeth selected for preperation were confined to posterior teeth which were free of caries, fractures or ony other defects, having been extracted primarily for orthodontic reasons. Porcelain teeth were used as a control system to monitor dimensional changes of the natural teeth occurring during the experimental period. All teeth prior to and after cavity preparation were stored in woter. 96. Table II.1 Selection of alloys Alloy Nonufocturer Composition Particle Distribution Amaleap "F* | Degusso, Germeny [AggSn (+ CugSn) | Lathe 100% * In Free (Fine) F400 Southern Dental [AggSn (+ CugSn) | Lethe 100% Industries, (nicrofine) Australia. Ultreceps Southern Dental [AggSa [+ CugSa) | Lethe 100% Industries, Austrolia 6.C."2 Hi 6.C. Chemical Co. |AggSn Spherical 100% Atomic Japon pmalcop-non- | Degussa, Germany |Lathe: Ag,Sn | Lathe 70% gonma-2 * (+ Cugsn) Spherical: Ag-Cu | Spherical 30% eutectic + ispersalloy | Johnson & Lethe: gn | Lathe 70% Johnson, U.S.A | (+ CugSn} Spherical: Ag-Cu | Sphericol 30% eutectic + lrytin 5.8. White, Single melt - | Sphericol 100% ULsA. high-copper ternary: Ag 60%, Sn 27%, Co 19% Jbeheriphase © | Southern Dentel [Lothe: Aggin | Lathe 90% Industries, (+ Cug5n) Australia $ Single melt - | Spherical 70% high-copper ternary: Ag 60% Sa 27%, Cu 13% + Composition of eutectic Encapsulated by Vivadent, Liechtenstein. Ag 71.9%, Cu 28.1% 11.3 97. Covity Preperation Stonderdizotion of all cavity preparations was attempted but onatomical form of the tooth for specimen preparation resulted in a voriation of cavity preparation form. (i) (ii) (iit) (iv) Differences in cavity preparation resulted due to: variation in the depth of enanel.on the occlusal surface. Some natural teeth had very deep fissures which terminated deeply in a pulpol direction. Other teeth had shallow, non retentive fissures. A range of enamel depths on the occlusal surface was observed, see Figure II.1 cuspal incline. The angulation of the cuspal incline with respect to the long oxis of the tooth varied greatly. An attempt to finish cavity margins for amalgam restorations at right angles to the external tooth surface is rarely possible, to attempt this in some cases would have resulted in undermining cusps. The covo-surface angle of Class restorations in teeth with steep cuspal inclines was more ocute then for those of teeth with flatter cuspal inclines See Figure 11.2; tooth type. Mandibular and moxillery teeth were selected for specimen preparation. The differences in anatomicol form influenced the chorocter of the cavity preparation see Figure 11.3, This is of greater importence for natural teeth than for porcelain teeth; size of the pulp chamber. The natural teeth were generally extracted from young patients and the pulp chamber wos very lorge. A large pulp chamber will not permit significant enlargement of the cavity preparation in a pulpal or bucco- 93. “seus pour yodsno so330TJ YFM es0y3 Udy, (sUTdJOW @00}4NS-CADD 94} WOL}) SUOT}OUDdOUd Ay TAD uadsap estnbou fow SeuTTOUT ToOdsn> ds0qs yt yzee, “soustouT TodsnD Jo uostupdwos y | *]] eunBry seur[our [odsno dees re +g efBuo ‘sourpour Todsns 4033073 yam posoduos "y eqBup ‘seuspouy qodsno dees soy sop [OWS oq [THM UTesOW e20jJAS-0AD9 043 SeyoDoUddD Apoq WOBTOUD ey} YOTYM 3D eTBUD oY, UIT esn6ry 100. suotgoundesd 447009 eouenTsuT Aow dtnd 043 yo wos TOoTWOZOUD oy) yopowssd seddp ell e4nBry uppowesd smo] lol. lingual direction. Porcelain teeth covity preparations were shallower than the preparations in natural teeth as the crown size was generolly smaller and the diatoric chamber limited depth. The preparations in porcelain teeth tended to be more uniform, and cuspal inclines were more stenderdized 1I-3.1 Types of cavity preparations (1) Notural teeth: Cavity preparation was performed using a flat fissure diomond bur with o rounded tip (size 010) in o high speed handpiece. A water spray prevented drying out of the tooth during cutting. Most of the cavity preparation was accomplished with high speed. Some slow speed rotary instruments (Flat fissure No. 2, No. 3) were gently drawn across the internal cavity walls to smooth and stondordise the preparation. The approximate depth of each cavity preparation was 2.5 - 3.0 mm from the height of the marginal ridge The bucco-lingual dimension of the cavity preporation was approximately 1.5 mm, The mesio-distal extension of the Class I preparation was left 1 mm from the marginal ridge. Unsupported enamel was removed with hand instruments if required. There were two types of Class I cavity preparations. See Figure I1.3. Cavity preparation A required the further instrumentation of slow speed inverted cone No. 1 steel bur to render the line angles on the buccal and lingual ospects.of the pulpal wall sharp and to aid retention of the restoration. In preparation the buccal and lingual walls of the cavity 102. Type A cavity preparation Type B cavity preparation Figure 11.3 Cavity preparation design included Type A with sharp retention placed in the pulpal aspect of the vertical walls, and Tyoe 8 with diverging walls and rounded line angles preparation B were sloped to provide retention, the line ongles adjoining the pulpal woll were rounded Class II cavity preparations were prepared with o high speed fine-grit diomond, flat fissure bur with a rounded tip (size 010) on a high speed hondpiece with water spray coolant. Slow speed flat fissure steel burs refined the cavity preparation and hond instruments were used to remove unsupported enamel. See Figure II.4. (ii) Porcelain teeth: Class I and Class II cavity preparations were prepared in porcelain teeth. It wos necessary to prepare the teeth using a high speed flat fissure dicmond bur (fine grit) with water coolant for most of the cavity preperation. Slow speed flat fissure diamond burs were used to smooth and standardize both types of covity preparation. Stondardizetion of cavity preparation for porcelain teeth was more easily accomplished as cusp height and cuspal inclines were similar ond the diatoric chamber restricted the occlusopulpal height of the verticel walls. The Class I cavity preparation wos opproximately 2.0 mn deep and 1.5 mm wide in the bucco-lingua direction. As porcelain specimens were primarily prepared as controls it seemed of little importance to prepare both types of Class I cavity preparations. Only the Class I cavity preparation with rounded line engles on the pulpel wall were prepared. 104, enomel Figure I1.4 Class IT cavity preparation. 105. 1I-3.2 Cleansing ond mounting of specimens prior to condensation. All specimens were stored in water after cavity preparation Pe and prior to condensation. Mounting of naturel ond porcelain teeth wos necessary to stabilise each specimen to enable condensation. Special mounting blocks of Velmix” die stone were prepared. Soft red wax was placed in the mounting cavity to stabilise the tooth during condensation. See Figure I1.5. Condensation of Class I restorations was easily accomplished with this method. Condensation of Class II restorations, particularly in porcelain teeth required stabilization of the toffelmire matrix bond, which was used during condensation. Stabilization of the matrix bond and retainer was achieved using green stick compound. Once mounted in the condensation blocks the cavity preparations were washed with a water and cir jet ond dried. Excessive use of air for drying the cavity preparction was avoided. Three percent hydrogen peroxide on a cotton pellet was opplied to the walls of the cavity preparation for five seconds, the tooth debris ond hydrogen peroxide were then washed away with water. The cavity preparation was dried ond was then ready for condensation of amalgam. No copa! resin varnish wos used. Dyring mounting ond condensation each specimen was removed from water storoge for approximately 12-15 minutes. * Velmix Kerr, U.S.A. 108. |_— supporting wax {| Velmix condensation block —— Velmix condensation block supporting wax Porcelain tooth Natural tooth Figure II.5 Stabilizction of specimens prior to condensation. 107. 11-4 Trituration A Silomat* ultra-speed amalgamator was used to achieve optimal trituration in the shortest possible time Trituretion was carried out according to the manufacturer's specification. Toble 11.2 lists the trituration times for each alloy. II-5 Condensation Procedures Variations in condensation procedures from ideal to unsatisfactory methods were employed to consider the effects on marginal adaptation. Differences in condensation force, increment size, packing and finishing procedures were exomined for all alloy types. (i) Force. A heavy force of 27-36 N was compared with a lighter force of 9-13 N. Condenser size was kept constant, a 1 mm diometer serrated packer was used. (11) Increment size. The effect of o large increment size on adaptation was examined with respect to adaptation of a smaller size increment. A large increment was a full load of amalgom from the carrier which was a cylinder with diameter opproximately 1.5 mm and length 3.00 mm, A small increment was @ cylinder with diometer 1.5 and length 0.75 mm. (411) Packing. For some specimens amalgam was condensed 1 mm past the covo-surface margin and the excess was carved back to “correct” final contour. Other specimens contained amalgam which wos condensed as close to the cavo-surface margin (or as Vivadent, Schaon-Liechtenstein 108. Table I1.2 Trituration time for omalgom alloys. Alloy Trituration time using Silomat Amaleap F 5 seconds F400 8 seconds Ultrocaps 8 seconds G.C.'s Hi Atomic 5 seconds Amaleop-non-gamma~2 5 seconds Dispersalloy (pellets) 10 seconds Tytin 5 seconds Spheriphase C 8 seconds 109. close to the final contour) os possible leaving little or no carvable excess. The technique of packing flushly is sometimes used by dentists who, using © low initiol mercury to alloy ratio, assume that this type of omelgom preparation requires no mercury removal and therefore no packing to excess ‘ond subsequent carving. II-6 Finishing Techniques The different finishing techniques of carving, corving and burnishing and carving and polishing were studied and compared Specimens were carved to finel contour after condensation using a Wall’s No. 3 carver to remove the excess (if any) and a + Hollenback carver to restore anatomy. Carving was completed about six minutes after commencement of trituration. Specimens which required burnishing, were burnished 10-12 minutes after commencement of trituration. Gentle burnishing with o ball burnisher for 30 to 60 seconds smoothed the carved surface and gove a lustre to the final surface Polishing the carved surface was carried out 8 hours after “condensation. Plug finishing burs smoothed the surface, removed excess amalgam and created flush amolgam-tooth margins. Pumice and glycerine on a natural bristle brush finished the amalgam to o mott surface and a high lustre polish was ottoined with @ zine oxide and alcohol paste on a noturcl bristle brush. ALL specimens hed the lingual half of the occlusal surface 110. either polished or burnished. The other half of the amalgam occlusal surfoce was left “as corved” to observe the effect that additional treatment of the restoration surface would have on the final restoration. See Figure 11.6. II-7 Variction in the Mercury to Alloy Ratio The effect of mercury to alloy ratio on adaptation was investigated for a spherical (G.C.’s Hi Atomic) ond a lathe-cut (Amalcap F) conventional Ag,Sn emalgom alloys. The ratio differences were also subjected to different manipulator variables of heavy and light condensation force. The rotics for Analcop F included the ratio selected by the manufacturer (normal), a ratio where mercury wos less thon suggested (dry) and a ratio where more mercury than suggested by the monufecturer was mixed with the alloy ponder (wet). The selected rotios are: (4) normal mercury: alloy —5.5:5 by mass (ii) dry” mercury: alloy 4:5 by moss wet mercury: alloy 8:5 by mass Two ratios were selected for G.C.'s Hi Atomic (i) dry mercury:alloy 3.7:5 by mass, (ii) wet mercury:alloy 5.4:5 by mass. Adaptation of the two different types of G.C.‘s Hi Atomic mixes could then be compared with the other sphericol alloy system (Tytin). See Toble 11.3 for mercury to alloy ratios. We carved margin Figure 11.6 ) A, burnished or polished margin Lingual (palatal) ‘omalgem enamel Types of marginal finishing performed. The buccal cavo- surface margin was carved, the lingel cavo-surface margin was polished or burnished. 12. 11-8 Operator Variables A series of specimens wos packed by five different operators to attempt to assess differences due to operator variables. To standardize this procedure Closs I omalgam covity preparations with sharp line angles were prepared in upper and lower premolar teeth Each operator was given a selection of packing and carving instruments and allowed to pack each restoration as he would in a clinical situation. The alloy type used was Amalcap F at the manufecturer's suggested mercury to alloy retio. The premolar teeth to be restored were mounted in o phontom head in both the maxillary and mondibular arch. This permitted comparison of upper and lower restorations to determine what effect location may have on restorative procedures. Each operator contoured his packed restoration as he would normally in a clinical situation. Some operators performed burnishing after carving. The anatomical form of the occlusal surface may affect the edaptation of the restorative material to the covo-surface margin. The cuspal inclines of mandibular premolars tend to be flotter than the cuspal inclines of maxillary premolars. The larger the cavo-surface angle the more difficult it would be to “finish to the correct contour. See Figure II.7. Both male ond female operators condensed the analgom restorations. No polishing of these specimens was performed. 113. Toble 11.3: Disploys the mercury to alloy ratios of the amalgam . elloys used in this study. Alloy Mercury : Alloy Ratio by Mass Amalcop F 5.5: 5* 4:5 (dry) 8:5 (wet) G.C.'s Hi Atomic 3.7: 5** 5.4: 5 (wet) Tytin 3.75: 5° F400 5.6: 5* Ultracaps 535° ‘Amolcop-non-gamma-2 Dispersal loy 5.5: 5 Spheriphase C. + Rotios supplied by manufacturer in capsule. + Ratios dispensed from manufacturer's dispenser. +e Supplied in pellet form. W4. Steep cuspal incline enome! dentine Angle A is Jorge. It is assumed to be difficult to carve an omalgom restoration that is properly contoured to this margin. Flatter cuspal incline enamel dentine As angle A approaches 90° amalgam restorotions’ are more easily carved to the correct contour. Figure I1.7 Difference observed in finishing to the covity preparation margin due to cavo-surface angle, angle -A. Ms. 11-9 Specimen Preparation (1) Sectioning. After condensotion of amalgam into the cavity preparation specimens were stored in water, At no time were specimens permitted to dry out. Specimens were left for at least one week before sectioning To section both natural and porcelain amalgam containing specimens a diomond bur in a high speed handpiece with water coolant was used. Class I specimens were trimmed to the midline in a bucco- lingual direction. See Figure 11.8, This meant that half of the specimen was lost. Class II specimens were sectioned in a mesiodistal direction using the some technique for Class I specimens. See Figure 11.9. Before sectioning it wos necessary to coat the restored tooth in self curing acrylic resin so that during sectioning the amalgam restoration would not be dislodged from the cavity preperation. (11) Mounting. Each specimen was embedded into a perspex mounting dise with a clear self curing acrylic resin - Stellon*. See Figure 11.10, The mounted specimen was again stored in water for ot least one week before polishing * Stellon - AD Internetional, England. 116. UoTIOUTWOXe J0J SUOTzOs03Se4 I S80]9 Guturoqucs suewroeds yo GuTuoTz0ag — g*]] eunBTy uoygouTWoxe soy uoq329s TonBuyTooong (TT) MeTA [OsnT299 (1. qonéury, wo8ouo pods yBry uatm peonpes ‘jeusve Tensury Tes0ng re20ng re 118. perspex mounting block mounted specimen self cure ocrylic resin for mounting specimen Figure I1.10 Mounted specimen 119. II-10 Microstructural Preparation Specimens were prepared for metallographic examination using the method described by Wing (1961, 1963, 1965 (a), 1966 (a)). The mounted specimens were abraded to a flat surface by using successively finer grits of silicon carbide poper (220, 240 and 600). Specimens were abraded under a stream of running water at 20°C to prevent heating of the specimen during polishing. Following silicon carbide abrasion the specimens were polished on a rotating, motor driven pad*, using a selection of diamond pastes (15, 6 and 1 micrometers) with a mixture of 50% propylene glycol and water as o lubricant. Some specimens were etched using a two stage cyanide etch (Wing, 1961, 1965 (a); Wing and Ryge, 1965 (a), 1965 (b)). In one beaker, two grams of potassium cyanide (KCN), one gram of iodine crystals ond 25 mls of water were stirred to form the first solution. In onother beoker the second solution contained 2 grams of potassium ferricyonide (KFeCN), 2 mls ommonia (NH,0H) and 40 mls of water which were stirred into solution. Polished specimens were etched by first wiping the specimen for 5 to 8 seconds with cotton wool socked in the first solution until the surface become frosty. The specimen wos then washed in water ond inmersed in the second solution for 4 to 5 seconds. The specimen wos then washed in water and swabbed well in olcohol, then dried with compressed air. Buehler Ltd., Metallurgical polishing wheel, 1425/950 r.p.m. Evanston, Illinois, U.S.A. 60204 120. The etched specimens were then viewed on a metallographic microscope (Neophot I1)* ot magnifications of 50 X and 500 X to ‘examine marginal adaptation of omalgam to cavity walls, covo-surface engles and retentive undercuts. The effects of varying olloy type and condensation procedure wos examined with respect to adaptotion ond porosity in the set omalgam. Differences due to finishing techniques were also studied. After optical exomination the specimens were ultrasonically cleaned, gold coated to a thickness of 20A° and examined with o Scanning Electron Microscope** at magnifications of 12.5 X, 50 X and 250 X. This allowed for further examination of specimens and permitted comparison of results. Limited exomination of the tooth-amalgam interface wos performed with a electron probe microanalyser!** II-11 Photography The optical microscope observations were photographically recorded using a 4" x 5" negative film format and printed on black and white photographic paper at magnifications of 50 X. Corl Zeiss, Jena, East Germany: ** TS.1. Super IIIA, Jopon. *** Siemens, Etek Autoprobe, West Germany. 121. 11-12 Dimensional Change A dimensional chenge test was performed to determine the expension or contraction of different alloy types and systems, and with variations of the mercury to alloy ratio, This test wos performed under controlled laboratory conditions at 37°C. The specimens used for dimensional change testing were 8 mm x 4mm cylindrical specimens. These specimens were prepared froma mould. See Figure II.11. This mould consisted of « solid steel cylinder with a hollow cylinder machined within. A matching steel piston wos used to obturate the lower part of the cylinder during amalgam condensation and to eject the specimen after condensation. Two spacers (1 mm thick x 25.4 mm diometer and 8 mm thick x 25.4 mn diometer) were used to establish the length of the specimen. Before condensation, the piston, with the two spacers, wos placed in the cylinder to leave a mould space of 9 mm by 4 mm. Immediately following condensotion, the 1 mm spacer was removed ond the specimen extruded 1 mm to allow cerving with a sharp razor blede. This top layer represents the mercury-rich excess amalgam. The 8 am by 25.4 mm spacer was then removed and the completed specimen ejected and dimensional change measured using a Mikrokator* S5I0Y at a measuring load of 1 gm. The final dimensional change was recorded in micrometers per centimeter at 24 hours after condensation. Mikrokator Johansson, Eskilstuna, Sweden. 83803 eBuoyo TouoysuawTp 404 suewtoads woByous ToaTupuyTA2 yo UoTzOIBdeWd Joy pasn p[now jo voTzO;UeSeude4 oTZOUMOsBOTg [1 "11, euNB TY sussods wu g pud wu, yo <—— Jorowes 4934y wu g 4990ds 122. wu | 4e20ds uewyoeds wob tows ww y Aq wu 6 uw 6 ptnow 44709, 123. CHAPTER IIT RESULTS III.1 Adaptation of Amalgam to the Cavity Wall The adaptetion of different amalgam alloys to the vertical and pulpal walls of Closs I restorations for natural teeth is presented in Tobles 11.1.1, I1f.1.2, [11.1.3 and I1T.1.4. The adaptation of all alloys to the vertical walls is better thon the adaptation to the pulpal wall in both natural and porcelain teeth. See Figure III.1.1. The adoptation of alloys to the internol cavity walls is not the some for natural and porcelain teeth. Generally, omolgam alloy is better adapted to the natural tooth surface thon to the porcelain tooth surface. Although the adaptation of omalgom to porcelain was not os good as the adaptation to natural tooth structure, mony trends in adaptation following changes in condensation techniques were noted to be similor to that of the natural tooth. The porcelain specimens were used'as control to determine the effect of microscopy techniques on the size of the restoration-tooth interface gap. Some cracking and crazing of natural tooth structure was noted after examination using « scanning electron microscope, see Figure III.1.2 Light ond sconning electron microscopy measurements were taken of each specimen and both measurements compared favourably. Some cracking was seen with the natural tooth specimens in the dentine, at the junction of the vertical and pulpal walls when using « scanning electron microscope (SEM). See Figure III.1.2. There was 124, a very slight increase in the omalgom-naturel tooth interface along the pulpel floor for some specimens viewed with the S.E.M. The differences in adaptation for natural and porcelain tooth structure may be due to a difference in the surface energy ond molecular structure of eoch matericl. Analysis of the adaptation of different elloys will be mainly confined to naturel tooth structure. For statistical onalysis all dota wos analysed using the foctoricl onalysis of voriance technique, followed by multiple pair- wise testing of meons. Analysis was performed with the Generalised Linear Interactive Modelling (GLIM) statistical package (Baker and Nelder, 1978) on the PDP-1090 computer at the University of Queensland. (i) Adaptation to vertical walls From Table III.1.1 it can be seen that the mean interface gap distonce of oll omalgam alloys when using o heavy force, is less than five micrometers. Some alloys appear to adapt more closely to the vertical wall thon others. Generally, when a heavy force, small increment condensation technique is employed, adaptation to the vertical wall of a Closs I cavity preperation is good with only slight differences between the vorious alloy systems. When a light force, smoll increment condensation technique is employed the adaptation of different alloy systems to the vertical wall is similar to or slightly less than when a heovy force is used. The adaptation of G.C.’s Hi Atomic ond Spheriphase C has changed only slightly. 125. Table 11.1.1 Comparison of the mean natural tooth-restoration gap width for vertical wails, for all alloys. Condensation techniques included heavy force/smali increment ond light force/small increment. Alloy Mean gop size for vertical walls (pm) Heavy force Light force Amolcop F 2.7 (0.83)* 4.2 (1.33) G.C.'s Hi Atomic 4.7 (1.70) 4.9 (2.36) ‘Amoleap-non-gamma-2 1.9 (0.85) 5.2 (1.82) Disperselloy 2.9 (1.10) 5.8 (2.03) Spheriphase C 2.9 (1.08) 2.8 (0.56) F400 3.8 (0.81) 4.4 (1.43) Ultracops 2.5 (0.62) 3.1 (1.28) Tytin 2.0 © (0.83) "3.3 (0.02) stondord deviation in brackets Table 11.1.2 Comparison of the mean natural tooth-restoretion gap width for pulpal walls, for all alloys. Condensation techniques included heavy force/small increment and light force/small increment. Alloy Mean gap size for pulpal walls (pm) Heavy force Light force Amoleop F 4a (2.44)¢ 6.4 (2.91) 6.C.'s Hi Atomic 8.8 (2.92) 8.5 (3.02) Amalcap-non-gomma-2 8.9 (2.74) 10.9 (3.05) Dispersal loy 6.5 (2.28) 9.5 (1.87) Spheriphase C 7.4 (2.00) 7.1 (1.52) F400 8.0 (3.15) 12.5 (3.69) Ultroceps 5.1 (0.91) 7.1 (1,36) Tytin 5.3 (1.31) 8.3 (0.14) stondard deviation in brackets 126. Table III.1.8 Comparison of the mean porcelain tooth-restoration gap width for vertical wolls, for oll alloys. Condensation techniques included heavy force/small increment and light force/small increment Alloy ; Mean gop size for vertical walls (pm) Heavy force _ Light force Amaleop F 3.9 (1.14)* 4,5 (2.24) G.C.'s Hi Atomic 7.8 (2.51) 11.0 (2.22) ‘Amalcap-nen-ganma-2 3.8 (1.36) 6.8 (1.26) Dispersal loy 4.4 (1.03) 8.7 (2.63) Spheriphase C 3.7 (0.73) 4.9 (1.68) F400 6.6 (1.08) 10.5 (2.50) Ultrocops 3.4 (0.50) 4.7 (1.42) standard deviation in brackets Table I1I.1.4 Comparison of the mecn porcelain tooth-restoration gap width for pulpal walls, for all alloys Condensation techniques included heavy force/small increment and light force/small increment. Alloy Mean gop size for pulpal walls (ym) Heavy force Light force Amalcop F 6.5 (1.74)* 7.1 (0.72) G.C.'s Hi Atomic 13.8 (3.31) 25.8 (5.00) ‘Amaleop-non-gonma-2 9.5 (0.81) 10.5 (3.35 Dispersalloy 11.9 (4.08) 20.1 (3.81) Spheriphose C 10.9 (3.30) 15.8 (2.19) F400 14.6 (1.78) 22.8 (4.80) Ultrecops 7.4 (1.32) 10.3 (2.34) * stondord deviation in brackets (i) 127, (iii) tooth omal gam porcelain omelgam Figure I1I.1.1 Scanning electron photomicrographs comparing the adaptation of amalgam to natural tooth structure (i) vertical walls ond (it) pulpal walls, and porcelain tooth structure (iii) vertical walls ond (iv) pulpal walls. (Magnification 275 x 128. (ii) Figure I11.1.2 Scanning electron photomicrographs of (i) natural tooth showing cracking after scanning electron microscopy ond (ii) a porcelain tooth containing an amalgam restoration (Magnification 12.5 x) 129. Statisticol anolysis using oll the data from all the specimen types prepared from different condensation techniques, compares the adaptation of each alloy to the vertical, walls. See Toble III.1.5 ond Figure I1I.1.3. The statistical analysis set out in Table III.1.5 ond Figure III.1.3 identifies any significant difference in adaptation of different olloy systems to the vertical walls. The tooth-restoration gop is (i) significantly greater for G.C.’s Hi Atomic than for Spheriphase C, ot the confidence level of p < 0.01 (ii) significantly greater for G.C.’s Hi Atomic thon for Tytin, Ultracaps and Amalcop F, ot the confidence level of p< 0.05; (iii) significantly greater for Dispersalloy thon for Tytin, at the confidence level of p< 0.05. There is no significant difference between the remcining alloys in their adaptation to the vertical walls: (ii) Adoptation to pulpal walls Tobles III.1.2 and III.1.4 disploy the adaptation of ol amalgam alloys to the pulpal walls of porcelain and natural teeth when heavy and light condensation forces are used on small increments. The adaptation of onalgam to the pulpal wall of noturol teeth is generally better thon the odoptation to porcelain, for both condensation techniques. Adaptation to the pulpal wall is better when a heavy force is employed for both porcelain ond natural teeth. G.C.'s Hi Atomic and Spheriphose C show similar odeptation to the pulpal wall of a natural tooth for light and heavy force condensction methods. 130. Table I11.1.5 Comparison of the overell means and standerd errors of the mean, for the adaptation of elloys to the Vertical wolls of natural teeth. Alloy Notural tooth-vertical wall mean gop (jm) Amaleop F 3.6 (0.41)* G.C.'s Hi Atomic 4.8 (0.44) Amalcap-non-gomno-2 3.5 (0.65) Dispersal loy 4.5 (0.57) Spheriphase C 2.8 (0.62) F400 4. (0.65) Ultracaps 2.8 (0.69) Tytin 2.5 (0.92) stondard error of the mean in brackets c Aor non-gonma-2 G.C.'s Hi Amalcap F Atomic Dispersolloy Amalcap- Ultracaps F400 Spheriphase 2 (uxt) e0un3s tp doB uvew 131, Ao{T2 yone Joy 'STTOM TOOTZI0A 40} do6 uoTz0403504-43003 UDew 043 yo Ydoug E*{“]]] ouNdTY 132. _The size of the interface gop distance along the pulpal wall is greater than for the vertical wall. The statistical onolysis for all data concerning the pulpal wall is displayed in Table I1I.1.6 and Figure III.1.4. The tooth- restoration interface gop is: (a) significantly greater for F400 than for Ultracaps, ot the confidence level of p < 0.05; (144) significontly greater for G.C.’s Hi Atomic than for Ultracops ond Ancleap F, at the confidence level of p < 0.05. There is no significant difference in the adoptation of the remaining alloys to the pulpol well. TIT.1.1 Lathe and Spherical Alloys (i) Vertical wolls The results in Tables III.7.7 ond III.7.8 show that the adaptation to the vertical walls of natural tooth structure is slightly better for Analcop F than for G.C.’s Hi Atomic. When a light packing force on a smoll increment is employed the adoptation of Amalcop F ond G.C.'s Hi Atomic is similor. The heavy force condensotion technique produces closer adaptation of Amalcap F to vertical wolls thon it does for G.C.’s Hi Atomic. Increment size does not appear to affect the adeptation of either spherical or lathe-cut alloys to the vertical walls. (ii) Pulpel wolls The results are displayed in Tables III.7.9 ond I1I.7.10. Adoptetion of Analcap F (lathe) to the pulpal wall is best when a 133. heavy force is employed during condensation. Adaptation of the lathe-cut alloy is better than the spherical alloy when o heavy force small increment is used. Increasing the increment size can reduce the availble packing force which is transferred to the pulpal aspect of the increment and so result ino less well adapting restoration to the pulpal wall. This slightly affects the adaptation of Amalcap F to the pulpcl wall but hos little effect on G.C.'s Hi Atomic which exhibits good adaptation to the pulpal wall for « heavy force; large increment packing technique A light force reduces the adaptation of lothe-cut amalgam to the pulpol wall but adaptation of the sphericol alloy appears to change only slightly. A large increment combined with o ligh condensation force reduces the adaptation of both alloy types to the pulpol well. Lack of adaptation is similar for both alloys. The lothe-cut alloy, although it can produce o better adapting restoration under ideal condensation situations, its adaptation may be reduced by changing force and incrément size. The sphericol omalgom produces reasonably adapting restoration for small increment with heavy or light condensation force. This confirms the findings of Wing (1970) on the clinical use of sphericel particle amalgams Good adaptation is also achieved when a heavy load is used to pack a large increment. Adaptation is reduced when o light load/large inerement is used, as condensation force is unable to be satisfactorily tronemitted to the pulpol aspect of the cavity preparation to adapt the omolgam in this area. The lathe-cut alloys, which have more “packability” than the sphericol particle alloys require more force pulpally to produce a 134. Table III.7.6 Comparison of the overall means and standard errors of the mean, for the adaptation of alloys to the pulpal walls of natural teeth. Alloy Notural tooth-pulpal wall mean gap (ym) Amaleap F 6.9 (0.91)* G.C.'s Hi Atomic 10.0 (1.00) Amalcap-non-ganma~2 9.9 (1.45) Dispersalloy 8.1 (1.27) Spheriphase C 7.3 (1.38) F400 10.2 (1.45) Ultracaps 6.0 (1.52) - Tytin 6.5 (2.04) standard error of the mean in brackets 135. 4 2 , 8 3 8 Bb, BR o8 8 En? 5 B 8s “2 gy 2 &£ 8 ¢ rs = 3 # ae fey oz FC —- T T T T T T — ‘ z = iy 1S 9 L 8 6 Ob be +hoTTO Yooe Jog ‘sTTOM TodTnd 404 do6 vor30s0350/-43003 uDals 043 Jo ydoup we LITT ese6sy xoTY (why 90u038Tp dob voew 136, Toble I11.1.7 Comparison of the meen tooth-vertical wall space for natural teeth when different condensation techniques are employed. Gap size (micrometers) Condensation technique Alloy heavy force/ | light force/| heavy force/|light force/ small ine | smell ine | lorge ine _|lorge ine Amalcap F 2.7 (0.83)* | 4.2 (1.33) 2.6 (0.86) }4.8 (3.02) G.C.'s Hi 4.7 (1.70) 4.9 (2.36) 4.6 (0.95) |5.2 (2.66) Atomic standard deviation in brackets Table III.1.8 Comparison of the mean ‘tooth-vertical wall space for porcelain teeth when different condensation techniques are employed. Gap size (micrometers) Condensation technique Alloy heavy force/| light force/| heavy force/|light force/ small inc__| small ine large inc [large ine Amaleop F |.3.9 (1.14)* | 4.5 (2.24) | 3.6 (0.77) |7.7 (2.18) G.c.'s Hi [7.8 (2.51) | 11.0(2.22) | 9.8 (3.36) |12.5(2.59) Atomic Toble III.1.9 standard deviotion in brackets Comparison of the mecn tooth-pulpal wall space for natural teeth when different condensation techniques are employed. Gop size (micrometers) Condensation technique Alloy heovy force/| light force/| heavy force/|light force/ smoll inc | small ine | large ine |large ine Amoleap F 4.1 (2.44)* | 6.4 (2.91) 4.5 (1.44) 12.6 (7.79) G.C.'s Hi 8.8 (2.32) 8.5 (3.02) 7.9 (3.79) 15.3 (3.89) Atomic standard deviation in brackets Toble III.1.10 Comporison of the mean tooth-pulpal wall space for porcelain teeth when different condensation techniques ore employed Gop size (micrometers) Condensation technique Alloy heavy force/ |light force/| heavy force/ light force/ small inc _|small ine | large ine _|large ine Amoleap F 6.5 (1.74]* |7.1 (0.72) | 5.9 (0.40) [14.8 (6.91) one HE Ji3.8(3.31) — |25.8(5.00) | 19.0(2.11) [30.2 (6.10) standard deviation in brackets 137. similor adoptation to the pulpal wall. A small increment is more likely to allow the transmission of o heavy force to the pulpal well, then @ large increment. I11.1.2 The Effect of Force Tobles I1I.1.1, III.1.2, 111.1.3 and III.1.4 present the adaptation of the alloy to. the cavity preparation when heavy ond light forces are employed. (i) Vertical walls For vertical walls increasing the condensation force from light to heavy improves the adaptation for most alloys. No real improvement is seen in the adaptation of G.C.‘s Hi Atomic and Spheriphase C to the vertical wall of natural teeth when condensation force is increased from Light to heavy ono small inerement A stotistical analysis of the effect of force on the adoptation of oll omalgam alloys to the verticel wall of natural and porcelain teeth wos performed. See Table III.1.11. Table III.1.1 ond Figure I1I.1.5 show that under similar conditioné the adaptation of amalgam to natural tooth is better than the adaptation to porcelain. Force does influence the adaptation of amalgam to the vertical wall for both porcelain and natural teeth. The adaptation of amalgam to the vertical walls of natural tooth structure is better for a heavy force than when a light force is used. This is significant at the p < 0.001 level 138, Toble III.1.11 Comparison of the effect of force on the adaptation of amalgam to the vertical walls of natural and porcelain tooth structure Tooth type Mean gop size (micrometers Force Heavy Light Natural 3.1 (0.29)* 4.5 (0.29) Porcelain 5.1 (0.35) 8.0 (0.36) stondord error of the mean in brackets Toble I1I.1.12 Comparison of the effect of force on the adaptation of amalgom to the pulpal walls of natural and porcelain tooth structure. Tooth type Mean gap size (micrometers) Force Heavy Light Notural 7.0 (0.64)* 10.2 (0.65) Porcelain 10.3 (0.78) 17.4 (0.79) standard error of the mean in brackets 139. 9 3 © Porcelain 7 6 eel © Porcelain gop © Noturel distance 4 (um) 3 © Natural Heavy Light Force Force Figure I11.1.5 Graphs the difference ‘in adaptation to the vertical wall when light and heavy condensation forces art used on porcelain and natural teeth, 140. The adaptation of amalgam to the vertical walls of porcelain tooth structure is better for a heavier force than when a light force is used. This is significant at the p 4 0.001 level. (41) Pulpol_ walls From Tables III.1.2 and III.1.4 it can be seen increasing the condensation force from light to heavy improves the adaptation of most amalgam alloys to the pulpal wall of the covity preparation. Force does not appear to improve the adaptotion of G.C.’s Hi Atomic (3 7:5) and Spheriphose C to the pulpal wall when a small increment size is condensed. Table III.1.12 and Figure III.1.6 show that under similar conditions the adaptation of omalgam to tooth is better thon the adaptation to porcelain. Force influences the adaptation of the amelgom alloy to the pulpal wall for both porcelain and natural teeth. The statisticol onolysis performed showed the adaptation of omalgom to the pulpal walls of natural tooth structure is better with 0 heavier force than when a light force is used. This is significant at the p <0.001 level. The adaptation of amalgam to the pulpal walls of porcelain tooth structure is better for a heavier force than when a light force is used. This is signficont at the p< 0.001 level Generally, @ heavy force will produce a better adapted amalgam restoration to vertical and pulpal walls of both natural and 141. porcelain tooth structure. 11I.1.3 Increment Size Changing the increment size wos only exomined within the G.C.'s Hi Atomic (3.7:5) ond Amalcap F (5.5:5) amalgam alloys From Table III.1.13 it can be seen thot changing increment size for Anolcop F ond G.C.'s Hi Atomic does not affect the adaptation of amelgam to the vertical walls of notural and porcelain tooth ‘structure. The condensation force has a greater influence on vertical woll adaptation. From Table III.1.14 it can be seen that changing the increment size from small to large may affect the adaptetion of Amalcap F and G.C.’s Hi Atomic to the pulpol wall. The réduetion in edeptation is greater for a large increment when a light force is used, than for a small increment. From a statistical analysis of the date from Amalcap F and G.C.'s Hi Atomic it is shown there is a force/increment interaction. When o heavy force is used to condense smoll and large increments to the pulpal woll, the increment size hos little influence on the edoptation. See Toble III.1.15 and Figure I11.1.7. Light condensetion force on small increments produce better adaptation to the pulpal wall thon a light force on a lerge increment. This difference is significant at the p< 0,001 level. A heavy condensation force on a small increment produces better adoptotion to the pulpal wall than a light force. This is 16 1" 16 TTT 15 BE iP mean gop 10 distance (um © porcelain © porcelain — enctural enatural gh Figure III.1.6 Heavy Light Force Force Graphs the difference in adaptation to the pulpal woll when light and heavy condensation forces are used on porcelain ond natura! teeth. 143, Table III.1.13 Comparison of the mean gap distance along the vertical wall for natural and porcelain teeth: for different condensation methods Gap size mean (micrometers) Alloy/ Condensation technique tooth type | heovy force/{ heavy force/ | light force/flight force/| small inc | large ine small inc | lorge ine Amaleop F noturol | 2.7 (0.83)* | 2.6 (0.86) | 4.2 (1.33) |4.8 (3.02) porceloin | 3.9 (1.14) | 3.6 (0.77) | 4.5 (2.24) [7.7 (2.18) G.C.'s Hi Atomic naturel | 4.7 (1.70) | 4.6 (0.95) | 4.9 (2.36) |5.2 (2.66: porcelain | 7.8 (2.51) | 9.8 (3.36) 11.0(2.22) |12.5(2.59) standard deviation in brackets Table III.1.14 Comparison of the mean gap distance along the pulpal wall for natural ond porcelain teeth, for different condensation methods Gap size mean (micrometers) : Alloy/ Condensation technique tooth type | heavy force/|heavy force/ { light force/[light force/ small ine | large inc | small inc | large ine Amalcop F naturel 4.1 (2.44) 14.5 (1.44) | 6.4 (2,91) [12.6 (7.79) porcelain | 6.5 (1.74) |5.9 (0.40) | 7.1 (0.72) ] 14.8 (6.91) 6.C.'s Hi Atomic natural 8.8 (2.32) |7.9 (3.79) | 8.5 (3.02) | 18.3 (3.89) porceloin | 13.8(3.31), |19.0(2.11) | 25.8(5.00) | 30.2 (6.10) * stondard deviation in brackets 144. significant at the p<0.001 level. A heavy condensation force on a large increment produces better edaptation to the pulpal woll than a light force. This i: significant at the p < 0.001 level. Ideally, a heavy condensation force and small increment size should be employed to attain the maximum adaptation of amalgam to the pulpal wall of @ Closs I restoration TI1.1.4 Effect of Alloy Type All alloys were better adopted to vertical walls thon to pulpol wolls. See Table I11.1.16. The behaviour of the adaptation of the lathe-cut omalgom alloys wos similor. All showed reduced adaptation to the’ vertical ond pulpol wolls when. a light condensation force was used. Adaptation of these amalgam alloys did change with changes in monipulative procedures. The adaptation of G.C.'s Hi Atomic alloy to vertical walls did not oppear to be altered by heavy or light condensation forces. See Toble I1I.1.16. Tytin, the other spherical alloy, showed better adoptetion to all walls with a heavy condensation force. Overall. Tytin appeored to be better adapted than G.C./s Hi Atomic. The better adaptation of Tytin may be a reflection of the greater pockability of this alloy compared with G.C.'s Hi Atomic. In spherical cmalgom systems, packability usuolly relates to the groded sizes of sphericol particles used. 145. Condensation force did not appear to alter the adaptation of Spheriphase C to both pulpel ond vertical walls of the cavity preparation. All alloys showed good adoptation to the vertical walls for o heavy force ond small increment condensation technique. G.C.'s Hi Atomic (spherical "Ag,Sn") showed the least adaptation. All alloys adopted similarly to the verticel wall for light force/small increment condensation. Lothe-cut alloys, Analcap F ond Ultracaps, and spherical alloy Tytin, were best adopted to the pulpal walls when condensed with a heavy load on small increments. The remaining alloys adapted to the pulpal wall similarly. Lothe-cut alloys Amoleap F and Ultracaps and a spherical-lathe mixed alloy Spheriphose C were best adapted to the pulpal wall for the light force/small increment condensation technique. There did not oppear to be any real correlation between the type of alloy ond adoptction. Certain trends were evident, but it is difficult to estimcte the adeptotion of an amelgom alloy system eccording to the porticle types present. Other factors such as the mercury to clloy ratio, plosticity of the mix, trituration time composition of the alloy end dimensional change moy affect the adoptation of on amolgom clley. From the alloys tested in this study it wos not possible to determine the likely behaviour of different alloy systems. See Tables I1I.1.17 and III.1.18. 146. III.1.5 Largest Marginal Gop Generally, the largest tooth-restoration gap was seen in the undercut area, especially in the cose of sharp undercuts. Difficulty is experienced when condensing oll omalgom types in this region Spheriphase C was well adapted to the undercut areas. Areas of marginal porosity along the internal cavity wall are areas of poor adoptation. Generally the maximum gop area is greater for specimens in which light condensation has been employed. See Toble III.1.9. Conclusion The overall adaptation of all alloys tested was of an ecceptable nature, No alloy, generolly, was poorly adapted to the covity woll from point of view of clinical acceptability. Adoptation may be influenced by the alloy particle: size and shope, the mercury to alloy ratio offecting the plasticity of the trituroted mass, the composition ond the behaviour of that composition ond the dimensional change of the alloy. With so many variables effecting the adaptation it is not possible to determine the behoviou of on elloy on the basis of one varicble alone. Qilo (1979) tried to determine odaptotion according to dimensional change. From the present study there was no observed relationship between dimensional change ond adaptation. Other properties such os plasticity of the mix and alloy type must be considered. 147. Toble III.1.15 Comparison of the mean adaptation of the omalgam alloy to the pulpal wall when different condensation forces are used to pack large ond small size increments. Gop size mean (micrometers) Increment size Condensation force Heavy L Light Small 8.4 (0.60)* 12.0 (0.62) Large 8.3 (1.33) 17.2 (1,23) * standerd error of the mean in brockets 148. or we © light force 16 15 14) 13] rab © light force ath meon gop 10h distance (pm) st Sheovy force . heavy force Smal Large Increment Size Figure I1I.1.7 Grophs the mean gop distonce of the pulpal wall for oll alloys when condensation procedures ore altered. 149. Toble I1I.1.16 Comparison of the interface gap distance for all alloys to the pulpol and vertical wells of natural teeth when heavy ond light condensation forces ore used to condense small increments Alloy Particle Gap size mean (micrometers ie Vertical walls Pulpal walls Force Force heavy light heavy light Amalcap F {100% lothe | 2.7(0.83)*| 4.2(2.24) | 4.1(2.44) | 6.4(2.91) F400 00% lothe | 3.8(0.81) | 4.4(1.43) | 8.0(3.15) | 12.5(3.69 Ultracaps 100% lothe | 2.50.62) | 3.1(1.28) | 5.1(0.91) | 7.1(1.36 G.C.'s Hi [100% 4.7(1.70) 4.9(2.36) 8.8(2.32) 8.5(3.02) Atomic sphericol Tytin jLoo% 2.0(0.83) 3.3(0.02) 5.3(1.31) 8.3(0.14) spherical Amalcap-non| 70% lathe | 1.9(0.85) | 5.2(1.82) | 8.9(2.74) | 10.9(305) =gomma-2 | 30% spherical Dispers- | 70% lathe |2.9(1.10) | 5.8 (2.03) | 6.5(2.28) | 9.5(1.87)) alloy 30% spherical [Spheriphase | 70% 2.9(1,08) | 2.8(0.56) | 7.4(2.00) | 7.1(1.52)] c spherical 30% lathe * stondard deviotion in brackets Table II1.1.17 Lists the mereury to alloy ratio for each elloy Alloy Mercury to alloy ratio Analcop F 5.5:5 F400 5.6:5 Ultroceps 5.0:5 G.C.'s Hi Atomic 3.735 Tytin 3.7535 Analcep-non-ganma-2 6.55:5 Dispersal loy 5.5:5 Spheriphase 5.535 150. Toble I1I.1.18 Dimensional change of amalgam alloys after 24 hours Alloy Dimensionel chenge jm/em Amaleop F 5.5:5 2 Amalcop F 8:5 427 Amalcap F 4:5 4 G.C."s Hi Atomic 3.7:5 -14 G.C.'s Hi Atomic 5.4:5 -14 ‘Amalcop-non-gamma-2 42 Dispersal loy 45 Spheriphase C “8 F400 -3 Ultrocops 16 Tytin 7 151. Toble III.1.19 Lists the greatest notural tooth-restoration gop size and'the location Gop size (micrometers) Alloy Condensation technique heavy force/small inc | light force/small inc gop size location gap size location Amalcap F 160 undercuts |260 undercuts F400 120 undercuts |120 undercuts Ultrocaps 60 undercuts 80 undercuts pulpal ond vertical walls vertical walls -C.'s Hi Atomic | 100 undercuts {100 undercuts ytin 50 vertical 80 undercuts woll elcap-non-gomma-2| 160 undercuts |120 undercuts pulpol wall ispersal loy 160 undercuts |240 undercuts pulpal wall pheriphase C 100 vertical .wall|200 verticol woll 152. II1.2 Mercury to Alloy Ratio Voriotion in the operative procédure of condensing an amalgam into a cavity preparation will affect the life ond quality of the restoration. Varying the mercury to alloy ratio may also affect the finished restoration. It is thought that porosity and mercury wecken the restoration (Jérgensen et ol, 1966) and lead to early marginal breakdown. The mercury to alloy ratio was varied from the manufacturers recommended ratio. Class I cavity preparations were restored ond the resultant restoration examined for internal cavity wall adoptation and porosity. Amalcap F was used ot the recommended ratio 5.5:5,' a wet rotio 8:5 end a‘dry mercury to alloy ratio of 4 ~ G.C."s Hi Atomic was exomined at the ratio of 3.7:5 and the wetter mix of 5. 5 mercury to alloy ratio. Closs I restorations were prepared in natural and porcelain teeth. During condensation increment size wos always small restorations were packed to excess and carved back, the only variation involved wos the difference in condensation force. Half of the specimens were packed using a heavy load (27-36N) ond the other half packed with o light lood (9-13N). Table II11.2.1 displays the mean width (in micrometers) of the space at the tooth-restoration interface for both spherical ond lothe- cut omalgoms at oll ratios. The vertical ond pulpal wall adaptation were difficult ond were exomined separately. (i) Amalcap F A, Heavy lood, small increment 153, Table I11.2.1 A comparison of the mean widths of the natural tooth- restoration interface gap for both Analcap F and G.C.'s Hi Atomic amalgam alloys ot ratios tested Alloy Ratio |lVerticol wall gap (pm) | Pulpol wall gap (ym) Imercury: [Heavy force Light force| Heavy force] Light force clloy malcap F | 5.5:5 2.7 (0.83)) 4.2 (1.93) | 4.1 (2.44) [6.4 (2.91) JAnaleap F 8:5 2.9 (2.48) 3.5 (2.56) }7.5 (2.62) ]7.9 (4.48) Jmalcap F 435 2.9 (0.73)} 5.2 (2.44) | 9.9 (4.71) [14.7 (9.21) 6.c.'s Hi | 3.7:5 4.7 (1,70) 4.9 (2.36) | 8.8 (2.32) | 8.5 (3.02 Atomic lo.c.’s Hi | 5.425 3.8 (1,80)] 4.8 (2.05) | 9.0 (0.90) [13.5 (5.68 Atomic stondard deviation in brackets 154. Three initiol ratios of Amalcap F were compared in Class I cavity preparations. The adaptation to the verticol walls was similar for oll ratios. The pulpal walls were better odapted by the 5.5:5 Amclcap F than for the adaptation of the wet ond dry mixes. The wet mix wos better adapted to the pulpal wall than the dry mix. It is difficult to adapt dry mixes of lathe-cut omalgom alloy to the pulpal woll as they resist condensation. Adaptation to undercut areas was good for ell mercury to alloy ratios with the adaptation of the 5.5:5 mercury to alloy ratio being best. Adaptation of the dry mix was slightly less thon the wet mix to undercut creas. The manufacturer's ratio Analcap F produced anolgam containing fine to moderate sized porous regions widely distributed. The wet ratio Analcop F produced omalgams with fine sized porous regions which were densely distributed throughout the body of the restoration. The porous regions of the dry Analcap F mix varied from small to lorge in size ond wore densely distributed throughout the omalgam restoration. See Figures I11.2.1, I11.2.2 and III.2.3. B. Light load, smell incremen: Using a lighter load on the wet, dry ond normal mixes of Amalcap F omalgam resulted in a greater difference in the observed obility of each ratio type to adapt to undercuts, and vertical and pulpal walls. Amalcap F normal ratio produced the best adapted restorations to the pulpal walls and undercuts, adaptation to vertical walls was good. The 8:5 Amaleap F mercury to alloy ratio amalgam restorations were less well adapted to the undercut crecs and pulpal well, the edaptation to the vertical walls wos best for ol Amolcop F rotios restorations. The dey mix Analcop F 155. Figure 111.2.1 Closs I restoration of Amolcap F at 5.5:5 mercury to alloy rotio. Restoration is well adapted to verticel ond pulpal walls ond undercuts, porosity is minimal. Condensation technique of heavy force, small increment was employed. Light photomicrograph magnificetion x 50. 157. restorations were less well adopted to undercuts, pulpal ond vertical walls. The dry mix of Analcap F produced the least adaptation of the three ratios. Porosity of the normal Amalcap F ratio restorations varied from fine to the occasional large area. The distribution of the porous areas throughout these restorations wos relatively dense. The wet Analeap F mix produced porous creas of the predominontly fine size with the occosionel gross orea and the porous areas were also very dense for the restorations of the dry Analcap F mix. The numerous porous creas were much larger in size than for the other restorations. See Figures I1I.2.4 and II1.2.5. A light condensation force for Analcap F normel ond dry ratios did not seem adequate to permit proper condensation of the amalgam alloy in the pulpel third in some restorations. See Figure IIT.2.6. (ii) 6. ts Hi Atomic Two ratios were selected, the dry 3.7:5 ond a wet 5.4:5 mercury to alloy for G.C.'s Hi Atomic alloy to examine adaptation ond porosity. Two condensation techniques were used. The incremen| size wos maintained as smell, the force used wos heavy (27-36N) or Light (9-13N). A. Heavy force, small increment The heavy condensation force for both clloy to mercury ratios produced good adaptation to undercut areas ond to the vertical and pulpol wolls. Adaptation of 5.4:5 mercury to olloy ratio G.C.’s H! Atomic to the vertical walls of the cavity was slightly better Figure 11.2.4 158. Class I restoration of Amalcap F at 5.5:5 mercury to alloy retie. Condensation technique of light force, small increment employed. Light photomicrograph, magnification x 50. yootjyuBow ‘ydosBossTwozoyd 3y8TT F TTOWs ‘eou0y 4Y8TT yo enbruyse, F304 Kop {o 03 Aunouow gig 30 “pg X UoT;D9TyTUBOW ‘YdosBososwozoyd 34817 “ee * “pedotdwe quowessut [TOws ‘0405 348TT 40 ent spokoqdue yuewos vorzosuspucy “oT;0u AOTTO 03 AunoueW giy 30 uoT3psuspucy g dospowy yo voT;os03seu | SSOT) 9°Z°II] SuNBTy 4 dooToWy yo voTz0s03504 | SSDI) g°Z" ITT eunbry 159. 160. Similar adaptation of both alloy ratios to the pulpal wall was observed. The porosity of the dry ratio G.C.’s Hi Atomic renged from fine to moderate in size and was very densely scattered through the onalgom body. The porosity of the 5.4:5 mercury to alloy ratio omclgam wos slightly smaller in size, ronging from small to moderate size and was moderately distributed through the amalgam body but not as densely os for the drier mix. See Figures III.2.7 and 11.2.8. B. Light lood, small increment. The adaptation of dry ratio G.C.’s Hi Atomic to undercuts wos not as well adopted os the 5.4:5 mercury to alloy ratio restorations, The lack of adaptation to the vertical wall wos similar for both ratios. G.C.’s Hi Atomic dry ratio wos better odapted to the pulpal woll than the 5.4:5 mercury to alloy ratio, See Figure III.2.9 and 1I1.2.10. Discussion Differing the elloy to mercury retie for en omalgom con offect adaptation to the cavity wall and porosity within the body of the final restoration, It appears best to use the optimum mercury to alloy ratio. Increasing the mercury to alloy ratio of Amalcap F to 8:5 produced on omalgam thet wos too wet ond difficult to handle. This alloy lacked body and was difficult to manipulate. Removal of excess mercury by squeezing was required. The wet mix produced mercury rich restorations that contained a large number of fine porous regions 161. The 4:5 mercury to alloy ratio Amalcop F was also too difficult to manipulate as the resultant amalgam mix wos dry ond difficult to condense. A light condensation load on Amalcap F 4:5 mercury to alloy rotio resulted in an unsatisfactory restoration of poor marginal adaptation and numerous, large porous creas The amalgam of the dry ratio of G.C.'s Hi Atomic produced restorations thet were well adopted ond of less porosity when heavy condensation force was used. A lighter load on 6.C.'s Hi Atomic alloy at this ratio resulted in poorer adaptation and porosity of higher frequency and size, less mercury removal was possible. The ratio of 5.4:5 mercury to alloy for G.C.'s Hi Atomic wos well odapted with light or heavy forces ond the porosity was less than for the drier mix of spherical amalgom. Eose of monipuletion must be considered. The mercury-rich mix of spherical alloy was more difficult to condense than the drier mix. It was not possible to pock the restoration well using a heovy force os the wet alloy slipped from under the condenser. Mercury removal of the wet mixes of omalgom is time consuming, access and visibility are hampered with the excess mercury lying on top of the amalgam during condensation. Nercury-rich amalgams are weak ofter setting. Dry mixes should not be used as they contain much porosity which may weaken the analgom mix. Comparison of Lathe-cut and Spherical Alloys The adaptation of lathe-cut elloy to the vertical walls was better than the adaptation of the spherical alloy for a heavy force small increment condensation technique. Light condensation force on small increments produced similer adaptation of all lathe-cut omalgom 162. “eg x UoT;D9TyTUBOW ‘ydosBosoTWoZOYd 44817 pakoydue yUewesoUs TTOWS ‘92404 kaosy Jo enbTUYoos seg x UoT;09TyTUBOU ‘YdosBoustuORoY4d 3Y8T7 juewesour [TOUS 'e0J05 AaDey Jo pokoTdu~ enbruysez uvosqesuepucy “oT301 AoTTO 03 Aunosow giy*g 30 oTWOIy uorzosuspucy ‘043041 OTTO 03 Aundsew GiZ"¢ 30 TWOIY TH S,°9°9 JO UOT}0403804 | SSOT) g°Z-III eun6ty TH S,°D°9 40 UOT}Ds03S04 ] SST L°Z7IIT eun6ts 163 “ee X LoTzDOTyTUBOU ‘YdosBossTwox0Yg “pedoTdWe “eg x uoTz09TyTUBOW “YdosBo4stWoI0Yyd 246T7 quewosouT T[Ows puoyayB1] feos0y 3481 go enbruyoo =peAoydue yuewss0UT TTOWS PUD 9010} 4YBTT Jo enbruysoa Yoryosuspucy 073d AoT TO 03 Asnouew Gip"G 30 oTWOIy UoT}OSuapUc) -OT}04 AOTTO 0} AundueW Giz"¢ 30 OTHOW TH 8,°)"9 Jo UoTqDs0380e4 | ssOT) OL“ZIIT eun6T4 TH S,°D°9 go uoTg0103Se4 | SSOTD }°Z"III e408TS | | 164. ratios and spherical amalgam ratios to the vertical walls of o Class I cavity preparation. Closest adaptetion to the pulpal floor for a heavy force small increment condensation technique wos seen for Amalcap F normal ratio. Amalcap 5.5:5, Amalcop 8:5 ond G.C.’s Hi Atomic 3.7:5 mercury to alloy ratio for light condensation locd on @ small increment adapted to the pulpal wall in a similor monner. Data was onalysed using the foctorial analysis of variance technique, followed by multiple pair-wise testing of means. Statistical onlysis wos performed with the Generalised Linear Interactive Modelling (GLIM) statistical package (Baker & Nelder, 1978) on a PDP - 1090 Computer. Taking into account both condensation techniques, the adeptotion of Analeap F (5.5:5) to the vertical wall was significantly better statisticolly thon thot of G.C.’s Hi Atomic (3.5:5) at a confidence level of p< 0.05. See Table 11.2.2. Toking into account both condensation techniques, the adaptation of Analcap F 4:5 is significantly worse statistically than the adaptation of Analcap F 5.5:5 to the pulpal wall (significant at p< 0.05). The marginal gaps of G.C.'s Hi Atomic (3.7:5) and G.C.'s Hi Atomic (5.4:5) are significantly greater than the size of the marginal gap along the pulpal wall for Amaleop F (5.5:5) at a confidence level of p< 0,05. See Toble 111.2.3. From Teble III.2.1 it can be seen that as the condensation 165. force decreases the adaptation of Analcap F (5.5:5) end Amalcap F (4:5) to the vertical wall is reduced. There does not appear to be a change in the adaptation of Anclcop F (8:5) to the vertical wall when condensation force is reduced. A reduction in the condensation force greatly reduces the adaptation of Analeap F (4:5) to the pulpal wall. A lighter force will also increase the gap between the tooth and restoration along the pulpal wall for Analcap F (5.5:5). Force does not appear to alter the adaptation of Analeap-F (8:5) to the pulpal wall. From Table III.2.1 it con be seen that as the condensation force decreases, the adaptation of G.C.‘s Hi Atomic (3.7:5) and G.C.'s Hi Atomic (5.4:5) to the vertical walls is not greatly changed. The adeptation of G.C.’s Hi Atomic (3.7:5) to the pulpal wall os force is reduced, shows little interface gap difference. The adaptation of G.C.’s Hi Atomic (5.4:5) is greatly reduced when o light condensation force is used. Overall, it oppears thot adaptation to the vertical walls for both alloys is better than the adaptation to the pulpal wall. Adaptation of the amalgom to the verticol walls is less subject to manipulative vericbles than the adaptation to the pulpal walls. This must be kept in mind when the only retention for a Class I cavity preporation is placed in the buccopulpal or linguopulpal Line ongles.

You might also like