Open access

Heat Treatment of Dental Alloys: A Review

Written By

William Brantley and Satish B. Alapati

Submitted: 02 December 2011 Published: 19 September 2012

DOI: 10.5772/52398

From the Edited Volume

Metallurgy - Advances in Materials and Processes

Edited by Yogiraj Pardhi

Chapter metrics overview

7,074 Chapter Downloads

View Full Metrics

1. Introduction

Metallic materials have widespread use in dentistry for clinical treatment and restoration of teeth. Major areas of usage are: (1) restorative dentistry and prosthodontics (dental amalgam and gold alloy restorations for single teeth, metallic restorations for multiple teeth, including metal-ceramic restorations, removable partial denture frameworks, and dental implants), (2) orthodontics (wires which provide the biomechanical force for tooth movement), and (3) endodontics (rotary and hand instruments for treatment of root canals). Heat treatment of the metal can be performed by the manufacturer, dental laboratory, or dentist to alter properties intentionally and improve clinical performance. Heat treatment of the metal also occurs during the normal sequence of preparing a metal-ceramic restoration, when dental porcelain is bonded to the underlying alloy substrate. Moreover, intraoral heat treatment of some metallic restorations occurs over long periods of time. There is an enormous scientific literature on the heat treatment of metals for dentistry. A search of the biomedical literature in May 2012, using PubMed [http://www.ncbi.nlm.nih.gov/pubmed/] revealed nearly 450 articles on heat treatment of dental alloys. The purpose of this chapter is to provide a review of the heat treatment of metallic dental materials in the foregoing important areas, describing the important property changes, with a focus on the underlying metallurgical principles.

Advertisement

2. Restorative dentistry and prosthodontics

2.1. Dental amalgams

Dental amalgams are prepared in the dental office by mixing particles of a silver-tin-copper alloy for dental amalgam that may contain other trace metals with liquid mercury. The initially mixed (termed triturated) material is in a moldable condition and is placed (termed condensed) directly by the dentist into the prepared tooth cavity, where it undergoes a setting process that produces multiple phases and can require up to one day for near completion. Extensive information about the several different types of dental amalgams are provided in textbooks on dental materials [1,2]. Particles of the alloy for dental amalgam are manufactured by either lathe-cutting a cast ingot or directing the molten alloy through a special nozzle. Both the machining of the lathe-cut particles and the rapid solidification of the spherical particles create residual stress. In addition, the microstructure of the solidified silver-tin-copper alloy has substantial microsegregation. Consequently, manufacturers of the alloy powder for dental amalgam perform a proprietary heat treatment to relieve residual stresses and obtain a more homogeneous microstructure. This heat treatment is of considerable practical importance since it affects the setting time of the dental amalgam after the powder is mixed with mercury. Subsequently, the dental amalgam restorations undergo intraoral aging, which can be regarded as heat treatment, and detailed information about the microstructural phase changes for prolonged intraoral time periods has been obtained from clinically retrieved dental amalgam restorations [3].

2.2. Gold alloys for all-metal restorations

Gold alloys are principally used for all-metal restorations (inlays, crowns and onlays) in single posterior teeth. These alloys are cast by a precision investment process, and the restorations are cemented by the dentist into the prepared tooth cavity. The original gold casting alloys contained over approximately 70 wt.% gold, but the very high price of gold has led to the development of alloys that contain approximately 50 wt.% gold. These alloys also contain silver, copper, platinum, palladium, zinc, and other trace elements, including iridium for grain refinement. Information about the dental casting process and the gold alloys is available in dental materials textbooks [1,2]. Detailed compositions and mechanical properties of specific alloys are available on the website of the major manufacturers. Another valuable reference is the current ISO Standard on metallic materials for fixed and removable dental appliances [4], which stipulates mechanical property requirements. In the normal dental laboratory procedure, gold castings for all-metal restorations are water-quenched after solidification, following loss of the red heat appearance for the sprue. This results in formation of a disordered substitutional solid solution and leaves the alloy in the soft condition, which is preferable since adjustments are more easily made on the restoration by the dental laboratory or dentist. The gold alloy casting can also be placed in the soft condition by heating at 700°C for 15 minutes and water-quenching. The quenched gold casting may be placed in the hard condition by heat treatment at 350°C for 15 minutes and air-cooling. This heat treatment results in formation of ordered AuCu or AuCu3 regions in the disordered matrix of the high-gold or lower-gold alloys, respectively. Examples of changes in clinically important mechanical properties from heat treatment are shown in Table 1 for two gold alloys, where (S) and (H) represent the soft and hard conditions.

In practice, dental laboratories do not perform heat treatments on the cast gold restorations because of the time involved. However, it appears to be fortunate that the gold alloys that contain sufficient copper to undergo ordering will undergo age hardening in the mouth. Figure 1 compares the intraoral aging behavior of a traditional high-gold dental alloy (Type lV) and a special gold alloy containing gallium (AuCu-3wt%Ga) [5].

AlloyVickers Hardness0.2% Offset Yield StrengthPercentage Elongation
Firmilay (74.5% Au)121 (S)182 (H)207 MPa (S)276 MPa (H)39% (S)19% (H)
Midas (46% Au)135 (S)230 (H)345 MPa (S)579 MPa (H)30% (S)13% (H)

Table 1.

Summary of property changes resulting from heat treatment of two gold alloys for all-metal restorations. [http://www.jelenko.com/, accessed August 15, 2012]

Figure 1.

Comparison of the two-week aging behavior at 37°C for a high-gold dental alloy and a dental gold alloy containing gallium that was designed to undergo intraoral aging. From [5] and reproduced with permission.

2.3. Alloys for fixed prosthodontics (metal-ceramic restorations)

Metal-ceramic restorations are in widespread clinical use for restorative and prosthetic dentistry, and are employed for single-tooth restorations and for restorations involving multiple adjacent teeth (fixed prostheses or crown-and-bridgework). An alloy is cast using the precision investment procedure in dental laboratories to fit accurately to the prepared tooth or teeth, and to form a substrate (termed the coping) for the porcelain. After an initial oxidation step that forms a native oxide on the metal surface, one or two layers of opaque porcelain are bonded to the metal, followed by the application of a layer of body porcelain and a surface glaze [1,2]. In order to have a strong bond between the porcelain and metal, which is essential for clinical longevity of the metal-ceramic restoration, the coefficients of thermal contraction for the metal and porcelain must be closely matched, and a difference not exceeding 0.5 ppm/°C is generally desired. Mechanical property requirements for the alloys are stipulated in ANSI/ADA Specification No. 38 (ISO 9693) [6], and the minimum value of 250 MPa for the 0.2% offset yield strength is important, since the thin coping must withstand intraoral forces without undergoing permanent deformation. The metal-ceramic bond strength (termed the bond compatibility index) is measured with a three-point bending test that uses thin cast alloy strip specimens having a centrally located area of sintered porcelain, and a minimum bond strength (shear stress) of 25 MPa is stipulated.

Both noble and base metal alloys are used for bonding to dental porcelain. The current American Dental Association classification has four alloy groups for fixed prosthodontics [7]: (1) high-noble (gold-platinum-palladium, gold-palladium-silver and gold-palladium); (2) noble (palladium-silver, palladium-copper-gallium, and palladium-gallium); (3) predominantly base metal (nickel-chromium and cobalt-chromium); (4) titanium and titanium alloys. Information about these alloys for metal-ceramic bonding is summarized in a textbook on fixed prosthodontics [8]. The principal mechanisms for metal-ceramic bonding are (a) mechanical interlocking from the initially viscous porcelain at the elevated sintering temperatures flowing into microirregularities on the air-abraded cast metal surface and (b) chemical bonding associated with an interfacial oxide layer between the metal and ceramic. These two mechanisms are evident from photomicrographs, found in numerous references [8], of the fracture surfaces for metal-ceramic specimens prepared from a wide variety of dental alloys. This native oxide forms on the cast alloy during the initial oxidation firing step in the dental porcelain furnace. Noble alloys for bonding to dental porcelain contain small amounts of secondary elements, such as tin, indium and iron, which form the native oxide and also increase the alloy strength. However, Mackert et al [9] found that during initial oxidation heat treatment, metallic Pd-Ag nodules formed on the surface of a palladium-silver alloy for metal-ceramic restorations and only internal oxidation occurred for the tin and indium present in the alloy composition. They concluded that porcelain bonding arose predominantly from mechanical interlocking with the nodules. Internal oxidation has also been reported for high-gold [10] and high-palladium [11] alloys for bonding to porcelain, but both alloy types also formed surface oxides [10,12].

The initial oxidation step and subsequent sintering (also termed baking or firing) of the dental porcelain layers causes the alloy to experience substantial heat-treatment effects. Under normal dental laboratory conditions, the porcelain firing sequence is performed rapidly. For example, in one study heating of high-palladium alloys in the dental porcelain furnace was performed at approximately 30°C/min over a temperature range from 650°C to above 900°C, and the total heating time for the several firing cycles at these elevated temperatures was about 45 minutes [11]. Studies [13-15] have shown that the as-cast microstructures of noble metal alloys for bonding to porcelain are highly inhomogeneous in the initial as-cast condition, presumably from substantial elemental microsegregation that occurs during the rapid solidification involved with casting into much cooler investment [1,2]. After simulation of the dental porcelain firing sequence, the noble metal alloy microstructures become substantially homogeneous, and there are accompanying changes in the mechanical properties, as shown in Table 2.

Peaks in Vickers hardness for heat treatments at temperatures that span the porcelain-firing temperature range indicate that influential precipitation processes can occur in some noble alloys for fixed prosthodontics [13,16]. For the gold-palladium-silver alloy in Table 1, heating an as-cast specimen to 980°C caused a pronounced decrease in Vickers hardness, and subsequent heat treatments at temperatures from 200° to 980°C revealed a pronounced peak in Vickers hardness at approximately 760°C. The absence of substantial changes in Vickers hardness for similar heat treatments of the gold-palladium alloy in Table 2 arises from differences in the precipitates that form in the two complex alloy compositions. Figure 2 presents the age hardening behavior of a palladium-silver alloy, where specimens were subjected to isothermal annealing for 30 minute time periods at temperatures from 400°C to 900°C that span the range for the porcelain firing cycles [16]. Bulk values of Vickers hardness were obtained with 1 kg loads, and 25 g loads were used to obtain hardness values for specific microstructural regions. In contrast, research suggests that microstructures of popular nickel-chromium base metal alloys used with dental porcelain are not changed substantially during dental laboratory processing [17].

Alloy TypeVickers Hardness0.2% Offset Yield StrengthPercentage Elongation
Au-Pd-Ag (Neydium)199 (C)218 (P)420 MPa (C)490 MPa (F)6% (C)8% (F)
Au-Pd (Olympia)213 (C)225 (P)500 MPa (C)540 MPa (F)13% (C)20% (F)

Table 2.

Mechanical properties for two noble metal alloy types used with dental porcelain, comparing the as-cast condition (C) and simulated porcelain firing heat treatment (F) [13].

Figure 2.

Annealing behavior of a palladium-silver alloy for fixed prosthodontics, showing changes in Vickers hardness for a heat treatment temperature range that spans the porcelain firing cycles. Reproduced from [16] with permission.

2.4. Alloys for removable prosthodontics

Base metal casting alloys (nickel-chromium, cobalt-chromium and cobalt-chromium-nickel) are popular for fabricating the metallic frameworks for removable partial dentures because of their lower cost [1,2]. Once an active area of dental metallurgy research, studies have found that these alloys have dendritic microstructures in the as-cast condition, because of the absence of suitable grain-refining elements, and that heat treatment is ineffective for producing improved mechanical properties [18]. A more recent publication shows the dendritic microstructures of some current alloys and their mechanical properties [19].

Removable partial denture frameworks have clasps that engage the teeth. These clasps can be cast as part of the entire framework, or alternatively wire clasps can be joined to the cast framework in the dental laboratory [1,2]. Both noble metal and base metal wires for clasps are available [20]. Because of their superior strength compared to the cast base metal alloys, wire clasps with smaller cross-section dimensions can be used with the frameworks, but caution is required during joining in the dental laboratory to avoid overheating that will cause loss of the wrought microstructure. Wire clasps are used in the as-received condition; heat treatment is not recommended before joining to the framework.

2.5. Dental implant alloys

Dental implants in current widespread clinical use are manufactured from CP (commercially pure) titanium or Ti-6Al-4V, and some implants have a thin bioceramic surface coating (typically hydroxyapatite, the principal inorganic constituent of bone and tooth structure). Proprietary heat treatments [21] are performed on Ti-6Al-4V by manufacturers to obtain optimum microstructures for the implants; minimal information is currently available about these microstructures in the dental scientific literature.

Recently, there has been considerable research interest in the development of new titanium implant alloys for orthopedic applications that have improved biocompatibility compared to the Ti-6Al-4V alloy in widespread current use. There is particular interest in the beta-titanium alloys which have lower elastic modulus than Ti-6Al-4V to minimize stress shielding and subsequent loss of the surrounding bone which has a much lower elastic modulus. Stress shielding does not seem to be of concern for dental implants, presumably because of the threaded designs. Biocompatible titanium-niobium-zirconium beta alloys have been investigated, and oxide nanotubes can be grown on the alloy surface by an anodization technique, and subsequent heat treatment can be employed to modify the structure of the nanotubes [22]. In another exciting research area, titanium oxide nanowires have been recently grown on both CP titanium and Ti-6Al-4V using special elevated-temperature oxidation heat treatments in an argon atmosphere with low oxygen concentrations [23]. Both of these special types of surface oxide layers may prove to be useful for dental and orthopedic implants, but future testing in animals will be needed to examine their efficacy.

Advertisement

3. Orthodontics

3.1. Background

Orthodontic wires engaged in brackets that are bonded to teeth, after being deformed elastically during initial placement, provide the biomechanical force for tooth movement during unloading. There are four wire types in current clinical practice: stainless steel, cobalt-chromium, beta-titanium and nickel-titanium [24]. The clinically important mechanical properties are (a) elastic modulus, which is proportional to the biomechanical force when wires of similar dimensions are compared; (b) springback, which is generally expressed as the quotient of yield strength and elastic modulus (YS/E), and represents the approximate strain at the end of the clinically important elastic range; and (c) modulus of resilience, expressed as YS2/2E and representing the spring energy available for tooth movement. (The permanent deformation portion of orthodontic wire activation is ineffective for tooth movement.) Round orthodontic wires are manufactured by a proprietary drawing sequence that involves several stages with intermediate annealing heat treatments. Rectangular orthodontic wires are manufactured by a rolling process utilizing a Turk’s head apparatus. The wire drawing process with the heat treatments greatly affects mechanical properties.

3.2. Stainless steel orthodontic wires

A recent study that investigated stainless steel wires used in orthodontic practice found that most products were AISI Type 304 and that AISI Type 316L (low carbon) and nickel-free ASTM Type F2229 were also available [25]. While standard physical metallurgy textbooks consider the elastic modulus to be a structure-insensitive property, research has shown that the permanent deformation and heat treatments involved with the wire drawing process can substantially affect the elastic modulus of stainless steel orthodontic wires [26,27]. X-ray diffraction has revealed that conventional orthodontic wires manufactured from AISI Types 302 and 304, while predominantly austenitic structure, can contain the α′ martensitic phase, depending upon the carbon content and temperatures involved with the processing [28]. The presence of this martensitic phase accounts for the reduction in elastic modulus for some conventional stainless steel orthodontic wires. In addition, when fabricating complex stainless steel appliances, it is recommended that orthodontists perform a stress-relief heat treatment to prevent fracture during manipulation; a heating time up to 15 minutes and a temperature range of 300° to 500°C appears to be acceptable [29-31]. Heating austenitic stainless steel to temperatures between 400° and 900°C can result in chromium carbide precipitation at grain boundaries and cause the alloy to become susceptible to intergranular corrosion, and heating of austenitic stainless steel wires above 650°C should not be done because loss of the wrought microstructure causes degradation of mechanical properties.

3.3. Cobalt-chromium orthodontic wires

The cobalt-chromium orthodontic wire (Elgiloy) marketed by Rocky Mountain Orthodontics (Denver, CO, USA) contains 40% Co, 20% Cr, 15.81% Fe, 15% Ni, 7% Mo, 2% Mn, 0.15% C carbon and 0.04% Be beryllium (https://www.rmortho.com/, accessed August 15, 2012). Four different tempers (spring quality) are available, and the soft Blue temper is favored by many orthodontists because the wire is easily manipulated in the as-received condition, and then heat treated to increase the yield strength and modulus of resilience. Heat treatment (not recommended for the most resilient temper) is conveniently performed with the electrical resistance welding apparatus commonly used in orthodontic practice, and the manufacturer provides a special paste that indicates when the heat treatment is complete. Alternatively, furnace heat treatment performed at 480°C for 5 minutes has been found to give satisfactory results [32]. An extensive study employing furnace heat treatment (480°C for 10 minutes) for three tempers and numerous sizes of the Elgiloy wires observed increases of 10% – 20% in elastic modulus and 10% – 20% in 0.1% offset yield strength, which resulted in substantial improvement of the modulus of resilience [27]. These changes in mechanical properties arise from complex precipitation processes during heat treatment that are not understood. Many other companies now market cobalt-chromium orthodontic wires, but studies of their mechanical properties and the results of heat treatment have not been reported.

3.4. Beta-titanium and other titanium-based orthodontic wires

Beta-titanium orthodontic wires have the advantages of: (a) known biocompatibility from the absence of nickel in the alloy composition; (b) lower elastic modulus than stainless steel and cobalt-chromium wires, which provides more desirable lower orthodontic force for tooth movement; (c) higher springback than stainless steel and cobalt-chromium wires, which is desirable for the archwire to have greater elastic range; and (d) high formability and weldability, which are needed for fabrication of certain appliances [24]. A recent study [25] of commercially available titanium-based orthodontic wires revealed that most products are Beta III alloys [21] containing approximately 11.5 Mo, 6 Zr, and 4.5 Sn, similar to the original beta-titanium wire introduced to orthodontics [33,34]. Beta C [21] and Ti-45Nb beta-titanium and Ti-6Al-4V (alpha-beta) wire products are also available [25]. Heat treatment is not performed by the orthodontist on these wires, but care with the wire drawing and intermediate heat treatments by the manufacturer are essential for obtaining the desired mechanical properties. These processes must be conducted under well-controlled conditions because of the highly reactive nature of titanium.

3.5. Nickel-titanium orthodontic wires

Following the pioneering work of Andreasen and his colleagues [35,36], near-equiatomic nickel-titanium (NiTi) wire was introduced to orthodontics by the Unitek Corporation (now 3M Unitek) [37]. This wire had the advantages of a much lower elastic modulus than the stainless steel and cobalt-chromium wires available at the time and a very large elastic range. The clinical disadvantage is that substantial permanent deformation of this wire is not possible to obtain certain orthodontic appliances that can be fabricated with the three preceding, highly formable, alloys. The original nickel-titanium wire had a work-hardened martensitic structure and did not exhibit the superelastic behavior (termed pseudoelasticity in engineering materials science) or the true shape memory characteristics displayed by subsequently introduced NiTi wires [1,38-41]. These nickel-titanium wires have been a very active area of research.

The mechanical properties of the nickel-titanium orthodontic wires are determined by the proportions and character of three microstructural phases: (a) austenite, which occurs under conditions of high temperature and low stress; (b) martensite, which occurs under conditions of low temperature and high stress; and (c) R-phase, which forms as an intermediate phase during the transformation between martensite and austenite. Very careful control of the wire processing and associated heat treatments, along with precise compositional control, by the manufacturer are needed to produce nickel-titanium wires with the desired superelastic, nonsuperelastic, or shape memory character [42,43].

Heat treatments have been exploited by manufacturers to control the orthodontic force ranges produced by nickel-titanium archwires [39]. Heat treatment temperatures have ranged from 400° to 600°C with times from 5 minutes to 2 hours [39,40]. Effects of heat treatment on cantilever bending plots for two sizes of a round superelastic nickel-titanium wire are presented in Figure 3 [40].

Figure 3.

Effects of heat treatments on cantilever bending plots for 6 mm test spans of a superelastic nickel-titanium orthodontic wire. Reproduced from [40] with permission.

Loss of superelastic behavior occurs for the 2 hour heat treatment at 600°C, evidenced by the large decrease in springback (difference between the original deflection of 80 degrees and the final angular position on unloading). Heat treatment at 500°C for 10 minutes had minimal effect, while heat treatment for 2 hours caused a decrease in the average superelastic bending moment during the unloading region of clinical importance. Bending properties for nonsuperelastic wires were only slightly affected by these heat treatments. In addition to the use of furnace heat treatment, electrical resistance heat treatment [44] has also been exploited by one manufacturer to produce archwires where the level of biomechanical force varies with position along the wire [24].

Microstructural phases at varying temperatures in nickel-titanium orthodontic wires and their transformations are conveniently studied by differential scanning calorimetry (DSC) [45]. Temperature-modulated DSC provides greater insight into the transformations than conventional DSC [46]. Figures 4 and 5 present temperature-modulated DSC heating curves for shape memory and superelastic nickel-titanium orthodontic wires, respectively. The transformations involving austenite (A), martensite (M) and R-phase (R) are labeled. The austenite-finish (Af) temperature for completion of the transformation from martensite to austenite on heating is determined by the intersection with the adjacent baseline of a tangent line to the peak for the final transformation to austenite [47].

Figure 4.

Heating temperature-modulated DSC plot for a shape memory nickel-titanium orthodontic wire. Reproduced from [46] with permission.

Figure 5.

Heating temperature-modulated DSC plot for a superelastic nickel-titanium orthodontic wire. Reproduced from [46] with permission.

The Af temperature is below body temperature (37°C) for nickel-titanium wires that exhibit shape memory in the oral environment. The superelastic nickel-titanium wires have Af temperatures that are greater than mouth temperature and have more widely separated peaks for the successive transformations from M →R and R → A. The nonsuperelastic wires have much weaker transformations (lower values of enthalpy [ΔH]) and Af temperatures that are also greater than mouth temperature [45]. Examination of x-ray diffraction patterns for nickel-titanium orthodontic wires revealed the effects of heat treatment on the Ms temperature for the start of the cooling transformation to martensite as well as the occurrence of stress relief and perhaps some recrystallization [24,48].

Transformation of a low temperature martensite phase (M′) to the higher temperature form of martensite (M), shown in Figures 5 and 6, is readily detected as a large exothermic peak on the nonreversing heat flow curves from temperature-modulated DSC. Transmission electron microscopy has revealed that this transformation arise from low-temperature twinning within the martensite structure [49].

Advertisement

4. Endodontics

4.1. Stainless steel instruments

Traditionally, endodontic treatment was performed with stainless steel hand files and reamers to remove the injured or diseased dental pulp from the root canals of teeth. While conventional elevated-temperature heat treatment is not recommended for these instruments, they are subjected to sterilization procedures before being using again with a different patient. One study found that dry heat sterilization (180°C for 2 hours) and autoclave sterilization (220 kPa pressure and 136°C for 10 minutes) slightly decreased the flexibility and resistance to torsional fracture of the instruments but they still satisfied the requirements for minimum angular deflection in the ISO standard [50]. Further research is needed to gain insight into the metallurgical origins of the property changes.

4.2. Nickel-titanium instruments

Following the pioneering work of Walia et al that introduced the nickel-titanium hand file to the endodontics profession [51], engine-driven rotary instruments were introduced that enable rapid instrumentation of root canals. These instruments are in widespread clinical use, and research on the nickel-titanium files has been a highly intensive area of research.

The major mechanical property of the equiatomic nickel-titanium alloy that led to replacement of the traditional austenitic stainless steel files was the much lower elastic modulus of NiTi, which enabled curved root canals to be negotiated with facility. An excellent review article [52] describes the manufacturing process for the nickel-titanium files, which are generally machined from starting wire blanks. The conventional nickel-titanium rotary instruments have been fabricated from superelastic nickel-titanium blanks.

Defects caused by the machining process and metallurgical flaws in the starting blanks, along with inadvertent overloading by the clinician, can result in fracture of the file within the root canal, which causes considerable patient anguish since the broken fragments often cannot be easily retrieved [53,54].

A recent study investigated the effect of heat treatment on conventional nickel-titanium rotary instruments, using temperature-modulated DSC and Micro-X-ray diffraction [55]. Results are shown in Figure 6 (a) – (d) for heat treatment at temperatures from 400° to 800°C in a flowing nitrogen atmosphere.

Figure 6.

Temperature-modulated DSC reversing (R), nonreversing (NR) and total (T) heat flow curves for specimens from conventional rotary endodontic instruments after heat treatment in flowing nitrogen for 15 minutes at (a) 400°, (b) 500°, (c) 600° and (d) 850°C. From [55] and reproduced with permission.

Heat treatment between 400° and 600°C increased the Af temperature for as-received conventional NiTi rotary instruments to approximately 45° – 50°C, and the transformations between martensite and austenite were changed to a more reversing character than nonreversing character [55]. Heat treatment in a nitrogen atmosphere might lead to a harder surface from the formation of nitrides [56], which is beneficial for cutting efficiency of the rotary instrument. This research suggested that heat treatment at temperatures near 500°C in a nitrogen atmosphere might yield the optimum microstructure and mechanical properties, with improved resistance to deformation and fracture for conventional NiTi rotary instruments. Heat treatment at temperatures exceeding 600°C should not be performed, since the superelastic behavior is lost along with potential degradation of the wrought microstructure [24]. Another study has reported that heat treatment at 430° and 440°C greatly improved the fatigue resistance of one conventional rotary instrument product [57].

New nickel-titanium rotary instruments have been marketed, for which the wire blanks were improved by special proprietary processing techniques, including heat treatment. The first notable example was M-Wire, named for its stable martensitic structure [58]. Previous conventional rotary instruments were fabricated from superelastic wire blanks with evident transformable austenite detected by conventional DSC [59]. However, when the conventional instruments were cooled far below room temperature to attain the fully martensite condition, the enthalpy changes for transformations from martensite to austenite were far below those for superelastic orthodontic wires [44,45], indicating that these instruments contain a substantial proportion of stable martensite in their microstructures.

Two different batches of M-Wire (termed Type 1 and Type 2), with unknown differences in proprietary processing, were obtained for characterization by temperature-modulated DSC and Micro-X-ray diffraction [58]. Figure 7 shows the differences in the temperature-modulated DSC plots for (a) conventional superelastic wire and (b) Type 1 M-Wire.

Figure 7.

Comparison of temperature-modulated DSC total heat flow for (a) conventional superelastic wire and (b) Type 1 M-Wire. Lower curves are the plots for the heating cycles. Reproduced from [58] with permission.

The general appearances of the temperature-modulated DSC plots in Figure 7 (a) and (b) are similar. However, the approximate Af temperatures for the conventional superelastic wire and Type 1 M-Wire were approximately 15°C and 50°C, respectively. The approximate Af temperature for the Type 2 M-Wire was 45°C. The proportions of the different NiTi phases were quite different for Type 1 and Type 2 M-Wire, as shown in Figure 8.

The Micro-X-ray diffraction pattern indicated that Type 1 M-Wire had an austenitic structure, and the Micro-X-ray diffraction pattern from the conventional superelastic wire was similar. In contrast, the Micro-X-ray diffraction pattern from Type 2 M-Wire contained additional peaks for martensite and R-phase, along with peaks for austenite. However, when M-Wire was examined by transmission electron microscopy, a heavily deformed martensitic structure was found [58]. The explanation is that the DSC peaks only reveal NiTi phases that are capable of undergoing transformation and that (stable) heavily deformed martensitic NiTi only produces weak x-ray diffraction peaks. Rotary instruments fabricated from M-Wire have been found to have similar Af values, microstructures and Vickers hardness, so the machining process and other proprietary fabrication steps do not appear to markedly alter the inherent structure and properties of the starting blanks [60].

Figure 8.

Micro-X-ray diffraction patterns for (a) Type 1 M-Wire and (b) Type 2 M-Wire. Peaks for austenite (A), martensite (M) and R-phase (R) are labeled. Reproduced from [58] with permission.

Recently, new nickel-titanium rotary instruments have been introduced, in which the wire blank is heated to an appropriate temperature for transformation to the R-phase and twisted, along with repeated heat treatment and other subsequent thermal processing; instruments have been characterized by conventional DSC and cantilever bending tests [61]. Another recent study has characterized several new nickel-titanium rotary instruments by DSC and conventional x-ray diffraction, along with optical and scanning electron microscopic examination of their microstructures, including use of energy-dispersive x-ray spectroscopic analyses (SEM/EDS), to investigate the martensitic microstructures and composition of precipitates [62]. Because of the potentially great commercial importance, development of new rotary instruments with improved clinical performance is expected to remain an area of intensive research, along with study of the role of heat treatment [63].

It is essential to appreciate the complexity of the physical metallurgy of the nickel-titanium alloys and the effects of the severe thermomechanical processing of the starting wire blanks, along with heat treatments and machining of the wire blanks, on the metallurgical structure. Transmission electron microscopy and electron diffraction remain the best techniques to gain insight into the instrument microstructures and elucidate the relationships with mechanical properties and clinical performance.

References

  1. 1. AnusaviceK. J.(editorPhillips’ Science of Dental Materials, 11th edition. St. Louis: Saunders/Elsevier Science; 2003
  2. 2. Powers JM, Sakaguchi RL (editors).Craig’s Restorative Dental Materials, 12th edition. St. Louis: Mosby/Elsevier; 2006
  3. 3. Boyer DB, Edie JW.Composition of clinically aged amalgam restorations. Dent Mater 199063146150
  4. 4. International Organization for Standardization: ISO 22674. Dentistry- Metallic materials for fixed and removable appliances; 2006.
  5. 5. WatanabeI.AtsutaM.YasudaK.HisatsuneK.Dimensional changes related to ordering in an AuCu-3wt%Ga alloy at intraoral temperature. Dent Mater 1994106369374
  6. 6. American National Standard/American Dental Association Specification No.Metal-ceramic dental restorative systems: 2000 (Reaffirmed 2010This specification is a modified adoption of ISO 9693:1999.
  7. 7. American Dental Association, Council on Scientific Affairs.Revised classification system for alloys for fixed prosthodontics. [http://www.ada.org/2190.aspx]
  8. 8. RosenstielS. F.LandM. F.FujimotoJ.ContemporaryFixed.Prosthodontics4th.editionMosby.Mosby/Elsevier; 2006
  9. 9. Mackert JR Jr, Ringle RD, Fairhurst CW.High-temperature behavior of a Pd-Ag Alloy for porcelain. J Dent Res 1983621212291235
  10. 10. OhnoH.KanzawaY.Structural changes in the oxidation zones of gold alloys for porcelain bonding containing small amounts of Fe and Sn. J Dent Res 19856416773
  11. 11. PapazoglouE.BrantleyW. A.MitchellJ. C.CaiZ.CarrA. B.Newhigh-palladium.alloysstudies.ofthe.interfacewith.porcelainInt J Prosthodont 199694315322
  12. 12. PapazoglouE.BrantleyW. A.CarrA. B.JohnstonW. M.Porcelain adherence to high-palladium alloys. J Prosthet Dent 1993705386394
  13. 13. Vermilyea SG, Huget EF, Vilca JM.Observations on gold-palladium-silver and gold-palladium alloys. J Prosthet Dent 1980443294299
  14. 14. CarrA. B.CaiZ.BrantleyW. A.MitchellJ. C.New high-palladium casting alloys. Part 2. Effects of heat treatment and burnout temperature. Int J Prosthodont 199363233241
  15. 15. VermilyeaS. G.CaiZ.BrantleyW. A.MitchellJ. C.Metallurgical structure and microhardness of four new palladium-based alloys. J Prosthodont 199654288294
  16. 16. GuoW. H.BrantleyW. A.LiD.ClarkW. A.MonaghanP.HeshmatiR. H.Annealing study palladium-silver dental alloys: Vickers hardness measurements and SEM microstructural observations. J Mater Sci Mater Med 2007181111118
  17. 17. Baran GR.The metallurgy of Ni-Cr alloys for fixed prosthodontics. J Prosthet Dent 1983505639650
  18. 18. MorrisH. F.AsgarK.RoweA. P.CENasjletiThe influence of heat treatments on several types of base-metal removable partial denture alloys. J Prosthet Dent 1979414388395
  19. 19. Bridgeport DA, Brantley WA, Herman PF.Cobalt-chromium and nickel-chromium alloys for removable prosthodontics, Part 1: Mechanical properties. J Prosthodont 199323144150
  20. 20. MDWaldmeierGrasso. J. E.NorburgG. J.MDNowakBend testing of wrought wire removable partial denture alloys. J Prosthet Dent 1996765559565
  21. 21. MJDonachie JrTitanium. A.TechnicalGuide.2ndedition.MaterialsPark. O. H. A. S. M.International; 2000
  22. 22. Jeong Y-H, Choe H-C, Brantley WA.Nanostructured thin film formation on femtosecond laser-textured Ti-35Nb-xZr alloy for biomedical applications. Thin Solid Films 20105191546684675
  23. 23. LeeH.DregiaS.AkbarS.AlhoshanS.Growthof. .TiD.nanowiresO.onTi.Tialloys.byoxidation.J Nanomaterials 2010DOI:10.1155/2010/503186].
  24. 24. BrantleyW. A.EliadesT.(editorsOrthodonticMaterials.ScientificClinicalAspects.Stuttgart: Thieme; 2001
  25. 25. VerstryngeA.Van HumbeeckJ.WillemsG.In-vitro evaluation of the material characteristics of stainless steel and beta-titanium orthodontic wires. Am J Orthod Dentofacial Orthop 20061304460470
  26. 26. GoldbergA. J.VanderbyR.Jr BurstoneC. J.Reduction in the modulus of elasticity in orthodontic wires. J Dent Res 1977561012271231
  27. 27. Asgharnia MK, Brantley WA.Comparison of bending and tension tests for orthodontic wires. Am J Orthod 1986893228236
  28. 28. Khier SE, Brantley WA, Fournelle RA.Structure and mechanical properties of as-received and heat-treated stainless steel orthodontic wires. Am J Orthod Dentofacial Orthop 199199310318
  29. 29. Backofen WA, Gales GF.Heat treating stainless steel wire for orthodontics. Am J Orthod 1951212117124
  30. 30. Funk AC.The heat treatment of stainless steel. Angle Orthod 1951129138
  31. 31. Howe GL, Greener EH, Crimmins DS.Mechanical properties and stress relief of stainless steel orthodontic wire. Angle Orthod 1968383244249
  32. 32. Fillmore GM, Tomlinson JL.Heat treatment of cobalt-chromium alloys of various tempers. Angle Orthod 1979492126130
  33. 33. GoldbergJ.BurstoneC. J.An evaluation of beta titanium alloys for use in orthodontic appliances. J Dent Res 1979582593599
  34. 34. BurstoneC. J.GoldbergA. J.Betatitanium. A.neworthodontic.alloyAm. J.Orthod1980772121132
  35. 35. Andreasen GF, Hilleman TB. An evaluation of 55 cobalt substituted nitinol wire for use in orthodontics. J Am Dent Assoc 197182613731375
  36. 36. AndreasenG. F.BradyP. R. A.usehypothesis.for.nitinolwire.fororthodontics.Angle Orthod 1972422172177
  37. 37. Andreasen GF, Morrow RE.Laboratory and clinical analyses of nitinol wire. Am J Orthod 1978732142151
  38. 38. BurstoneC. J.QinB.MortonJ. Y.ChineseNi.Ti-awire.neworthodontic.alloyAm J Orthod 1985876445452
  39. 39. MiuraF.MogiM.OhuraY.HamanakaH.The super-elastic property of the Japanese NiTi alloy wire for use in orthodontics. Am J Orthod Dentofacial Orthop 1986901110
  40. 40. Khier SE, Brantley WA Fournelle RA.Bending properties of superelastic and nonsuperelastic nickel-titanium orthodontic wires. Am J Orthod Dentofacial Orthop 1991994310318
  41. 41. FletcherM. L.MiyakeS.BrantleyW. A.BMCulbertsonD. S. C.bendingstudies.ofa.newshape-memory.orthodonticwire.J Dent Res 1992Spec Iss A): 169, Abstract 505
  42. 42. DuerigT. W.MeltonK. N.StöckelD.WaymanC. M.(editorsEngineering Aspects of Shape Memory Alloys. London: Butterworth-Heinemann; 1990
  43. 43. KusyR. P. A.reviewof.contemporaryarchwires.Theirproperties.characteristicsAngle Orthod 1997633197207
  44. 44. MiuraF.MogiM.OhuraY.Japanese NiTi alloy wire: Use of the direct electric resistance heat treatment method. Eur J Orthod 1988103187191
  45. 45. Bradley TG, Brantley WA, Culbertson BM.Differential scanning calorimetry (DSC) analyses of superelastic and nonsuperelastic nickel-titanium orthodontic wires. Am J Orthod Dentofacial Orthop 19961096589597
  46. 46. BrantleyW. A.IijimaM.GrentzerT. H.Temperature-modulatedD. S. C.providesnew.insightabout.nickel-titaniumwire.transformationsAm J Orthod Dentofacial Orthop 20031244387394
  47. 47. International Organization for Standardization: ISO 15841. Dentistry- Wires for use in orthodontics;2006ANSI/ADA Specification 32-OrthodonticWires: 2006 is an identical adoption of this ISO standard.
  48. 48. Khier SE. Structural Characterization, Biomechanical Properties, and Potentiodynamic Polarization Behavior of Nickel-Titanium Orthodontic Wire Alloys [Ph.D.Dissertation]. Milwaukee, WI, USA: Marquette University; 1988
  49. 49. BrantleyW. A.GuoW.ClarkW. A.IijimaM.Microstructural studies of 35°C copper Ni-Ti orthodontic wire and TEM confirmation of low-temperature martensite transformation. Dent Mater 2008242204210
  50. 50. Canalda-SahliC.Brau-AguadéE.Sentís-VilaltaJ.The effect of sterilization on bending and torsional properties of K-files manufactured with different metallic alloys. Int Endod J 19983114852
  51. 51. WaliaH.BrantleyW. A.GersteinH.An initial investigation of the bending and torsional properties of Nitinol root canal files. J Endod 198814346351
  52. 52. Thompson SA.An overview of nickel-titanium alloys used in dentistry. Int Endod J 2000334297310
  53. 53. Alapati SB, Brantley WA, Svec TA, PowersJM, Nusstein JM, Daehn GS. SEM observations of nickel-titanium rotary endodontic instruments that fractured during clinical Use. J Endod 20053114043
  54. 54. ParashosP.MesserH. H.Rotary NiTi instrument fracture and its consequences. J Endod 2006321110311043
  55. 55. AlapatiS. B.BrantleyW. A.IijimaM.SchrickerS. R.NussteinJ. M.Li-MU.SvecT. A.Micro-XR. D.temperature-modulatedD. S. C.investigationof.nickel-titaniumrotary.endodonticinstruments.Dent Mater 2009251012211229
  56. 56. TripiT. R.BonaccorsoA.RapisardaE.TripiV.CondorelliG. G.MarinoR.FragalàI.Depositions of nitrogen on NiTi instruments. J Endod 2002287497500
  57. 57. ZinelisS.DarabaraM.TakaseT.OganeK.PapadimitriouG. D.The effect of thermal treatment on the resistance of nickel-titanium rotary files in cyclic fatigue. Oral Surg Oral Med Oral Pathol Oral Radiol Endod 20071036843847
  58. 58. AlapatiS. B.BrantleyW. A.IijimaM.ClarkW. A.KovarikL.BuieC.LiuJ.JohnsonW. B.Metallurgical characterization of a new nickel-titanium wire for rotary endodontic instruments. J Endod 2009351115891593
  59. 59. BrantleyW. A.SvecT. A.IijimaM.PowersJ. M.GrentzerT. H.Differential scanning calorimetric studies of nickel titanium rotary endodontic instruments. J Endod 2002288567572
  60. 60. LiuJ.Characterization of New Rotary Endodontic Instruments Fabricated from Special Thermomechanically Processed NiTi Wire [Ph.D. Dissertation]. Columbus, OH, USA: Ohio State University; 2009
  61. 61. HouX.YahataY.HayashiY.EbiharaA.HanawaT.SudaH.Phase transformation behaviour and bending property of twisted nickel-titanium endodontic instruments. Int Endod J 2011443253258
  62. 62. ShenY.ZhouH. M.ZhengY. F.CampbellL.PengB.HaapasaloM.Metallurgical characterization of controlled memory wire nickel-titanium rotary instruments. J Endod 2011371115661571
  63. 63. YahataY.YoneyamaT.HayashiY.EbiharaA.DoiH.HanawaT.SudaH.Effect of heat treatment on transformation temperatures and bending properties of nickel-titanium endodontic instruments. Int Endod J 2009427621626
  64. 64. OtsukaK.RenX.Physical metallurgy of Ti-Ni-based shape memory alloys. Prog Mater Sci 2005505511678

Written By

William Brantley and Satish B. Alapati

Submitted: 02 December 2011 Published: 19 September 2012