Skip to main content

MINI REVIEW article

Front. Dent. Med, 28 October 2021
Sec. Dental Materials
This article is part of the Research Topic Debates in Endodontics: 2023/2024 View all 3 articles

Heat Treatment and Surface Treatment of Nickel–Titanium Endodontic Instruments

  • 1Department of Conservative Dentistry, School of Dentistry, Dental and Life Science Institute, Dental Research Institute, Pusan National University, Yangsan, South Korea
  • 2Faculty of Dentistry, Division of Endodontics, Department of Oral Biological and Medical Sciences, University of British Columbia, Vancouver, BC, Canada
  • 3Department of Stomatology, Affiliated Hospital of Jining Medical University, Jining, China

Knowledge and thorough understanding of the characteristics of endodontic nickel–titanium (NiTi) files is paramount for dentists performing root canal treatments to patients. Understanding the behavior of the NiTi files guides the clinicians in choosing the correct instruments for different clinical and anatomical situations. This review focuses on the metallurgical properties of endodontic NiTi files, with a special emphasis on recent developments and improvements in metallurgy and the effects of heat treatment and surface treatment. In this study, the impact that such developments have on the properties of endodontic NiTi files is discussed.

Introduction

Since Walia and colleagues first introduced nickel–titanium (NiTi) instruments in the late 1980s, NiTi instruments have revolutionized the root canal instrumentation by reducing the majority of iatrogenic instrumentation issues commonly associated with stainless steel files such as zipping, ledges, transportation, and perforation (1, 2).

The first NiTi rotary instruments were marketed in the 1990s (3). Despite significant advancements in file design and manufacturing procedures for NiTi rotary instruments during the last two decades, fracture of rotary instruments induced by torsional or cyclic fatigue remains a concern for clinicians, particularly in calcified or severely curved root canals (46). The relative proportions and properties of the microstructural phases govern the mechanical behavior of NiTi alloy. Heat treatment (thermal processing) has been reported to influence the fatigue resistance of NiTi instruments and is one of the most common methods for adjusting NiTi alloy transition temperatures (710).

Nickel–titanium alloy has found a unique commercial application in the endodontic industry, because of its shape memory effect and corrosion resistance results from phase transformation. Novel NiTi instruments produced by using thermomechanical techniques, such as M-wire, R-phase, and controlled memory (CM) files, have been launched in recent years and shown to have enhanced flexibility and cyclic fatigue resistance when compared to conventional superelastic NiTi files (1113). New NiTi instrument systems with a titanium oxide surface layer [e.g., WaveOne Gold (Dentsply Sirona, York, Pennsylvania, United States) and Reciproc Blue (VDW, Munich, Germany)] are made from NiTi alloy heat-treated in a special way.

In recent years, many new NiTi instruments have been introduced, and understanding the nature of NiTi alloy depending on their phases and their effects on instrument performance is important for clinicians for attaining favorable clinical results.

Properties of Each Phase (Austenitic, Martensitic, and R-Phase)

The NiTi alloys of the endodontic instruments are made of ~56% nickel and 44% titanium by weight, or a 1:1 atomic ratio (equiatomic) (14). Similar to other metallic systems, NiTi alloys can come in a variety of crystallographic forms. Near-equiatomic NiTi alloys have three microstructural phases (austenite, martensite, and R-phase), and their properties and their respective proportions influence the mechanical properties of the metal (15).

The austenite phase with the B2 cubic crystal structure exists at higher temperatures and is stronger (~80–90 GPa) and stiffer than the martensite (14, 15), while the martensite phase is a low-temperature monoclinic phase (B19′) with a lower Young's modulus and yield strength (~30–40 GPa) than the austenite phase (16, 17). This demonstrated that the martensite can be easily deformed at low stress, while the austenite has a substantially higher yield and flow stresses. The martensite phase also supports reducing the risk of file fracture under high stress conditions since it can be deformed rather than fractured. Therefore, the abundant effort has been dedicated to the introduction of martensitic alloys such as M-wire and CM wire instruments into the NiTi instruments market. Various studies investigated the performance of the instruments made from M-wire and CM NiTi and reported enhanced flexibility and fatigue resistance than those of conventional NiTi instruments (12, 13, 18).

The phase transformation from martensite to austenite and austenite to martensite can occur in one or two steps, with the two-stage transformation involving the formation of an intermediate R-phase (19). The R-phase is a “Rhombohedral phase” that differs from the cubic B2 phase in the austenite phase (20). The R-phase transformation occurs before the B2-B19′ transition and shows thermoelastic martensitic transformation features. The R-phase transformation can be induced by both temperature and stress. The recoverable strain of the R-phase/austenite transformation (~0.5%) is smaller than that of the martensite/austenite transformation in NiTi alloy, and the temperature hysteresis is exceedingly modest (19). Furthermore, the R-phase/austenite transformation has remarkable cyclic stability and Young's modulus of R-phase is lower than that of the austenite (21).

The Phase Transformation of NiTi Alloy

Conventional superelastic NiTi alloys present in the austenite form at room temperature. As austenite cools down, it begins to transform into martensite at the martensite transformation start temperature (Ms) and completes the transition at the martensite transformation finish temperature (Mf). On the other hand, when martensitic NiTi is heated past the austenite transformation start temperature (As), the crystal structure of the NiTi begins to transition to austenite, and once heated past the higher austenite finish (Af) temperature, the NiTi crystal structure becomes entirely austenite (Figure 1).

FIGURE 1
www.frontiersin.org

Figure 1. Differential scanning calorimetry curves of nickel–titanium (NiTi) instruments. The austenite-finish (Af) temperature shown with blue dotted line is 17°C for a “conventional” superelastic NiTi file (A) and 55°C for heat-treated NiTi file shown with a red dotted line (B). The blue bar represents the temperature range between room temperature and body temperature.

The phase transformation from austenite to martensite can also be caused by stress or external force, which can accommodate greater stress without increasing strain (14, 22). Compared to stainless steel, the superelasticity of NiTi allows the total recovery of the deformations up to 8% of the deformation (14). As a result, a conventional NiTi instrument, in an austenitic state at body temperature, exhibits transformational elasticity or the capacity to return to its original shape after being distorted. When external stress such as torsional stress or file friction against canal walls is applied, the stress-induced martensitic transformation occurs, resulting in more resilient materials with a higher ultimate tensile strength (23). Because the stress-induced martensitic state is not stable at room temperature, once the stress is relieved, the deformed NiTi alloy immediately reverts to the austenite phase. The NiTi files can shape the root canal with a constant cutting force in this manner, even in a curved root canal. When the NiTi alloy is deformed in the martensite state by an external force, it can be also fully recovered when heated. Deformation via martensite reorientation can be observed at temperatures below As, the starting temperature which is important for the reverse transformation of martensite upon heating and is completed at Af (19).

Superelasticity and Shape Memory

Nickel–titanium alloys show unique superelasticity and shape memory properties (14, 24). When the ambient temperature is higher than the Af temperature of the NiTi alloy, superelasticity is interrelated to the occurrence of a phase transition of the alloy when stress is applied above a critical level. When the stress is relieved, the stress-induced martensitic change reverses spontaneously and the material returns to its previous shape and size (24). In other words, when the endodontic instrument is removed from the root canal, it reverts to its original shape (25). The enhanced flexibility of NiTi instruments over stainless steel instruments is due to this reversible thermoelastic martensitic transition, which makes the instrumentation of curved root canals easier and safer (14). Superelasticity occurs when austenite and martensite undergo a reversible phase change. As a result, transformation temperatures have a significant influence on the mechanical characteristics and behavior of NiTi, which can be influenced by minor compositional changes, impurities, and heat treatments during the manufacturing process (26).

Heat Treatment of NiTi Alloys

The goal of heat treatments is to change the transition temperatures of NiTi alloys and, as a result, modify fatigue resistance. Superelastic conventional NiTi instruments existed in the austenite phase at room and body temperatures, which limited their usage in severely curved canals due to the stiffness of the instrument and low fatigue resistance (7, 27). The heat treatment process releases the internal strain of NiTi alloy and increases the phase transformation temperature of NiTi, resulting in more martensite phase at clinically relevant temperatures (7, 10), which makes heat-treated NiTi instruments higher flexibility and fatigue resistance than those of conventional NiTi instruments (Figure 2).

FIGURE 2
www.frontiersin.org

Figure 2. The photomicrograph of the fracture surface of NiTi instruments with the region of crack propagation and dimple area outlined (dotted line).

In early 2000, a new method for optimizing the structure of NiTi wire blanks for rotary instruments has been developed. Several proprietary thermomechanical processing techniques have been established with the goal of creating superelastic NiTi wire blanks that contain the significantly stable martensite phase in clinical conditions. M-wire (Dentsply Tulsa Dental Specialties) was introduced in 2007. It was developed by applying a series of heat treatments processes and contains three phases: martensite, R-phase, and austenite (11). M-wire instruments include Dentsply's ProFile GT Series X, ProFile Vortex, ProTaper Next, and WaveOne. In 2008, a new manufacturing process was developed by SybronEndo: Twisted Files (TF). TF is manufactured by twisting the NiTi rod, while most NiTi files are manufactured by the grinding method. The manufacturer claims that TF instruments were created by thermally transforming a raw NiTi wire in the austenite phase into the R-phase. R-phase occurs within a very narrow temperature range on the heating or cooling curve between martensitic and austenitic forms and made it possible to twist the NiTi rod. Previous studies reported that TF has greater cyclic fatigue resistance than files that have been manufactured by grinding, while the torsional resistance of R-phase files was significantly lower than that of ground files (2830).

In 2010, CM wire (DS Dental, Johnson City, TN, United States) was introduced as a new NiTi alloy with high flexible properties. CM NiTi files are made by a specific thermomechanical technique that controls the memory feature of NiTi alloy, making them exceptionally flexible but lacking the shape memory which is seen in other superelastic NiTi files. In other words, CM NiTi files do not rebound after unloading, and their original shape is restored following the application of heat. Thermally treated CM alloys would be primarily or entirely in the martensite phase at body temperature because the Af temperature of CM wire is ~55 and 50°C (7–9, 102). CM NiTi files include HyFlex CM and EDM (Coltène/Whaledent, Altstätten, Switzerland), Typhoon Infinite Flex NiTi Files (Clinician's Choice Dental Products, New Milford, CT, United States), and VTaper 2H (SS White, Lakewood, NJ, United States).

Thermal treatments have been reported to influence the mechanical properties and transformation features of NiTi alloys based on their thermomechanical history (31). Heat treatment would be applied prior to machining the instrument to reduce the work hardening of the alloy (32, 33). Recently, the application of this heating process also has been applied after the machining of the files, with the aim of transforming the alloy into a slightly different crystalline phase structure with enhanced mechanical properties (improved flexibility with superior mechanical resistance) (33). Post-machining heat treatment is applied to K3XF (SybronEndo, Orange, CA, United States) instrument. K3XF showed similar torsional properties, but greater flexibility and resistance to cyclic fatigue than those of the original K3 instrument (33, 34).

In comparison to conventional superelastic NiTi rotary instruments, heat-treated NiTi instruments have greater flexibility and cyclic fatigue resistance (7). It can be assumed that the cutting efficiency of the comparatively soft and flexible NiTi instruments is lower than that of the relatively stiff NiTi instruments. However, two investigations (35, 36) indicated that a heat-treated NiTi instrument (HyFlex CM) showed the most efficient cutting instrument in lateral action against dentine and acrylic resin when compared to other coronal flaring instruments such as BioRace (FKG Dentsaire SA, La Chaux-de-Fonds, Switzerland), ProFile (Dentsply Maillefer, Ballaigues, Switzerland), and ProTaper (Dentsply Maillefer).

Surface Treatment of NiTi Alloys

Surface treatment of NiTi instruments reduces inherent defects, increases surface hardness and flexibility, and improves fatigue resistance and cutting efficiency (3739). Microcracks are frequently formed on the surface of the instrument, indicating the very first stage of the fatigue phenomenon (40). Thus, a treatment that improves surface smoothness is expected to inhibit crack initiation and increase fatigue resistance. Electropolishing refers to any electrochemical procedure that aims to reduce the surface irregularities of material and achieve a high gloss finishing. It is carried out by immersing the part in a specially formulated, usually acidic, electrolyte solution and passing a direct electric current to facilitate a selective dissolution of the material (39). RaCe (FKG Dentaire) and EndoSequence (Brasseler, Savannah, GA, United States) NiTi file systems have undergone the electropolishing process. Previous studies (41, 42) have demonstrated that electropolishing improves the fatigue resistance of NiTi instruments, while some other studies have shown that the benefits of electropolishing may vary depending on the instrument type, design, and cross-sectional area (43, 44).

The surface hardness and wear resistance of heat-treated NiTi instruments have been reported to be improved using surface engineering techniques. Physical vapor deposition describes a variety of vacuum deposition methods that can be used to produce thin films and coatings. Several manufacturers have devised thermomechanical processing sequences to generate a titanium oxide surface layer for the NiTi instrument. Gao et al. found that the comparatively hard titanium oxide surface layer of the Vortex Blue (Dentsply Tulsa Dental, Tulsa, OK, United States) instrument may compensate for the loss of hardness when compared to ProFile Vortex M-wire while enhancing cutting efficiency and wear resistance (45). HyFlex electrical discharge machining (EDM) is manufactured via EDM, a non-contact thermal erosion process that partially melts and evaporates the wire by high-frequency spark discharges and shows higher resistance to cyclic fatigue than HyFlex CM (46).

Future Prospects

Endodontic hand- and engine-driven NiTi files have been available for clinicians already for almost 30 years. Continuous development has taken place since the introduction of the first NiTi files. It is highly likely that this path of incremental improvements will continue in the foreseeable future. However, there is currently no specification or international standard for assessing the fracture resistance of endodontic rotary instruments. Despite the fact that the cyclic fatigue and torsional test do not accurately represent clinical settings, it is required for evaluating the mechanical properties of endodontic instruments. By modifying the microstructure of the NiTi alloy through heat treatment, the mechanical properties of the alloy can be enhanced. Because the heat-treated files have a higher resistance to cyclic fatigue as well as an increase in ductility, the incidence of file fracture during clinical use might be reduced (46, 47). It is assumed that the higher ductility assessed by angular distortion gives the heat-treated instrument a higher “safety factor,” because files with more observable distortion of the cutting spirals are more likely to be discarded before breakage (48). However, a decrease in cutting efficiency has been reported in the heat-treated instruments (49). One of the potentially interesting recent observations is the effect of cold on the fatigue resistance of NiTi files (50). Whether this result can be transferred to the clinical situation of the root canal, remains to be seen. As rotary NiTi files fracture mainly due to limited fatigue resistance, it can be assumed that success in improving this characteristic of the NiTi files will be a key factor in future improvements of NiTi files in the instrumentation of the most difficult anatomies.

Conclusions

The changes in transformation behavior as a result of heat treatment have been found to affect the mechanical characteristics, enhancing clinical performance compared with files of similar design and size made from conventional NiTi alloy. Heat-treated and CM NiTi instruments are frequently employed by clinicians for endodontic treatment nowadays. Although the details of the thermomechanical treatment history of the new NiTi wires are unknown until now, it appears that thermomechanical processing is a very promising method for improving the efficiency and safety of contemporary endodontic instruments. However, it is important to remember that all instruments have strengths and weaknesses and that properties are determined by a variety of factors such as alloy type and degree of taper and cross-sectional design.

Author Contributions

SK contributed to the drafting and the critical revision of the manuscript. MH contributed to the conception, designing, and the critical revision of the manuscript. HCK contributed to the critical revision of the manuscript. ZW and HL contributed to the drawing of the figures and the critical revision of the manuscript. YS contributed to the conception, designing, drafting, and drawing of the figures as well as the critical revision of the manuscript. All authors contributed to the article and approved the submitted version.

Conflict of Interest

The authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Publisher's Note

All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article, or claim that may be made by its manufacturer, is not guaranteed or endorsed by the publisher.

References

1. Walia H, Brantley WA, Gerstein H. An initial investigation of the bending and torsional properties of nitinol root-canal files. J Endod. (1988) 14:346–51. doi: 10.1016/S0099-2399(88)80196-1

PubMed Abstract | CrossRef Full Text | Google Scholar

2. Peters OA, Barbakow F, Peters CI. An analysis of endodontic treatment with three nickel-titanium rotary root canal preparation techniques. Int Endod J. (2004) 37:849–59. doi: 10.1111/j.1365-2591.2004.00882.x

PubMed Abstract | CrossRef Full Text | Google Scholar

3. Haapasalo M, Shen Y. Evolution of nickel – titanium instruments: from past to future. Endod Topics. (2013) 29:3–17. doi: 10.1111/etp.12049

CrossRef Full Text | Google Scholar

4. Sattapan B, Nervo GJ, Palamara JEA, Messer HH. Defects in rotary nickel-titanium files after clinical use. J Endod. (2000) 26:161–5. doi: 10.1097/00004770-200003000-00008

PubMed Abstract | CrossRef Full Text | Google Scholar

5. Cheung GS, Peng B, Bian Z, Shen Y, Darvell BW. Defects in ProTaper S1 instruments after clinical use: fractographic examination. Int Endod J. (2005) 38:802–9. doi: 10.1111/j.1365-2591.2005.01020.x

PubMed Abstract | CrossRef Full Text | Google Scholar

6. Shen Y, Cheung GS, Peng B, Haapasalo M. Defects in nickel-titanium instruments after clinical use. Part 2: fractographic analysis of fractured surface in a cohort study. J Endod. (2009) 35:133–6. doi: 10.1016/j.joen.2008.10.013

PubMed Abstract | CrossRef Full Text | Google Scholar

7. Shen Y, Zhou HM, Zheng YF, Campbell L, Peng B, Haapasalo M. Metallurgical characterization of controlled memory wire nickel-titanium rotary instruments. J Endod. (2011) 37:1566–71. doi: 10.1016/j.joen.2011.08.005

PubMed Abstract | CrossRef Full Text | Google Scholar

8. Zhou H, Shen Y, Zheng W, Li L, Zheng Y, Haapasalo M. Mechanical properties of controlled memory and superelastic NiTi wires used in the manufacture of rotary endodontic instruments. J Endod. (2012) 38:1535–40. doi: 10.1016/j.joen.2012.07.006

PubMed Abstract | CrossRef Full Text | Google Scholar

9. Hou XM, Yahata Y, Hayashi Y, Ebihara A, Hanawa T, Suda H. Phase transformation behaviour and bending property of twisted nickel-titanium endodontic instruments. Int Endod J. (2011) 44:253–8. doi: 10.1111/j.1365-2591.2010.01818.x

PubMed Abstract | CrossRef Full Text | Google Scholar

10. Hieawy A, Haapasalo M, Zhou H, Wang Z, Shen Y. Phase transformation behavior and resistance to bending and cyclic fatigue of ProTaper Gold and ProTaper Universal instruments. J Endod. (2015) 41:1134–8. doi: 10.1016/j.joen.2015.02.030

PubMed Abstract | CrossRef Full Text | Google Scholar

11. Ye J, Gao Y. Metallurgical characterization of M-wire nickel-titanium shape memory alloy used for endodontic rotary instruments during low-cycle fatigue. J Endod. (2012) 38:105–7. doi: 10.1016/j.joen.2011.09.028

PubMed Abstract | CrossRef Full Text | Google Scholar

12. Shen Y, Qian W, Abtin H, Gao Y, Haapasalo M. Fatigue testing of controlled memory wire nickel–titanium rotary instruments. J Endod. (2011) 37:997–1001. doi: 10.1016/j.joen.2011.03.023

PubMed Abstract | CrossRef Full Text | Google Scholar

13. Peters OA, Gluskin AK, Weiss RA, Han JT. An in vitro assessment of the physical properties of novel Hyflex nickel–titanium rotary instruments. Int Endod J. (2012) 45:1027–34. doi: 10.1111/j.1365-2591.2012.02067.x

PubMed Abstract | CrossRef Full Text | Google Scholar

14. Thompson SA. An overview of nickel–titanium alloys used in dentistry. Int Endod J. (2000) 33:297–310. doi: 10.1046/j.1365-2591.2000.00339.x

PubMed Abstract | CrossRef Full Text | Google Scholar

15. Buehler WJ, Wang FE. A summary of recent research on the Nitinol alloys and their potential application in ocean engineering. Ocean Eng. (1967) 1:105–20. doi: 10.1016/0029-8018(68)90019-X

CrossRef Full Text | Google Scholar

16. Zadno R, Duerig TW. Linear superelasticity in cold worked NiTi. In: Duerig TW, Melton KN, editors. Engineering Aspects of Shape Memory Alloys. London: Butterworth-Heinemann (1990). p. 414–9. doi: 10.1016/B978-0-7506-1009-4.50038-X

CrossRef Full Text

17. Hodson DE, Wu MH. Shape Memory Alloys. Metals Handbook. 10th ed. Materials Park, OH: ASM International (1990). p. 897–902. doi: 10.31399/asm.hb.v02.a0001100

CrossRef Full Text

18. Kim HC, Kwak SW, Cheung GS, Ko DH, Chung SM, Lee W. Cyclic fatigue and torsional resistance of two new nickel-titanium instruments used in reciprocation motion: reciproc versus WaveOne. J Endod. (2012) 38:541–4. doi: 10.1016/j.joen.2011.11.014

PubMed Abstract | CrossRef Full Text | Google Scholar

19. Otsuka K, Ren X. Physical metallurgy of Ti – Ni-based shape memory alloys. Prog Mater Sci. (2005) 50:511–678. doi: 10.1016/j.pmatsci.2004.10.001

CrossRef Full Text | Google Scholar

20. Zhang XY, Sehitoglu H. Crystallography of the B2→ R→ B19′ phase transformations in NiTi. Mater Sci Eng A. (2004) 374:292–302. doi: 10.1016/j.msea.2004.03.013

CrossRef Full Text | Google Scholar

21. Tobushi H, Yamada S, Hachisuka T, Ikai A, Tanaka K. Cyclic deformation properties in a wire of TiNi shape memory alloy subjected to cyclic loadings. In: Abe T, Tsuta T, editors. Advances in Engineering Plasticity and Its Applications. Oxford: Pergamon Press (1996). p. 611–6. doi: 10.1016/B978-0-08-042824-6.50100-8

CrossRef Full Text

22. Zhou H, Peng B, Zheng Y. An overview of the mechanical properties of nickel – titanium endodontic instruments. Endod Top. (2013) 29:42–54. doi: 10.1111/etp.12045

CrossRef Full Text | Google Scholar

23. Torrisi L. The NiTi superelastic alloy application to the dentistry field. Biomed Mater Eng. (1999) 9:39–47.

PubMed Abstract | Google Scholar

24. Saburi T. Ti-Ni shape memory alloys. In: Otsuka K, Wayman CM, editors. Shape Memory Materials. Cambridge, NY: Cambridge University Press (1998). p. 49–96.

Google Scholar

25. Viana AC, Chaves Craveiro de Melo M, Guiomar de Azevedo Bahia M, Lopes Buono VT. Relationship between flexibility and physical, chemical, and geometric characteristics of rotary nickel-titanium instruments. Oral Surg Oral Med Oral Pathol Oral Radiol Endod. (2010) 110:527–33. doi: 10.1016/j.tripleo.2010.05.006

PubMed Abstract | CrossRef Full Text | Google Scholar

26. Yoneyama T, Kobayashi C. Endodontic instruments for root canal treatment using Ti-Ni shape memory alloys. In: Yoneyama T, Miyazaki S, editors. Shape Memory Alloys for Biomedical Applications. Cambridge: Woodhead Pub (2009). p. 297–305. doi: 10.1533/9781845695248.2.297

CrossRef Full Text | Google Scholar

27. Brantley WA, Svec TA, Iijima M, Powers JM, Grentzer TH. Differential scanning calorimetric studies of nickel titanium rotary endodontic instruments. J Endod. (2002) 28:567–72. doi: 10.1097/00004770-200208000-00001

PubMed Abstract | CrossRef Full Text | Google Scholar

28. Larsen CM, Watanabe I, Glickman GN, He J. Cyclic fatigue analysis of a new generation of nickel–titanium rotary instruments. J Endod. (2009) 35:401–3. doi: 10.1016/j.joen.2008.12.010

PubMed Abstract | CrossRef Full Text | Google Scholar

29. Kim HC, Yum J, Hur B, Cheung GS. Cyclic fatigue and fracture characteristics of ground and twisted nickel–titanium rotary files. J Endod. (2010) 36:147–52. doi: 10.1016/j.joen.2009.09.037

PubMed Abstract | CrossRef Full Text | Google Scholar

30. Park SY, Cheung GS, Yum J, Hur B, Park JK, Kim HC. Dynamic torsional resistance of nickel-titanium rotary instruments. J Endod. (2010) 36:1200–4. doi: 10.1016/j.joen.2010.02.016

PubMed Abstract | CrossRef Full Text | Google Scholar

31. Kuhn G, Jordan L. Fatigue and mechanical properties of nickel-titanium endodontic instruments. J Endod. (2002) 28:716–20. doi: 10.1097/00004770-200210000-00009

PubMed Abstract | CrossRef Full Text | Google Scholar

32. Kuhn G, Tavernier B, Jordan L. Influence of structure on nickel-titanium endodontic instrument failure. J Endod. (2001) 27:516–20. doi: 10.1097/00004770-200108000-00005

PubMed Abstract | CrossRef Full Text | Google Scholar

33. Gambarini G, Plotino G, Grande NM, Al-Sudani D, De Luca M, Testarelli L. Mechanical properties of nickel-titanium rotary instruments produced with a new manufacturing technique. Int Endod J. (2011) 44:337–41. doi: 10.1111/j.1365-2591.2010.01835.x

PubMed Abstract | CrossRef Full Text | Google Scholar

34. Shen Y, Zhou HM, Wang Z, Campbell L, Zheng Y, Haapasalo M. Phase transformation behavior and mechanical properties of thermomechanically treated K3XF nickel-titanium instruments. J Endod. (2013) 39:919–23. doi: 10.1016/j.joen.2013.04.004

PubMed Abstract | CrossRef Full Text | Google Scholar

35. Morgental RD, Vier-Pelisser FV, Kopper PM, de Figueiredo JA, Peters OA. Cutting efficiency of conventional and martensitic nickel-titanium instruments for coronal flaring. J Endod. (2013) 39:1634–8. doi: 10.1016/j.joen.2013.08.016

PubMed Abstract | CrossRef Full Text | Google Scholar

36. Peters OA, Morgental RD, Schulze KA, Paque F, Kopper PM, Vier-Pelisser FV. Determining cutting efficiency of nickel-titanium coronal flaring instruments used in lateral action. Int Endod J. (2014) 47:505–13. doi: 10.1111/iej.12177

PubMed Abstract | CrossRef Full Text | Google Scholar

37. Rapisarda E, Bonaccorso A, Tripi TR, Fragalk I, Condorelli GG. The effect of surface treatments of nickel-titanium files on wear and cutting efficiency. Oral Surg Oral Med Oral Pathol Oral Radiol Endod. (2000) 89:363–8. doi: 10.1016/S1079-2104(00)70103-X

PubMed Abstract | CrossRef Full Text | Google Scholar

38. Tripi RT, Bonaccorso A, Condorelli G. Fabrication of hard coatings on NiTi instruments. J Endod. (2003) 29:132–4. doi: 10.1097/00004770-200302000-00011

PubMed Abstract | CrossRef Full Text | Google Scholar

39. Gutmann JL, Gao Y. Alteration in the inherent metallic and surface properties of nickel-titanium root canal instruments to enhance performance, durability and safety: a focused review. Int Endod J. (2012) 45:113–28 doi: 10.1111/j.1365-2591.2011.01957.x

PubMed Abstract | CrossRef Full Text | Google Scholar

40. Schijve J. Fatigue of Structures and Materials. Dordrecht: Kluwer Academic Publisher (2001). p. 513.

Google Scholar

41. Condorelli GG, Bonaccorso A, Smecca E, Schäfer E, Cantatore G, Tripi TR. Improvement of the fatigue resistance of NiTi endodontic files by surface and bulk modifications. Int Endod J. (2010) 43:866–73. doi: 10.1111/j.1365-2591.2010.01759.x

PubMed Abstract | CrossRef Full Text | Google Scholar

42. Kwak SW, Lee JY, Goo HJ, Kim HC. Effect of surface treatment on the mechanical properties of nickel-titanium files with a similar cross-section. Restor Dent Endod. (2017) 42:216–23. doi: 10.5395/rde.2017.42.3.216

PubMed Abstract | CrossRef Full Text | Google Scholar

43. Cheung GSP, Shen Y, Darvell BW. Effect of environment on low-cycle fatigue of a nickel-titanium instrument. J Endod. (2007) 33:1433–7. doi: 10.1016/j.joen.2007.08.007

PubMed Abstract | CrossRef Full Text | Google Scholar

44. Oh SR, Chang SW, Lee Y, Gu Y, Son WJ, Lee W, et al. A comparison of nickeltitanium rotary instruments manufactured using different methods and cross-sectional areas: ability to resist cyclic fatigue. Oral Surg Oral Med Oral Pathol Oral Radiol Endod. (2010) 109:622–8. doi: 10.1016/j.tripleo.2009.12.025

PubMed Abstract | CrossRef Full Text | Google Scholar

45. Gao Y, Gutmann JL, Wilkinson K, Maxwell R, Ammon D. Evaluation of the impact of raw materials on the fatigue and mechanical properties of ProFile Vortex rotary instruments. J Endod. (2012) 38:398–401. doi: 10.1016/j.joen.2011.11.004

PubMed Abstract | CrossRef Full Text | Google Scholar

46. Goo HJ, Kwak SW, Ha JH, Pedullà E, Kim HC. Mechanical properties of various heat-treated nickel-titanium rotary instruments. J Endod. (2017) 43:1872–7. doi: 10.1016/j.joen.2017.05.025

PubMed Abstract | CrossRef Full Text | Google Scholar

47. Shen Y, Coil JM, Zhou H, Zheng Y, Haapasalo M. HyFlex nickel-titanium rotary instruments after clinical use: metallurgical properties. Int Endod J. (2013) 46:720–9. doi: 10.1111/iej.12049

PubMed Abstract | CrossRef Full Text | Google Scholar

48. Seto BG, Nicholls JI, Harrington GW. Torsional properties of twisted and machined endodontic files. J Endod. (1990) 16:355–60. doi: 10.1016/S0099-2399(06)81905-9

PubMed Abstract | CrossRef Full Text | Google Scholar

49. Chi CW, Lai EH, Liu CY, Lin CP, Shin CS. Influence of heat treatment on cyclic fatigue and cutting efficiency of ProTaper Universal F2 instruments. J Dent Sci. (2017) 12:21–6. doi: 10.1016/j.jds.2016.06.001

PubMed Abstract | CrossRef Full Text | Google Scholar

50. Shen Y, Huang X, Wang Z, Wei X, Haapasalo M. Low environmental temperature influences the fatigue resistance of nickel-titanium files. J Endod. (2018) 44:626–9. doi: 10.1016/j.joen.2017.11.004

PubMed Abstract | CrossRef Full Text | Google Scholar

Keywords: heat treatment, fatigue failure, nickel–titanium instrument, surface treatment, endodontics

Citation: Kwak SW, Shen Y, Liu H, Wang Z, Kim HC and Haapasalo M (2021) Heat Treatment and Surface Treatment of Nickel–Titanium Endodontic Instruments. Front. Dent. Med. 2:769977. doi: 10.3389/fdmed.2021.769977

Received: 03 September 2021; Accepted: 30 September 2021;
Published: 28 October 2021.

Edited by:

Vesna Miletic, The University of Sydney, Australia

Reviewed by:

Vladimir Ivanovic, University of Kragujevac, Serbia
Sebastian Bürklein, University Hospital Münster, Germany

Copyright © 2021 Kwak, Shen, Liu, Wang, Kim and Haapasalo. This is an open-access article distributed under the terms of the Creative Commons Attribution License (CC BY). The use, distribution or reproduction in other forums is permitted, provided the original author(s) and the copyright owner(s) are credited and that the original publication in this journal is cited, in accordance with accepted academic practice. No use, distribution or reproduction is permitted which does not comply with these terms.

*Correspondence: Markus Haapasalo, markush@dentistry.ubc.ca

Disclaimer: All claims expressed in this article are solely those of the authors and do not necessarily represent those of their affiliated organizations, or those of the publisher, the editors and the reviewers. Any product that may be evaluated in this article or claim that may be made by its manufacturer is not guaranteed or endorsed by the publisher.