Effect of Bonding Agent Application Method on Titanium-Ceramic Bond Strength Joseph G. Curtis, DMD,1 Jon Dossett, DDS,2 Thomas J. Prihoda, PhD,3 & Erica C. Teixeira, DDS, MSc, PhD4 1

Department of Prosthodontics, 579th Dental Squadron, United States Air Force, Washington, DC United States Air Force Dental Evaluation and Consultation Service, San Antonio, TX 3 Department of Pathology, University of Texas Health Science Center at San Antonio, San Antonio, TX 4 Department of Operative Dentistry, The University of Iowa, College of Dentistry and Dental Clinics, Iowa City, IA 2

Keywords Analysis of variance; dental porcelain; materials testing; metal ceramic alloys; microscopy, electron, scanning; titanium. Correspondence E. C. Teixeira, Department of Operative Dentistry, College of Dentistry, The University of Iowa, 229 Dental Science Building S, Iowa City, IA 52242-1001. E-mail: [email protected]. Presented as a poster presentation at the AADR/IADR 2013 annual meeting, Seattle, WA. The views expressed in this publication are those of the author and do not reflect the official policy of the Department of Defense, or other departments of the U.S. government. The author does not have any financial interest in the companies whose materials are discussed in this publication. Accepted May 7, 2014

Abstract Purpose: Although milled titanium may be used as a substructure in fixed and implant prosthodontics, the application of the veneering porcelain presents particular challenges compared to traditional alloys. To address these challenges, some Ti ceramic systems incorporate the application of a bonding agent prior to the opaque layer. Vita Titankeramik’s bonding agent is available as a powder, paste, and spray-on formulation. We examined the effect of these three application methods on the bond strength. Materials and Methods: Four titanium bars were milled from each of 11 wafers cut from grade II Ti using the Kavo Everest milling unit and a custom-designed milling toolpath. An experienced technician prepared the 25 × 3 × 0.5 mm3 metal bars and applied bonding agent using one of three application methods, and then applied opaque, dentin, and enamel porcelains according to manufacturer’s instructions to a 8 × 3 × 1 mm3 porcelain. A control group received no bonding agent prior to porcelain application. The four groups (n = 11) were blindly tested for differences in bond strength using a universal testing machine in a three-point bend test configuration, based on ISO 9693–1:2012. Results: The average (SD) bond strengths for the control, powder, paste, and sprayon groups, respectively, were: 24.8 (2.6), 24.6 (2.6), 25.3 (4.0), and 24.1 (3.9) MPa. One-way ANOVA and Tukey’s multiple comparison tests were performed between all groups. There were no statistically significant differences among groups (p = 0.951). Conclusion: Titanium-porcelain bond strength was not affected by the use of a bonding agent or its application method when tested by ISO 9693–1 standard.

doi: 10.1111/jopr.12234

Critics of metal ceramic restorations claim that conventional high-noble (HN) alloys are too costly, that predominately base (PB) alloys have compromised biocompatibility, and that casting procedures for all alloys are technique sensitive.1 Of the many PB alloys that are less expensive than HN alloys, titanium has favorable mechanical properties2 and exhibits superior biocompatibility, partly because of its stable oxide layer.3 Although cost and biocompatibility may favor the use of Ti over other HN and PB alloys for metal ceramic restorations, procedures for casting Ti restorations are significantly more technique sensitive than for other PB alloys. Casting investments for traditional alloys include gypsum and phosphate-bonded investments and contain components that form stable oxides at high temperature. But because titanium oxide is thermodynam-

ically favored over those oxides formed in traditional investments, a magnesium oxide investment must be used to avoid excessive reactivity. Even using a magnesium investment, an unstable α-case layer is formed that must be eliminated before bonding to ceramic.4-6 In addition, Ti reacts vigorously with oxygen at casting temperatures and must therefore be cast in an inert gas environment. These difficulties in casting have been circumvented by the development of CAD/CAM milling technologies. It was recently shown that these systems can produce copings that, after the addition and firing of porcelain, exhibit less marginal discrepancy than Ni–Cr cast alloy systems.7 Even though Ti copings can be milled accurately, the application of the veneering porcelain presents particular challenges. The quantity of the crystalline leucite phase must be altered

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

1

Titanium-Ceramic Bonding Methods

Curtis et al

from traditional formulations to obtain a compatible coefficient of thermal expansion.8 Additionally, Ti must be fired at a lower temperature than conventional metal ceramic alloys because it undergoes a phase change at 860 °C. Therefore, glass modifiers like boric oxide are added to decrease the firing temperature and control the oxide layer thickness and porcelain adherence.9 The oxide layer also may be controlled via application of a bonding agent between the porcelain and the metal. Metal oxides are added to such bonding agents to scavenge oxygen and optimize the thickness and properties of the oxide layer. The composition of bonding agents varies widely between manufacturers and may lead to important differences in bond strengths. Titankeramic (Vita, Bad Sackingen, Germany) bonding agent powder is estimated by energy-dispersive x-ray spectroscopy (EDS) to be composed of 71% SiO2 and 7% Al2 O3 . It has a uniform particle size, and the particles are probably of single composition.10 Bonding agents are supplied as powder/liquid systems, as pastes, and a unique commercial spray-on system (Vita SprayOn; Vita).11 A study employing a spray opaque and a PB alloy found that a more-even thickness was obtained with the spray application method, but bond strength was unaffected.12 Recent studies have sought to document improvements in titanium–ceramic bond strength through various surface treatments, such as chromium electroplating of the Ti,13 airborne particle abrasion,14 acid or caustic treatments,15 nitridation,16,17 and silicon coating by magnetron sputtering.18 No studies demonstrate the effect of bonding agent application methods on metal-ceramic bond strength in Ti systems. The three-point bending test proposed by Lenz et al commonly serves to facilitate comparison of materials and techniques involved in the metal-ceramic bond.19 Using this test, HN alloys with feldspathic porcelain systems demonstrate bond strengths averaging 50 to 60 MPa,20 while titanium-to-porcelain bond strengths average 20 to 40 MPa.21 A similar correlation is observed regarding clinical success. Walter et al22 showed 98% clinical success with HN metal-ceramic vs. 84% for titanium– ceramic systems over 6 years. The primary aim of this investigation was to evaluate the effect of three application methods of the bonding agent on the strength of the metal-ceramic bond in low-fusing porcelain/grade II commercially pure titanium systems. Because it was suspected that the spray-on technique would produce a layer of bonding agent more uniform in thickness and in coverage than brush-on application, it was hypothesized that the strongest bond identified by the three-point bending test would be associated with spray-on, followed by paste, and then brushon bonding agent. The null hypothesis was that there would be no significant difference in bond strengths between the three application methods.

Materials and methods The experiment was designed to observe the effects of bonding agent application on the bond strength of the Ti-porcelain system. The application method was tested at four levels: no bonding agent, paste bonding agent, powder bonding agent brush-on, and powder bonding agent spray-on. Based on a priori power analysis, eleven specimens were used in each of four 2

Figure 1 Image of milled beams embedded in resin. One wafer, previously cut from an ingot, is placed on removable custom-milled titanium positioning jig. Acrylic resin is poured into the ring. The jig is removed, and the wafer remains. The wafer is then milled using a custom toolpath to achieve proper thickness and width of four beams.

groups. The eleven specimens without bonding agent served as controls. Specimen preparation

Specimen preparation resulted in a 25 × 3 × 0.5-mm3 -milled beam with porcelain applied to a central 8 × 3 × 1.0 mm3 block R in accordance with ISO 9693.23 KaVo Everest T-B33/16 (Kavo, Biberach, Germany) Ti blanks in ASTM grade II cp Ti were machined into 3-mm-thick wafers using an industrial machining blade. These machined wafers were placed on a milled positioning jig and invested in acrylic resin for use in R the Everest milling unit. From each wafer, four Ti beams were milled to 3 mm wide x 0.5 mm thick (Fig 1). Each beam was standardized to 25 mm length using a thin carborundrum disc. Each set of four beams originating from a single wafer was randomized, received the assigned treatment, and underwent firing at the same time. A custom milling toolpath was designed using DentMill dental CADCAM suite (Delcam, Salt Lake City, UT) to replicate the surface texture of a milled coping. Once the beams were milled, all further specimen preparation was done by the same experienced dental technician. The surface was prepared according to manufacturer instructions using airborne particle abrasion and steam cleaning. The beams were abraded with airborne particles of 120-μm aluminum oxide at a 2-bar pressure, a 3 to 5 cm distance, and an angle of approximately 45° to the surface. Each specimen was cleaned with pressurized steam before bonding agent application. For beams in the brush-on powder/liquid groups, powder was mixed with the manufacturer’s supplied modeling liquid (Modeling liquid) and then applied using a sable brush. The paste bonding agent was applied in the same manner, using paste bonding agent liquid to maintain consistency. For those beams in the spray-on group, powder was mixed with the manufacturer’s solvent (Spray-on Liquid; Vita) in the jar of a HS #1 airbrush (Paache, Chicago, IL). Following the manufacturer’s instructions, the mix was standardized to 5 g of powder to 95 mL of liquid. It was sprayed at 6 to 8 cm from the surface,

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

Curtis et al

Titanium-Ceramic Bonding Methods

Table 2 Average bond strengths Group 1 (n = 11) 2 (n = 11) 3 (n = 11) 4 (n = 11)

Figure 2 Teflon device for porcelain application. The groove in the large block is 1.5 mm deep, allowing 0.5 mm for the Ti beam thickness and 1.0 mm for the thickness of the applied porcelain. The two additional removable Teflon pieces are secured by friction and serve to form the porcelain block. Table 1 Firing procedures used

Cycle Paste bonder Powder bonder Opaque Dentin Glaze without glaze

Temp (°C)

Hold (min)

Rate (°C/min)

Temp (°C)

Hold (min)

400 400 400 400 400

6 2 2 6 0

67 67 98 53 93

800 800 790 770 770

1 1 1 1 1

at 1 bar pressure, and through a tip opening of 0.63 mm and then fired according to manufacturer’s instructions. The solvent contained 90% ethanol, which allowed the bonding agent to dry quickly once on the metal. Eleven Ti beams served as the control group and did not receive any bonding agent. For all four groups, the opaque layer was applied next with a conventional brush-on technique. A device to facilitate the formation of the 1-mm-thick porcelain block (Fig 2) was made using three pieces of fluorocarbon resin (Teflon; DuPont, Wilmington, DE). The opaque layer was fired before 3D shade 2M2 dentin porcelain (Vita), and the corresponding EN2 shade enamel porcelain was next applied and fired simultaneously. The surface was then lightly ground using three strokes of a cylindrical diamond bur (S5117–023; Brasseler USA, Savannah, GA) to simulate a minor contour adjustment. The specimens were then fired in a calibrated Ney Centurion Qex oven (Denstply Ceramco; Burlington, NJ) according to manufacturer instructions24 (Table 1). Following bonding agent application, each set of four specimens was fired in the same oven for: (1) opaque; (2) body; (3) enamel; and (4) glaze (without addition of glaze material) firings. Mechanical testing

An MTS electromechanical universal testing machine (MTS Systems Corporation; Eden Prairie, MN) was used to load the prepared specimens using a 3-point bend apparatus. Each spec-

Mean bond strength (MPa)

Standard deviation

24.8 24.6 25.3 24.1

2.6 2.6 4.0 3.9

imen was placed on two rods of 1 mm radius separated by 20 mm. A third rod of 1 mm radius was attached to the crosshead. The crosshead was advanced at 1.0 mm/min, and the force at which the specimen failed was recorded using a software packR ; MTS Systems Corporation). In accordance age (Testworks with the ISO standard, the porcelain was placed on the side opposite the crosshead, and the crosshead contacted the Ti surface of each specimen. Following mechanical failure, each specimen was examined using a light microscope at 20× magnification, and the mode of failure was recorded. If unveneered opaque porcelain remained, the specimen was considered a mixed failure. If veneering porcelain remained without exposure of opaque, a cohesive failure was recorded. If no remnant of opaque porcelain could be observed, the failure was considered adhesive. For each specimen, τ b was calculated according to the equation described in ISO 9693: τb = k · F f ail where k is a constant determined from an algorithm appended to the ISO standard derived from finite element analysis.19 The thickness and the modulus of elasticity of the Ti beam are used to determine normalized average shear stress. For the Ti beams in this study with thickness 0.43 to 0.54 mm, k ranged from 4.1 to 6.3 mm−2 . One representative specimen from each group was examined under SEM without undergoing mechanical testing. The speciR ; Buehler, Lake mens were mounted in epoxy resin (Epoxicure Bluff, IL) and finished and polished with silicon carbide paper and then with aluminum oxide to a final grit of 0.1 μm. The specimens were then carbon coated and examined at 5000× by SEM (Zeiss Sigma VP 4.0; Zeiss, Oberkochen, Germany). One-way ANOVA with pairwise comparisons using least significant differences corresponding to this analysis was applied. Residuals were checked for homogeneity of variance and approximate normal distribution to assure the analysis assumptions were met. Based on the sample size of 11 and the pooled variance estimate of the treatment groups, there was 0.80 power to detect a difference of 4.4 between two treatments at a twotailed significance level of 0.05. Least significant difference pairwise comparisons were performed following the one-way ANOVA.

Results The average control, powder, paste, and spray-on bond strengths were calculated (Table 2). One-way ANOVA was performed, and no statistically significant differences between groups were identified. No post hoc test was indicated. Each

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

3

Titanium-Ceramic Bonding Methods

Curtis et al

Figure 3 SEM image of control, 5 k magnification (Ti on right). The superimposed plots represent EDS analysis along the horizontal yellow line. The highest lines are titanium and oxygen.

Figure 5 SEM image of paste, 5 k magnification (Ti on right). The superimposed plots represent EDS analysis along the horizontal yellow line. The highest lines are titanium and oxygen.

Figure 4 SEM image of powder, 5 k magnification (Ti on right). The superimposed plots represent EDS analysis along the horizontal yellow line. The highest lines are titanium and oxygen.

Figure 6 SEM image of spray-on, 5 k magnification (Ti on right). The superimposed plots represent EDS analysis along the horizontal yellow line. The highest lines are titanium and oxygen.

group had one adhesive failure (9.1%), with the remaining ten specimens (90.9%) failing in a mixed mode. Statistically, the null hypothesis was not rejected. Examination of the interface using SEM is shown in Figures 3 to 6. Under standardized in vitro test conditions, the ISO test failed to distinguish bond strengths between bonding agent application methods.

Discussion No increase in bond strength was observed when applying a bonding agent to the milled titanium surface, conflicting with previous findings using different titanium-porcelain systems. R In the early Procera Ti porcelain system (Nobel Biocare, Zurich, Switzerland), the bonding agent significantly improved bond strength.25 The Noritake product (Ti-22; Kuraray Noritake Dental Inc, Kurashiki, Japan) has also been tested, and the bonding agent shown to improve the bond strength.26 The GC bonding agent (GC Corp, Tokyo, Japan) improved (increased) 4

the strain energy release rate in one study.27 But the Vita Titankeramic system has not been tested against its un-bonded control. Chakmakchi et al10 showed that the Titankeramic system contains 15% to 30% more silica and tin oxide than other bonding agents. Perhaps our finding that the bonding agent did not affect the bond strength may be attributed to the difference in chemical composition of the Titankeramic bonding agent. In addition to the chemical factors attributed to the bonding agents, mechanical factors have been investigated and shown to affect the bond strength. Goeller et al28 and McLean and Sced29 showed a roughened metal surface increases bond strength. But Carpenter and Goodkind30 found that a roughened surface increases the oxide layer, which appears helpful in the case of other PB alloys, but may not be helpful in Ti. Of 37 previous in vitro studies on the titanium/porcelain bond, 1318,26,31-41 cast the Ti; two42,43 reported milling the Ti surface or receiving manufacturer-supplied milled beams, 185,9,14,15,17,21,25,27,44-53 polished the beams, and four13,54-56 did not describe the Ti

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

Curtis et al

Titanium-Ceramic Bonding Methods

Figure 7 Image of milled coping made using noncontact surface profilometer. Warm colors represent elevations, and cool colors represent depressions, with the z-axis scale on the right of the diagram. The x- and y-axis scales are on the bottom and left of the image.

Figure 8 Image of milled beam made using noncontact surface profilometer. Warm colors represent elevations, and cool colors represent depressions, with the z-axis scale on the right of the diagram. The x- and y-axis scales are on the bottom and left of the image.

surface preparation or roughness. Of nine clinical studies, only Kaus et al24 evaluated cast Ti; six studies22,57-61 evaluated the R Procera milling/EDM process, and two62,63 evaluated the R Everest milling process. In this investigation, we tried to duplicate the surface roughness and topography of clinical specimens. The surface roughness of a typical milled Ti coping (Fig 7) was compared with that of the beams (Fig 8) using a noncontact surface profilometer (ContourGT; Bruker USA, Camarillo, CA). The Ra calculated for the coping was 0.69 and for the beams 0.48. Reproduc-

ing the same surface is important when evaluating milled vs. cast Ti, since the milling process may affect the overall bond strength. Although we attempted to simulate clinical conditions, the results of this in vitro study should not compel a cautious clinician or technician to omit the application of bonding agent. Studies designed to elicit adhesive failures could provide additional insights and enhance the overall value of our data. In addition to better quantification of bond strength, additional studies could focus on a more objective analysis of the character

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

5

Titanium-Ceramic Bonding Methods

Curtis et al

and dimension of the oxide layer formed using different application techniques. 18.

Conclusion In a three-point bend test of Titankeramic porcelain and bonding agents on milled ASTM type II commercially pure titanium, there was no statistically significant difference between a control with no bonder, spray-on bonder, powder-liquid bonder applied with a brush, and paste bonder. Qualitative examination of SEM images revealed a thicker oxide layer in those specimens without a bonding agent and a thinner, more uniform oxide layer in those using a spray-on bonding agent.

20. 21.

22.

23.

References 1. Zarone F, Russo S, Sorrentino R: From porcelain-fused-to-metal to zirconia: clinical and experimental considerations. Dent Mater 2011;27:83-96 2. Wang RR, Fenton A: Titanium for prosthodontic applications: a review of the literature. Quintessence Int 1996;27:401-408 3. Haag P, Nilner K: Questions and answers on titanium-ceramic dental restorative systems: a literature study. Quintessence Int 2007;38:e5-13 4. Boening KW, Walter MH, Reppel PD: Non-cast titanium restorations in fixed prosthodontics. J Oral Rehabil 1992;19:281-287 5. Oka K, Hanawa T, Kon M, et al: Effect of barium in porcelain on bonding strength of titanium-porcelain system. Dent Mater J 1996;15:111-120 6. Guilin Y, Nan L, Yousheng L, et al: The effects of different types of investments on the alpha-case layer of titanium castings. J Prosthet Dent 2007;97:157-164 7. Shokry TE, Attia M, Mosleh I, et al: Effect of metal selection and porcelain firing on the marginal accuracy of titanium-based metal ceramic restorations. J Prosthet Dent 2010;103:45-52 8. Hautaniemi, JA: Effect of crystalline leucite on porcelain bonding on titanium. J Am Ceram Soc 1991;74:1449-1451 9. Adachi M, Mackert JR, Parry EE, et al: Oxide adherence and porcelain bonding to titanium and Ti-6Al-4V alloy. J Dent Res 1990;69:1230-1235 10. Chakmakchi M, Eliades G, Zinelis S: Bonding agents of low fusing cpTi porcelains: elemental and morphological characterization. J Prosthodont Res 2009;53:166-171 11. Espinosa-Marino J, Alvarez-Arenal A: Fabricating a spray gun for opaque porcelain application. J Prosthet Dent 1996;75:575. 12. J¨orn D, Waddell JN, Swain MV: The influence of opaque application methods on the bond strength and final shade of PFM restorations. J Dent 2010;38(supplement 2):e143-149 13. Elsaka SE, Hamouda IM, Elewady YA, et al: Effect of chromium interlayer on the shear bond strength between porcelain and pure titanium. Dent Mater 2010;26:793-798 14. Sadeq A, Cai Z, Woody RD, et al: Effects of interfacial variables on ceramic adherence to cast and machined commercially pure titanium. J Prosthet Dent 2003;90:10-17 15. Troia MG Jr, Henriques GEP, Mesquita MF, et al: The effect of surface modifications on titanium to enable titanium-porcelain bonding. Dent Mater 2008;24:28-33 16. Oshida Y, Fung LW, Isikbay SC: Titanium-porcelain system. Part II: bond strength of fired porcelain on nitrided pure titanium. Biomed Mater Eng 1997;7:13-34 17. Xia Y, Zhou S, Zhang F, et al: Effect of ZrN coating by magnetron sputtering and sol-gel processed silica coating on 6

19.

24.

25.

26.

27.

28.

29.

30.

31.

32.

33.

34.

35.

36. 37.

38.

titanium/porcelain interface bond strength. J Mater Sci Mater Med 2011;22:317-325 Ozcan I, Uysal H: Effects of silicon coating on bond strength of two different titanium ceramic to titanium. Dent Mater 2005;21:773-779 Lenz J, Schwarz S, Schwickerath H, et al: Bond strength of metal-ceramic systems in three-point flexure bond test. J Appl Biomater 1995;6:55-64 Fischer J, Zb¨aren C, Stawarczyk B, et al: The effect of thermal cycling on metal-ceramic bond strength. J Dent 2009;37:549-553 Zinelis S, Barmpagadaki X, Vergos V et al: Bond strength and interfacial characterization of eight low fusing porcelains to cp Ti. Dent Mater 2010;26:264-273 Walter M, Reppel PD, B¨oning K, et al: Six-year follow-up of titanium and high-gold porcelainfused- to-metal fixed partial dentures. J Oral Rehabil 1999;26:91-96 Schwickerath H: Pr¨ufung der Verbundfestigkeit Metall-Keramik [Test of the bonding strength of metal ceramics]. Dtsch Zahn¨arztl Z 1983;38:21-25 Kaus T, Pr¨obster L, Weber H: Clinical follow-up study of ceramic veneered titanium restorations- three-year results. Int J Prosthodont 1996;9:9-15 Gilbert JL, Covey DA, Lautenschlager EP: Bond characteristics of porcelain fused to milled titanium. Dent Mater 1994;10:134-140 Al-Hussaini I, Al-Wazzan KA: Effect of surface treatment on bond strength of low-fusing porcelain to commercially pure titanium. J Prosthet Dent 2005;94:350-356 Tholey MJ, Waddell JN, Swain MV: Influence of the bonder on the adhesion of porcelain to machined titanium as determined by the strain energy release rate. Dent Mater 2007;23:822-828 Goeller I, Meyer JM, Nally JN: Comparative study of three casting agents and their influence on bond strength of porcelain-fused to gold alloys. J Prosthet Dent 1972;28:504-508 McLean JW, Sced IR: Bonding of dental porcelain to metal: II. The base metal alloy/porcelain bond. Trans Brit Soc 1973;72:235-238 Carpenter MA, Goodkind RJ: Effect of varying surface texture on bond strength of one semi-precious non-precious ceramo alloy. J Prosthet Dent 1979;42:86-92 Ats¨u S, Berksun S: Bond strength of three porcelains to two forms of titanium using two firing atmospheres. J Prosthet Dent 2000;84:567-574 Bondioli IR, Bottino, MA: Evaluation of shear bond strength at the interface of two porcelains and pure titanium injected into the casting mold at three different temperatures. J Prosthet Dent 2004;91:541-547 Garbelini WJ: Evaluation of low-fusing ceramic systems combined with titanium grades II and V by bending test and scanning electron microscopy. J Appl Oral Sci 2003;11:354-360 Papadopoulos TD, Spyropoulos KD: The effect of a ceramic coating on the cpTi-porcelain bond strength. Dent Mater 2009;25:247-253 Pr¨obster L, Maiwald U, Weber, H: Three-point bending strength of ceramics fused to cast titanium. Eur J Oral Sci 1996;104:313-319 Suansuwan N, Swain MV: Adhesion of porcelain to titanium and a titanium alloy. J Dent 2003;31:509-518 Tanaka Y, Watanabe I, Okabe T: Cross-sectional TEM analysis of porcelain fused to gold-coated titanium. Dent Mater J 2007;26:84-88 Tr´oia MG, Henriques Jr GEP, N´obilo MAA, et al: The effect of thermal cycling on the bond strength of low-fusing porcelain to commercially pure titanium and titanium-aluminium-vanadium alloy. Dent Mater 2003;19:790-796

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

Curtis et al

39. V´asquez VZ, Ozcan M, Kimpara ET: Evaluation of interface characterization and adhesion of glass ceramics to commercially pure titanium and gold alloy after thermal- and mechanical-loading. Dent Mater 2009;25:221-231 40. Yilmaz H, Dinc¸er C: Comparison of the bond compatibility of titanium and a NiCr alloy to dental porcelain. J Dent 1999;27:215-222 41. Yoda M, Konno T, Takada Y, et al: Bond strength of binary titanium alloys to porcelain. Biomaterials 2001;22:1675-1681 42. Inan O, Acar A, Halkaci S: Effects of sandblasting and electrical discharge machining on porcelain adherence to cast and machined commercially pure titanium. J Biomed Mater Res B: Appl Biomater 2006;78:393-400 43. Pang IC, Gilbert JL, Chai J, et al: Bonding characteristics of low-fusing porcelain bonded to pure titanium and palladium-copper alloy. J Prosthet Dent 1995;73:17-25 44. Bienia´s J, Surowska B, Stoch A, et al: The influence of SiO2 and SiO2-TiO2 intermediate coatings on bond strength of titanium and Ti6Al4V alloy to dental porcelain. Dent Mater 2009;25:1128-1135 45. Galo R, Ribeiro RF, Rodrigues RC, et al: Effect of laser welding on the titanium ceramic tensile bond strength. J Appl Oral Sci 2011;19:301-305 46. Homann F, Waddell JN, Swain MV: Influence of water, loading rate and bonder on the adhesion of porcelain to titanium. J Dent 2006;34:485-490 47. Iseri U, Ozkurt Z, Kazazoglu E: Shear bond strengths of veneering porcelain to cast, machined and laser-sintered titanium. Dent Mater J 2011;30:274-280 48. Kim JT, Cho SA: The effects of laser etching on shear bond strength at the titanium ceramic interface. J Prosthet Dent 2009;101:101-106 49. Lim HP, Kim JH, Lee KM et al: Fracture load of titanium crowns coated with gold or titanium nitride and bonded to low-fusing porcelain. J Prosthet Dent 2011;105:164-170 50. Rammelsberg P, Aschl I, Pospiech P: Verbundfestigkeit niedrigschmelzender Keramiken zu Titan unter Ber¨ucksichtigung der Oberfl¨achenkonditionierung. Dtsch Zahn¨arztl Z 1998;53:200-203

Titanium-Ceramic Bonding Methods

51. Wang CS, Chen KK, Tajima K, et al: Effects of sandblasting media and steam cleaning on bond strength of titanium-porcelain. Dent Mater J 2010;29:381-391 52. White SN, Ho L, Caputo AA, et al: Strength of porcelain fused to titanium beams. J Prosthet Dent 1996;75:640-648 53. Zhang LJ, Wang ZY, Gao B, et al: [Effect of sand-blasting particle sizes on bonding strength between porcelain and titanium fabricated by rapid laser forming.]. Zhonghua Kou Qiang Yi Xue Za Zhi 2009;44:677-680 54. Oshida Y, Fung LW, Isikbay SC: Titanium-porcelain system. Part II: bond strength of fired porcelain on nitrided pure titanium. Biomed Mater Eng 1997;7:13-34 55. Reyes MJ, Oshida Y, Andres CJ, et al: Titanium-porcelain system. Part III: effects of surface modification on bond strengths. Biomed Mater Eng 2001;11:117-136 56. Sommer MC, Schwickerath H, Marx R, et al: [Mechanical strength of ceramic materials for fusing on titanium restorations]. Dtsch Zahnarztl Z 1991;46:746-749 57. Bergman B, Marklund S, Nilson H, et al: An intraindividual clinical comparison of 2 metal-ceramic systems. Int J Prosthodont 1999;12:444-447 58. Bergman B, Nilson H, Andersson M: A longitudinal clinical study of procera ceramic-veneered titanium copings. Int J Prosthodont 1999;12:135-139 59. L¨ovgren R, Andersson B, Carlsson GE, et al: Prospective clinical 5-year study of ceramic-veneered titanium restorations with the procera system. J Prosthet Dent 2000;84:514-521 60. Marklund S, Bergman B, Hedlund SO, et al: An intraindividual clinical comparison of two metalceramic systems: a 5-year prospective study. Int J Prosthodont 2003;16:70-73 61. Milleding P, Haag P, Neroth B, et al: Two years of clinical experience with Procera titanium crowns. Int J Prosthodont 1998;11:224-232 62. Boeckler AF, Lee H, Stadler A, et al: Prospective observation of CAD/CAM titanium ceramic single crowns: a three-year follow up. J Prosthet Dent 2009;102:290-297 63. Boeckler AF, Lee H, Psoch A, et al: Prospective observation of CAD/CAM titanium-ceramic-fixed partial dentures: 3-year follow-up. J Prosthodont 2010;19:592-597

C 2014 by the American College of Prosthodontists Journal of Prosthodontics 00 (2014) 1–7 

7

Effect of Bonding Agent Application Method on Titanium-Ceramic Bond Strength.

Although milled titanium may be used as a substructure in fixed and implant prosthodontics, the application of the veneering porcelain presents partic...
761KB Sizes 0 Downloads 6 Views