Você está na página 1de 8

Surface & Coatings Technology 207 (2012) 135142

Contents lists available at SciVerse ScienceDirect

Surface & Coatings Technology


journal homepage: www.elsevier.com/locate/surfcoat

Fretting wear and friction reduction of CP titanium and Ti6Al4V alloy by ultrasonic nanocrystalline surface modication
Auezhan Amanov a, b, In-Sik Cho c, Dae-Eun Kim d,, Young-Sik Pyun a,
a

Department of Mechanical Engineering, Sun Moon University, Asan 336708, South Korea Center for Nano-Wear, Yonsei University, Seoul 120749, South Korea Department of Hybrid Engineering, Sun Moon University, Asan 336708, South Korea d Department of Mechanical Engineering, Yonsei University, Seoul 120749, South Korea
b c

a r t i c l e

i n f o

a b s t r a c t
Application of surface modication techniques is expected to be a viable solution to mitigate fretting damage and to reduce friction. In this paper, the aim was to improve the fretting wear and friction characteristics of commercially pure titanium (CP Ti) and Ti6Al4V alloy by using an ultrasonic nanocrystalline surface modication (UNSM) technique. Lubricated fretting wear and friction tests were conducted with a ball-on-at conguration on untreated and UNSM-treated specimens using silicon nitride (Si3N4) balls. The results showed that the fretting wear and friction coefcient characteristics of the UNSM-treated specimens were improved compared to those of the untreated specimens. Moreover, it was found that the fretting wear scar diameter and depth of the UNSM-treated specimens were smaller and shallower compared to those of the untreated specimens. Surface analysis was performed using a scanning electron microscope (SEM). 2012 Elsevier B.V. All rights reserved.

Article history: Received 17 January 2012 Accepted in revised form 14 June 2012 Available online 26 June 2012 Keywords: CP Ti Ti6Al4V alloy Friction Fretting wear Ultrasonic surface modication

1. Introduction Fretting is a wear phenomenon that occurs when two contacting solids are subjected to a relative oscillatory tangential motion of small displacement amplitude typically less than 100 m [1]. The damage due to fretting wear can accelerate fatigue failure of components by creating crack initiation sites on the surface. Fretting wear is commonly encountered in various types of materials that are used as machine components, engineering structures and aerospace parts that experience vibration. Particularly, CP Ti and Ti6Al4V alloy that are mostly used in aerospace, biomedical and other applications are quite susceptible to fretting related failures. Despite their attractive mechanical and physical properties CP Ti and Ti6Al4V alloy display relatively poor fretting and wear resistance [26]. This is due to their high surface energy which promotes metal transfer, seizure and adhesive wear in tribological applications. It is therefore important to improve the friction and wear properties of these materials, particularly under fretting wear conditions [7,8]. In order to prevent fretting wear, modication of the surface to improve the tribological properties is necessary. To this end, it has been a great challenge to develop an effective surface modication technique for fretting applications.

Correspondence to: D.-E. Kim, Department of Mechanical Engineering, Yonsei University, Seoul 120749, South Korea. Tel.: +82 2 2123 2822; fax: +82 2 312 2159. Correspondence to: Y.-S. Pyun, Department of Mechanical Engineering, Sun Moon University, Asan 336708, South Korea. Tel.: +82 41 530 2333; fax: +82 41 530 2307. E-mail addresses: kimde@yonsei.ac.kr (D.-E. Kim), pyoun@sunmoon.ac.kr (Y.-S. Pyun). 0257-8972/$ see front matter 2012 Elsevier B.V. All rights reserved. doi:10.1016/j.surfcoat.2012.06.046

Currently, the dovetail surfaces of compressor blades are coated with plasma sprayed coatings and dry lm lubricants to impede fretting wear and prolong the life of the blades and disks [9]. However, there are many on-going investigations with the purpose of developing longer lasting coatings for compressor parts [10]. Surface modication techniques such as ion-implantation (II), laser beam quenching (LBQ), shot peening (SP), laser shot peening (LSP), plasma nitriding (PN), plasma immersion ion implantation (PIII) and surface mechanical attrition treatment (SMAT) have already been identied as effective methods to enhance the ability of materials to resist fretting wear [1117]. The SP process is probably the most popular surface modication technique among the above cited techniques. It is also adopted generally by industry due to its versatility in treating components of non-planar geometries. In the SP technique, spherical shots (balls) with sizes in the range of 0.251 mm are blasted onto the workpiece surface at impact velocities in the range of 20150 m/s under a controlled atmosphere [18]. The impact of the shots induces compressive residual stresses and work hardening to the surface region of the workpiece [19], but does not always generate a nanocrystalline surface layer [20,21]. Recently, a new surface modication technique called ultrasonic nanocrystalline surface modication (UNSM) which utilizes an ultrasonic vibration at 20 kHz was developed. The UNSM technique involves higher kinetic energies than the other techniques cited above since the ball (tip) strikes the workpiece surface under a high frequency of 20 kHz. Also, the surface roughness of the specimen after the UNSM treatment tends to be much smoother than what can be achieved with the SP process. In the UNSM technique, the ball (tip) trace can be

136

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142 Table 2 Chemical composition (wt.%) of CP Ti and Ti6Al4V alloy specimens. Material CP Ti Ti6Al4V alloy C 0.1 0.01 N 0.03 0.006 Fe 0.3 0.2 Al 6.47 V 3.89 H 0.015 O 0.25 0.17 Ti 99.305 89.254

controlled by a computer numerical control (CNC) machine. Subsequently, UNSM technique can result in homogenous microstructure, thicker nanocrystalline and work-hardened surface layers, and deeper surface regions with high compressive residual stress. Also, the thickness of the nanocrystalline surface layer can be controlled with better accuracy than process such as SP. In this work, UNSM technique was applied to CP Ti and Ti6Al4V alloy materials in order to improve their tribological properties. The UNSM is a novel surface modication technique which improves the tribological properties of interacting surfaces in relative motion. Detailed description of the UNSM process and its effects on metal and alloy properties as well as microstructure is available in the literature [2225]. In previous studies, fretting wear properties of CP Ti and Ti6Al4V alloy were investigated at a frequency range from 2 to 300 Hz in ambient environment at room temperature [2629]. However, all fretting wear tests in this study were conducted at a high-frequency of 20 kHz by using a newly developed fretting wear test rig. The high-frequency fretting wear behavior of UNSM-treated and untreated specimens made of AISI304 stainless steel was investigated in a previous study [23]. However, the effect of UNSM on the tribological properties of CP Ti and Ti6Al4V alloy was expected to be different from that of AISI304 steel because of the difference in the material structure and properties. The fretting wear mechanism and tribological characteristics of the specimens were investigated systematically through controlled fretting tests and rigorous surface characterization. The objective of this research was to improve the friction and fretting wear characteristics of CP Ti and Ti6Al4V alloy by applying the UNSM technique. 2. Experimental details 2.1. Ultrasonic nanocrystalline surface modication process UNSM is a technique that can be used to improve the mechanical surface properties of metals and alloys. The technique utilizes ultrasonic vibration energy. The principle of UNSM is based on the instrumental conversion of harmonic oscillations of an acoustically tuned body into resonant impulses of ultrasonic high-frequency. The energy generated from the oscillations is used to impact the workpiece surface from 20,000 to 40,000 shots per square millimeter. The roughness of the workpiece can be readily controlled by varying the impact load during the UNSM process. In this work, the CP Ti and Ti6Al4V alloy specimens were treated under the UNSM conditions as shown in Table 1. 2.2. Materials and test conditions The materials under investigation in this study were CP Ti and alpha/beta Ti6Al4V alloy that are widely used in aeronautics, especially for the compressor blades and disks, and surgical implants, automotive and marine parts, reactor vessels and heat exchangers. Their chemical composition and mechanical properties are given in Tables 2 and 3, respectively. It should be noted that Al helps to stabilize the alpha phase, while V stabilizes the beta phase [30]. Fig. 1 shows the microstructure of CP Ti and Ti6Al4V alloy identifying the alpha/beta phases. The phase has an HCP structure whereas

the phase has a BCC structure. In Fig. 1 the white phase is and the gray phase is . The bright spots in phase are due to the etching effect. The friction coefcients of both untreated and UNSM-treated disk specimens were obtained using a microtribometer (UMT-2 CETR, USA) with a ball-on-disk conguration under the conditions as shown in Table 4. Silicon nitride (Si3N4) balls with a diameter of 1.6mm were used as the counter surface in all the friction tests and silicon oil was used as a lubricant. In friction and fretting wear tests, all the specimens were prepared to have the same surface roughness value of Ra = 0.10 m in order to eliminate the surface roughness effect on the tribological behavior. The surface roughness was measured using a surface prolometer (Mitutoyo SJ-400). Fretting experiments were conducted using a newly developed high-frequency fretting wear test rig under the conditions as shown in Table 5. All experiments were conducted at a constant test frequency of 20 kHz for up to 110 6 cycles. At least three tests were conducted under each experimental condition. A schematic of the highfrequency fretting wear test rig is shown in Fig. 2. The fretting contact consisted of a nominally at specimen with dimensions of 4020 10 mm in contact with a Si3N4 ball with a diameter of 12.7 mm and a hardness of HV 1700. All the fretting tests were performed under lubricated conditions at room temperature with a relative humidity of 42%. As shown in the schematic the test specimen was mounted on a xture and placed on top of the ball attached to the ultrasonic horn. The normal load was applied by dead weight. The oscillatory motion of the ball was controlled by the ultrasonic booster system. The peak-to-peak displacement (stroke) was automatically maintained to be constant throughout the test by a photonic sensor. Under the very high-frequency of 20 kHz, the temperature of the contact zone may increase signicantly and this may affect the fretting wear mechanism due to oxidation as well as alteration in the mechanical properties of the specimen. Therefore, in order to prevent high temperature at the contact zone the contact region was cooled by an air spraying system. The average fretting wear scar dimensions measured using a surface prolometer were used for calculating the fretting wear volume loss. The wear volume loss, V, was calculated from the wear scar diameter using the approximate equation given by Halling [31]: V Rh 1h=4R
2 4

where R is the radius of the ball and h is the maximum depth of the wear scar. h is related to the measured mean wear scar diameter by the following equation:  1=2 2 2 h R R d =4 2

Table 1 UNSM treatment process conditions for CP Ti and Ti6Al4V alloy specimens. Material Frequency, Amplitude, Static Spindle Feed rate, Ball diameter, kHz m load, speed, mm/rev mm rpm N 30 30 60 30 0.07 2.38

Table 3 Mechanical properties of CP Ti and Ti6Al4V alloy specimens. Values UTS (MPa) Yield Stress, Y02 (MPa) Density (g/cm3) Elongation (%) CP Ti 434 275 4.52 20 Ti6Al4V alloy 1020 970 4.4 14

CP Ti 20 Ti6Al4V alloy

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142

137

(a)

(b)
+

Elem ment Ti K Total l

We eigh ht% 100 0.00 0 100 0.00 0

Ato omic c% 100 0.00 0

E Elem ment t A Al K T Ti K VK T Tota al

W Weig ght% % 5. .96 91.42 2 2. .62 10 00.0 00 -ph hase e

tom mic% % At 10 0.13 87 7.51 2.3 36

E Elem ment t A Al K T Ti K VK F Fe k T Tota al

Weig ght% % W 4. .17 78 8.51 1 15 5.41 1 1. .91 10 00.0 00 -ph hase e

At tom ic% % 7,2 25 76 6,94 14 4.20 1.6 61

Fig. 1. Microstructure of CP Ti (a) and + Ti6Al4V alloy (b).

where d is the mean wear scar diameter. An average of at least four measured fretting wear scar diameter values were used for calculating the wear scar depth. This method of quantifying fretting wear has been employed in other works as well [32,33]. Prior to testing, all the specimens were cleaned with acetone and ethanol for 5min each to remove all the contaminants from the surface.

After the tests, the specimens were again ultrasonically cleaned with acetone and ethanol for 5min each to eliminate the wear debris trapped in the fretting scar. The microstructure of the fretted surfaces was analyzed using an SEM (JEOL JSM-5610). For SEM observations, the specimens were polished with 1 m diamond paste and etched with a solution of 2ml HF+5ml H2O2 +100ml H2O. The Vickers microhardness

Table 4 Ball-on-disk friction test conditions performed using a silicon nitride ball with a diameter of 1.6 mm. Normal load, mN 50 Rotational speed, rpm 100 Sliding distance, m 37.68 Testing track radius, mm 6 Maximum contact pressure, GPa CP Ti 0.49 Ti6Al4V alloy 0.51 Temperature, C 21

138

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142

Table 5 Fretting wear test conditions performed using a silicon nitride ball with a diameter of 12.7 mm. Normal load, N 100, 200, 300 and 400 Displacement amplitude, m 30 Frequency, kHz 20 Fretting time, cycles 110
6

Maximum contact pressure, GPa CP Ti 1.55, 1.96, 2.24 and 2.47 Ti6Al4Valloy 1.6, 2.02, 2.31and 2.55

measurement was carried out using Mitutoyo HM-103 micro-Vickers hardness testing machine on the specimens at a load of 50gf. The residual stress measurement was performed using the X 3000 (X stress 3000) equipment with a tube current of 40mA at a tube voltage of 40kV. In order to obtain the residual stress values with respect to depth the surface layers of the specimens were removed by electropolishing which can remove the top surface layer without inducing additional residual stress. 3. Results and discussion 3.1. Microhardness and compressive residual stress results The microhardness and compressive residual stress values with respect to depth from the surface were obtained to assess the effective depth of UNSM treatment. The specimens were cut perpendicular to the surface using a diamond cutter and microhardness tests were performed on the cross section of the specimens. Relatively low load and speed conditions were applied during the specimen cutting process in order to minimize the possibility of specimen surface modication. Vickers-hardness and residual stress tests for the untreated and UNSM-treated CP Ti and Ti6Al4V alloy specimens were obtained as shown in Figs. 3 and 4, respectively. It could be seen from Fig. 3 that at the top surface the hardness of CP Ti increased from 146HV to 193 HV, while the hardness of Ti6Al4V alloy increased from 328 HV to 379 HV. However, the depth of the hardened layer was about 176 m and 200 m for the CP Ti and Ti6Al4V alloy specimens, respectively. The depth of the hardened layer could be estimated from Fig. 3 by noting the depth at which the microhardness equaled to that of the untreated specimen. It could also be found from Fig. 4 that the depth of the hardened layer become equal at a depth of about 250 m since the residual stress of the UNSM-treated specimens kept decreasing as a function of depth from the top surface, while the residual stress of the untreated specimens stabilized at a depth of 60 m. It has been

previously reported that improvement in surface hardness by SMAT was benecial in increasing the friction and fretting wear resistance [17,34]. The increase in hardness due to UNSM treatment can be attributed to both grain renement and work-hardening effects on the surface layer following the HallPetch relationship [17]. X-ray diffraction (XRD) is the most appropriate method for quantifying the residual stress produced by surface treatments [35]. In this study, the psi-splitting X-ray diffraction method was applied to determine the residual stress along the axial direction of the specimen by using the diffraction pattern of the Fe (2 1 1) crystal plane obtained by Cu K radiation. As shown in Fig. 4, the highest compressive residual stress at the top surface of CP Ti and Ti6Al4V alloy reached up to 1279.4 and 1142.7 MPa, respectively, as a result of the UNSM treatment due to local plastic deformation and increased strain hardening. The magnitude of compressive residual stress decreased with increasing depth from the top surface. The high residual compressive stresses are benecial for increasing the fretting wear resistance and the high hardness can be helpful to deter mechanical surface damage [36]. However, it was also shown in Fig. 4 that some compressive stresses exist at the surface of the untreated specimens. This is attributed to the surface nishing process that was used to prepare the specimen. Zhao et al. showed that surface nishing can introduce large compressive residual stress at the surface [37]. Chou et al. also demonstrated by using a thin plate bending method that surface nishing procedures such as polishing and grinding can generate compressive residual stress on the workpiece surface [38]. The presence of compressive residual stress at the surface is normally thought to be a result of local plastic deformation [39]. The type and magnitude of the compressive residual stress state is directly related to the machining and polishing conditions which vary signicantly with the technique used [40]. However, the observation of increased compressive residual stress in the UNSM-treated specimens, compared with those of the untreated specimens suggests that the surface and subsurface deformation in these specimens may be attributed to the greater extent of dislocation generation. Thus, it is

Fig. 2. Schematic of the high-frequency fretting wear test rig.

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142

139

Fig. 3. Variation in microhardness with respect to depth from the surface of untreated and UNSM-treated CP Ti (a) and Ti6Al4V alloy (b) specimens.

Fig. 4. Variation in residual stress with respect to depth from the surface of untreated and UNSM-treated CP Ti (a) and Ti6Al4V alloy (b) specimens.

reasonable to assume that the depth of compressive residual stress region extends to a depth commensurate with the depth of the dislocation generation zone.

3.2. Microstructure characteristics The cross-sectional microstructures of the UNSM-treated specimens were compared to those of the untreated specimens by SEM analysis. Fig. 5(b) shows the typical microstructure of plastically deformed CP Ti after the UNSM treatment. The grain size was measured by analyzing the electron backscatter diffraction (EBSD) observations using the TexSEM Laboratories (TSL) orientation imagining microscopy (OIM) Analysis 5 Program, which is a software for EBSD data acquisition and processing. The grain size measurement results revealed that the untreated CP Ti specimen had an initial grain size of about 35.5 m and it was rened to 200nm after the UNSM treatment. On the other hand, as for the Ti6Al4V alloy specimen, the initial grain size of and phases was 9.9 m and 3.8 m, respectively, and after the UNSM treatment they were rened to 1.2 m and 0.8 m, respectively. The grain size renement effect of the UNSM treatment was much more signicant for the CP Ti specimen than the Ti6Al4V alloy specimen. The rened grains are expected to lead to increase in hardness as a consequence of the predictions based on the HallPetch relationship [41].

Fig. 5(a) and (c) shows the cross-sectional microstructures of the untreated CP Ti and Ti6Al4V alloy specimens where the undeformed grains with a second phase can be found on the grain boundaries. It could be seen from Fig. 5(b) that the plastic deformation layers produced by the UNSM treatment on CP Ti have signicantly different features compared to those of the untreated specimen. It is well known that plastic deformed layers have a strong correlation with microstructures and mechanical properties in many metallic materials [42]. This plastically deformed layer leads to strain hardening and induces compressive stress at the surface of the metallic materials [43]. Also, grain boundaries became less apparent by the deformation of grains and mechanical twins were formed after the UNSM treatment. The precipitates were identied and located at the grain boundaries in the plastic deformation layer. As for the Ti6Al4V alloy specimen, Fig. 5(d) revealed clear evidence of changes in the microstructure and also showed that the initially continuous phase was fragmented after the UNSM treatment. In addition, the density of phase decreased from its initial state. 3.3. Frictional characteristics of untreated and UNSM-treated specimens Fig. 6 shows the average friction coefcient as a function of sliding distance for CP Ti and Ti6Al4V alloy specimens before and after the UNSM treatment. For the untreated CP Ti, the average friction coefcient increased within 5 m of sliding up to a value of 0.49 and reached a value of about 0.46. Also, the friction coefcient uctuated after it

140

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142

(a)

(b)

Precipitates

(c)

(d) +

Fig. 5. Cross-sectional SEM micrographs untreated CP Ti (a), UNSM-treated CP Ti (b), untreated Ti6Al4V alloy (c), and UNSM-treated Ti6Al4V alloy (d) specimens.

increased abruptly. The uctuation of the friction coefcient may be attributed to the localized fracture of the transfer layer and interaction of the particles at the sliding interface. It has been well reported that titanium alloys tend to experience material transfer to the counter surface when rubbed against other metals or ceramics [1,44,45]. For the UNSM-treated CP Ti, the average friction coefcient increased from an initial value of 0.28 up to a value of 0.37 during the rst 4 m and gradually decreased slightly. For the untreated Ti6Al4V, the average friction coefcient increased up to a value of 0.43 during the rst 15m and stabilized to a value of about 0.42. For the UNSM-treated Ti6Al4V, the average friction coefcient increased from the initial value of 0.25 up to a value of 0.32 within 25m of sliding and stabilized to a value of about 0.31. Comparable frictional behavior of Ti6Al4 V alloy has

been reported in previous studies [46,47]. It was postulated that the observed reduction in friction coefcient of the UNSM-treated specimen compared to that of the untreated specimen was related to the increase in hardness and compressive residual stress as well as alteration in the microstructure after the UNSM treatment. 3.4. Fretting wear characteristics of untreated and UNSM-treated specimens The effect of UNSM treatment on the reduction in fretting wear was evaluated by observing the wear characteristics of the untreated and UNSM-treated specimens. Fig. 7 shows the SEM images of the typical fretting wear scars generated in the untreated and UNSM-treated specimens after 110 6 fretting cycles at 20kHz and slip amplitude of 30 m. CP Ti and Ti6Al4V alloy images showed that the fretting wear scars for untreated and UNSM-treated specimens had a diameter of about 680 and 630 m and 520 and 480 m, respectively. The untreated specimens showed signicant evidence of wear debris formation on one side of the wear scar. The wear debris appeared to be adhered to the specimen surface, probably due to combined effects of shear and compressive stresses imparted by the counter surface. Also, as with fretting wear scars reported by Mohdtobi et al. [48] the stick and slip regions could be identied within the scars. One interesting point to note was that the general features of the fretting wear scar generated at a high frequency of 20kHz were quite similar to those observed at much lower frequencies reported in other works [49,50]. Though the SEM images gave qualitative information about the fretting wear behavior, quantication of wear was not possible from these images. Therefore, further surface analysis was conducted to quantify the fretting wear volume. The variation in fretting wear volume of the untreated and UNSM-treated specimens as a function of normal load was obtained as shown in Fig. 8. From the results, it was observed that the fretting wear volume increased as the normal load increased for both CP Ti and Ti6Al4V alloy specimens. The wear of UNSM-treated specimens

Fig. 6. Variation of the average friction coefcient of CP Ti and Ti6Al4V alloy specimens before and after the UNSM treatment.

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142

141

(a)

(b)

Fretting direction

Fretting direction 500 m 500 m

(c)

(d)
Fretting direction

Fretting direction 500 m 500 m

Fig. 7. SEM image of the fretting wear scars of the untreated CP Ti (a) and Ti6Al4V alloy (b) and UNSM-treated CP Ti (c) and Ti6Al4V alloy (d) specimens after the fretting test (1106 cycles at 20 kHz, slip amplitude of 30 m, normal load of 400 N).

was signicantly lower compared to those of the untreated specimens. Hence, it was conrmed that the fretting wear resistance of the UNSM-treated specimens was improved compared to the untreated specimens. This improvement may be attributed to the increased hardness and induced compressive residual stress of the UNSM-treated specimens. It could be summarized from this study that UNSM treatment was effective to mitigate fretting wear. The three mechanisms to enhance fretting resistance could be summarized as: (1) increased surface hardness; (2) induced compressive residual stress; (3) low friction coefcient. As for the surface roughness, it has been reported that to minimize the fretting wear rough surfaces are preferred [51]. However, in this study, the surface roughness value for all specimens was kept

identical. The generation of induced compressive residual stress in the surface layer by UNSM treatment is one of the most important phenomena to mitigate fretting wear. This observation is in accord with other surface treatment processes such as SP and ion-beam-enhanced deposition methods that are also useful in increasing the compressive residual stress at the surface [52,53]. Finally, the reduction in coefcient of friction can also improve the fretting wear resistance because of the decrease in the alternating tensile shear stresses.

4. Conclusions The effect of UNSM-treatment on the high-frequency fretting wear and friction characteristics of CP Ti and Ti6Al4V alloy was investigated. From the experimental results, the following conclusions may be drawn: By UNSM treatment, the CP Ti specimen grain size was rened from 35.5 m to 200nm, and for the Ti6Al4V alloy specimen and phases were rened from 9.9 m and 3.8 m to 1.2 m and 0.8 m, respectively. Surface hardness of CP Ti and Ti6Al4V alloy increased from 146 HV to 193 HV and from 328 HV to 379 HV, respectively, after the UNSM treatment. When the specimens were subjected to the UNSM treatment the highest compressive residual stress at the top surface of CP Ti and Ti6Al4V alloy reached up to 1279.4 and 1142.7 MPa, respectively. The UNSM-treated specimens showed an enhanced fretting wear resistance and low friction coefcient compared to those of the untreated specimens at higher loading rates. The reason for the improved fretting wear and frictional properties of the UNSM treated specimens was attributed to the increased hardness and compressive residual stress induced by the UNSM process.

Fig. 8. Variation in wear scar volume as a function of normal load for untreated and UNSM-treated specimen of CP Ti and Ti6Al4V alloy after 1106 fretting cycles at 20 kHz and a slip amplitude of 30 m.

142

A. Amanov et al. / Surface & Coatings Technology 207 (2012) 135142 [23] I.S. Cho, C.S. Lee, A. Amanov, Y.S. Pyoun, I.G. Park, J. Nanosci. Nanotechnol. 11 (1) (2010) 742. [24] A. Amanov, Y.S. Pyoun, B. Zhang, J.H. Park, J. Nohava, Tribol. Online 6 (2011) 284. [25] A. Amanov, Y.S. Pyoun, I.S. Cho, C.S. Lee, I.G. Park, J. Nanosci. Nanotechnol. 11 (2011) 701. [26] C.R. Ramos-Saenz, P.A. Sundaram, N. Diffoot-Carlo, J. Mech. Behav. Biomed. Mater. 3 (2011) 549. [27] R.S. Magaziner, V.K. Jian, S. Mall, Wear 1112 (2008) 1002. [28] S. Fouvry, C. Paulin, S. Deyber, Tribol. Int. 42 (2009) 461. [29] A. Shenhar, I. Gotmana, S. Radinb, P. Ducheyne, E.Y. Gutmanas, Surf. Coat. Technol. 126 (2000) 210. [30] D.M. Brunette, Titanium in Medicine: Material Science, Surface, Engineering, Biological Responses, and Medical Application, Springer, New York, NY, 2001. [31] J. Halling, Wear 4 (1961) 22. [32] R.C. Bill, ASLE Trans. 21 (1978) 236. [33] T. Kayaba, A. Iwabuchi, Wear 74 (1981) 229. [34] Y.S. Zhang, Z. Han, K. Lu, Wear 265 (2008) 396. [35] P.S. Prevey, In: Proceedings of IITT-International, 1990, p. 81. [36] K. Kubiak, S. Fouvry, A.M. Marechal, J.M. Vernet, Surf. Coat. Technol. 201 (7) (2006) 4323. [37] J. Zhao, L.C. Stearn, M.P. Harmer, H.M. Chan, G.A. Miller, R.E. Cook, J. Am. Ceram. Soc. 76 (1993) 503. [38] I.A. Chou, H.M. Chan, M.P. Harmer, J. Am. Ceram. Soc. 79 (1996) 2403. [39] H. Wu, S.G. Roberts, B. Derby, Acta Mater. 49 (2001) 507. [40] D. Johnson-Walls, A.G. Evans, D.B. Marshall, M.R. James, J. Am. Ceram. Soc. 69 (1986) 44. [41] H. Gleiter, Prog. Mater. Sci. 33 (1989) 223. [42] T. Inoue, Z. Horita, H. Somekawa, K. Ogawa, Acta Mater. 56 (2008) 6291. [43] Abu Fazal M. Arif, J. Mater. Process. Technol. 136 (2003) 120. [44] A. Molinari, T.B. Straffelini, T. Bacci, Wear 208 (1997) 105. [45] H. Dong, T. Bell, Wear 225229 (1999) 874. [46] B. van Peteghem, S. Fouvry, J. Petit, Wear 271 (2011) 1535. [47] H. Ji, L. Fia, X. Ma, Y. Sun, Wear 246 (2000) 40. [48] A.L. Mohdtobi, P.H. Shipway, S.B. Leen, Wear 271 (2011) 1572. [49] X. Huang, R.W. Neu, Wear 265 (2008) 971. [50] S. Soderberg, U. Bryggman, T. McCullough, Wear 110 (1986) 19. [51] Y. Fu, N.L. Loh, A.W. Batchelor, D. Liu, X. Zhu, J. He, K. Xu, Surf. Coat. Technol. 106 (1998) 193. [52] N.C. Horswill, K. Sridharan, J.R. Corad, J. Mater. Sci. Lett. 14 (1995) 1349. [53] R.B. Waterhouse, Wear 68 (1982) 310.

Acknowledgments This research was supported by a grant (No. 2011K000290) from the Center for Nanostructured Materials Technology under the 21st Century Frontier R&D Programs of the Ministry of Education, Science and Technology of Korea (MEST) and the National Research Foundation of Korea (NRF) grant funded by Korean Government (MEST) (No. 2011 0000409). References
[1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] R.B. Waterhouse, Int. Mater. Rev. 37 (1992) 77. R.A. Poggie, J.J. Wert, A.K. Mishra, J.A. Davidson, In: ASTM STP, 1145, 1992, p. 65, PA. K.G. Budinski, Wear 151 (1991) 203. P.D. Miller, J.W. Holladay, Wear 2 (1958/1959) 133. H. Hong, W.O. Winer, J. Tribol, Trans. ASME 111 (1989) 504. F.M. Kustas, M.S. Misra, In: ASM Handbook, 18, ASM International, Materials Park, Ohio, 1992, p. 778. R.B. Waterhouse, Fretting Fatigue, Elsevier Applied Science, London, 1981. D. Hoeppner, In: ASTM STP, 1159, 1992, p. 23. J. DeMasi-Marcin, D. Gupta, Surf. Coat. Technol. 6869 (1994) 1. A. Freimanis, A. Segall, J. Conway Jr., E. Whittney, Tribol. Trans. 45 (2002) 193. J.E. Barry, E.J. Tobin, P. Sioshansi, Surf. Coat. Technol. 51 (1992) 176. Z.D. Dai, S.C. Pan, M. Wang, S.R. Yang, X.S. Zhang, Q.J. Xue, Wear 213 (1997) 135. V. Fridrici, S. Fouvry, Ph. Kapsa, Wear 250 (2001) 642. S. Srinivasan, D.B. Garcia, M.C. Gean, H. Murthy, T.N. Farris, Tribol. Int. 42 (2009) 1324. M.M. Ali, S.G. Raman, S.D. Pathak, R. Gnanamoorthy, Tribol. Int. 43 (2010) 152. B. Prakash, E. Richter, H. Pattyn, J.P. Celis, Surf. Coat. Technol. 173 (2003) 150. S.A. Kumar, S.G. Raman, T.S.N. Narayanan, R. Gnanamoorthy, Adv. Mater. Res. 463464 (2012) 316. L. Shaw, Y.T. Zhu, In: in: J. Groza, J. Shackelford, E. Lavernia, M. Powers (Eds.), CRC Materials Processing Handbook, CRC Press, Boca Raton, FL, 2007, p. 311. L. Wagner, Mater. Sci. Eng., A 263 (1999) 210. I. Altenberger, B. Scholtes, U. Martin, H. Oettel, Mater. Sci. Eng., A 264 (1999) 1. W. Cao, R. Fathallah, L. Castex, Mater. Sci. Technol. 11 (1995) 967. A. Amanov, Y.S. Pyoun, I.S. Cho, C.S. Lee, I.G. Park, Wear 286287 (2012) 136.

Você também pode gostar