Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of...

11
J-STAGE Advance Publication date: 5 July, 2018 Paper No.17-00523 © 2018 The Japan Society of Mechanical Engineers [DOI: 10.1299/mej.17-00523] Vol.5, No.4, 2018 Bulletin of the JSME Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and Yoshiki KOMIYA* *Department of Aerospace Engineering, Nihon University 7-24-1, Narashinodai, Funabashi, Chiba 274-8501, Japan E-mail: [email protected] Abstract Titanium and its alloys have excellent strength properties and corrosion resistance, but they show poor wear resistance, which cannot be improved by heat treatment. The addition of hard ceramic particles to produce a titanium matrix composite is an effective method for enhancing the wear resistance of titanium and its alloys. In this study, TiC and TiB2 were used as reinforcement materials because a composite containing these reinforcements shows high tensile strength and different microstructure. TiB- and TiC-reinforced pure Ti matrix composites were fabricated by using the spark plasma sintering (SPS) method. Dry sliding wear tests were conducted on the composites using a ball-on-disk type testing machine. The effects of the reinforcement material, microstructural features, and reinforcement volume fraction on the wear behavior of the composites were investigated. The specific wear rates of both composites decreased with increasing reinforcement volume fraction. The specific wear rate of the TiC/Ti composite deceased drastically at a reinforcement volume fraction of 25 vol.%. TiC/Ti composites with reinforcement volume fractions of over 5 vol.% showed excellent wear resistance compared with the TiB/Ti composites. The wear behavior of the composite depended mainly on the distribution of reinforcement material and the nature of the reaction products between the matrix and reinforcement particles. 1. Introduction Titanium and its alloys exhibit low density, good corrosion resistance, high specific strength, and excellent fatigue strength. For years these materials have been considered suitable for aerospace applications; however, they are increasingly being considered for commercial applications, such as automotive, biomedical, and consumer applications. On the other hand, titanium and its alloys show poor wear resistance, which has hindered the widespread adoption of these materials. For titanium and its alloys, it is difficult to improve the wear resistance by subjecting them to heat treatment. However, the addition of ceramic particles into titanium and its alloys is an effective method to improve the wear resistance. Some reports on the dry sliding wear behavior of titanium matrix composites (TMC) with pure Ti or titanium alloys have been published (Threrujirapapong, et al., 2008, Chang, et al., 2013, Luo, et al., 2014, Rasool, et al., 2015) . Most of these studies used Ti-6Al-4V alloy as a matrix material (Alman, et al., 1999, Alma, et al., 2002, Candel, el at., 2010, Hu, el at., 2010, Mahamood, et al., 2013, Xie, et al., 2014). Particulate-reinforced titanium matrix composites exhibit not only improved wear resistance but also enhanced specific mechanical properties. In generally, the wear resistances of titanium matrix composites improve with increasing reinforcement volume fraction. For TiB/Ti- 6Al-4V alloy matrix composites, the highest reported tensile strength was obtained at 5 vol.% TiB (Sanpei, et at., 2015). For TiB/Ti composites, the highest reported tensile strength was observed at 20 vol.% TiB, and this tensile strength was higher than the tensile strength of TiB/Ti-6Al-4V alloy composites containing 20 vol.% TiB (Izui, et at., 2016). In other words, the reinforcement volume fraction that gives the maximum tensile strength depends on the matrix material. Therefore, pure titanium matrix composites with high reinforcement volume fraction can exhibit enhanced wear Keywords : Titanium matrix composite, TiB, TiC, Spark plasma sintering, Dry sliding wear, Ball-on-disk tests, Wear and friction properties, Specific wear rate 1 Received: 6 October 2017; Revised: 28 May 2018; Accepted: 26 June 2018

Transcript of Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of...

Page 1: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

J-STAGE Advance Publication date: 5 July, 2018Paper No.17-00523

© 2018 The Japan Society of Mechanical Engineers[DOI: 10.1299/mej.17-00523]

Vol.5, No.4, 2018Bulletin of the JSME

Mechanical Engineering Journal

Dry sliding wear behavior of TiB/Ti and TiC/Ti composites

Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and Yoshiki KOMIYA* *Department of Aerospace Engineering, Nihon University

7-24-1, Narashinodai, Funabashi, Chiba 274-8501, Japan

E-mail: [email protected]

Abstract

Titanium and its alloys have excellent strength properties and corrosion resistance, but they show poor wear

resistance, which cannot be improved by heat treatment. The addition of hard ceramic particles to produce a

titanium matrix composite is an effective method for enhancing the wear resistance of titanium and its alloys. In

this study, TiC and TiB2 were used as reinforcement materials because a composite containing these

reinforcements shows high tensile strength and different microstructure. TiB- and TiC-reinforced pure Ti matrix

composites were fabricated by using the spark plasma sintering (SPS) method. Dry sliding wear tests were

conducted on the composites using a ball-on-disk type testing machine. The effects of the reinforcement material,

microstructural features, and reinforcement volume fraction on the wear behavior of the composites were

investigated. The specific wear rates of both composites decreased with increasing reinforcement volume

fraction. The specific wear rate of the TiC/Ti composite deceased drastically at a reinforcement volume fraction

of 25 vol.%. TiC/Ti composites with reinforcement volume fractions of over 5 vol.% showed excellent wear

resistance compared with the TiB/Ti composites. The wear behavior of the composite depended mainly on the

distribution of reinforcement material and the nature of the reaction products between the matrix and

reinforcement particles.

1. Introduction

Titanium and its alloys exhibit low density, good corrosion resistance, high specific strength, and excellent fatigue

strength. For years these materials have been considered suitable for aerospace applications; however, they are

increasingly being considered for commercial applications, such as automotive, biomedical, and consumer applications.

On the other hand, titanium and its alloys show poor wear resistance, which has hindered the widespread adoption of

these materials. For titanium and its alloys, it is difficult to improve the wear resistance by subjecting them to heat

treatment. However, the addition of ceramic particles into titanium and its alloys is an effective method to improve the

wear resistance. Some reports on the dry sliding wear behavior of titanium matrix composites (TMC) with pure Ti or

titanium alloys have been published (Threrujirapapong, et al., 2008, Chang, et al., 2013, Luo, et al., 2014, Rasool, et al.,

2015) . Most of these studies used Ti-6Al-4V alloy as a matrix material (Alman, et al., 1999, Alma, et al., 2002, Candel,

el at., 2010, Hu, el at., 2010, Mahamood, et al., 2013, Xie, et al., 2014). Particulate-reinforced titanium matrix

composites exhibit not only improved wear resistance but also enhanced specific mechanical properties. In generally,

the wear resistances of titanium matrix composites improve with increasing reinforcement volume fraction. For TiB/Ti-

6Al-4V alloy matrix composites, the highest reported tensile strength was obtained at 5 vol.% TiB (Sanpei, et at., 2015).

For TiB/Ti composites, the highest reported tensile strength was observed at 20 vol.% TiB, and this tensile strength was

higher than the tensile strength of TiB/Ti-6Al-4V alloy composites containing 20 vol.% TiB (Izui, et at., 2016). In other

words, the reinforcement volume fraction that gives the maximum tensile strength depends on the matrix material.

Therefore, pure titanium matrix composites with high reinforcement volume fraction can exhibit enhanced wear

Keywords : Titanium matrix composite, TiB, TiC, Spark plasma sintering, Dry sliding wear, Ball-on-disk tests, Wear and friction properties, Specific wear rate

1

Received: 6 October 2017; Revised: 28 May 2018; Accepted: 26 June 2018

Page 2: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

resistance and strength properties at the same time.

In this study, pure Ti was used as a matrix material, and TiB2 and TiC particles were used as reinforcement materials,

because TiB- and TiC-reinforced titanium composites exhibit excellent tensile strengths and different microstructural

features (Izui, et al., 2016, Tsang, H. T., et al., 1996). The composites were fabricated by the spark plasma sintering (SPS)

method. TiB2 reacts with Ti to form TiB whiskers and its clusters during the sintering process. TiC reacts with Ti to form

a thin interface layer of Ti2C around TiC particles (Wanjara, P., et al., 2000, Vallauri, D., et al.,2008). The wear resistance

of TiB- and TiC-reinforced titanium matrix composites is expected to be favorable due to the very high Vickers hardness

of TiB and TiC. Dry sliding wear tests were conducted with the ball-on-disk method, because it is easy to make balls

perfectly contact the wear test specimen compared with the pin-on-disk method. This study investigated the effects of

the reinforcement materials, the microstructural features, and the reinforcement volume fraction on the wear and friction

behavior of TiB/Ti and TiC/Ti composites.

2. Experimental

2.1 Starting powders

Figure 1 shows SEM micrographs of the matrix and reinforcement powders. Gas-atomized (GA) pure Ti was used

as a matrix material. The particle size and oxygen content of Ti particles were less than 45 μm and 0.106 wt%,

respectively. Two types of reinforcement particles were used: TiB and TiC. The particle sizes of both materials were

1–2 μm.

(a) Pure Ti (Matrix) (b) TiB2 (c) TiC

Fig.1 SEM micrographs of pure Ti (matrix), TiB2, and TiC particles (reinforcements)

2.2 Composite fabrication process and microstructural analysis

Pure Ti particle was mixed with reinforcement particles using a planetary ball-mill at a rotation speed of 200 rpm

for 10 min. Then, the mixed particles were filled into a graphite die, as shown in Fig. 2. The filled particles were pre-

compacted under a pressure of 20 MPa with a hand press machine. The composites were prepared using a spark plasma

sintering machine (Dr. Sinter, SPS-3.20IV, Sumitomo Coal Mining Ltd., Japan). Table 1 shows the composite

fabrication conditions (Kamegawa, S., et al., 2015).

SEM observation and EDX mapping analysis of the composite samples were conducted using a scanning electron

microscope (Shimadzu SSX-550, Japan) . The Vickers microhardness of the samples was measured with a Shimadzu

Heating rate 20 °C/min (212 °F/min)

Sintering temperature 900 °C (1652 °F)

Sintering time 10 min

Sintering pressure 70 MPa

Pressure < 5 Pa

Cooling method Furnace cooling

Table 1 Composite fabrication conditions

Fig.2 Graphite die and punches

Composite fabrication conditions

2

Page 3: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

Microhardness tester HMV-2(T) with 9.8 N load and a dwell time of 30 s. The average microhardness for 10 indents

was taken for each composite specimen.

2.3 Dry sliding wear test

Dry sliding wear test of the composites were carried out using a three ball-on-disk machine. A high-carbon-

chromium (100Cr6) steel ball bearing with a diameter of 10 mm was employed as a counterpart material. The wear

tests were conducted at a constant sliding velocity of 100 mm/s and a sliding distance of 500 m at a constant load of 23

N. The composite surfaces were polished with 1500 grit emery paper followed polishing with 3 μm diamond grains to

achieve a uniform roughness. The average surface roughness (Ra) of the composites was less than 0.2 μm. Contact

between the composite sample and the balls was checked using pressure-sensitive paper.

The specific wear rate was calculated by the following formula:

Ws = (W1-W2) / PLρ (1)

where Ws is the specific wear rate, W1 is the sample mass before the test, W2 is the sample mass after the test, P is the

applied load, L is the sliding distance, and ρ is the sample density. Wear track profiles of composite surface were

measured with a digital microscope (Olympus DSX510, Japan). Furthermore, the surfaces of the composite before the

wear test and the wear tracks after the wear test were examined by X-ray diffraction analysis (XRD) (Rigaku Rint 2000,

Japan) with Cu-Kα radiation (50 kV and 100 mA).

3. Results and discussion

3.1 Microstructure and Vickers microhardness of the composites

Figure 3 and 4 present the microstructure and EDX mapping analysis of the composites containing TiB2 and TiC

reinforcements. TiB2 reacted with Ti to form TiB whiskers in the Ti particles, TiB clusters at the interfaces between them

during the sintering process. A number of fine needle-shaped TiB whiskers (TiBw) were formed in the Ti particles. From

Fig. 3 (b) and (c), the TiB whiskers and their clusters showed a high B content. Therefore, the microstructure of the

TiB/Ti composite consisted of Ti phases containing TiB whiskers and TiB clusters. For the composite containing the

TiC reinforcement, TiC particles were discretely distributed in the Ti matrix phase. TiC reacted with Ti to form a thin

layer of Ti2C around the TiC particles.

The microstructures of the TiB/Ti composite containing 15 and 25 vol.% reinforcements are shown in Fig. 5. These

microstructures formed a network of TiB whiskers within the Ti matrix. The local volume fractions of TiB

TiB clusters

(a) SEM micrograph (b) Ti (c) B

Fig. 3 SEM micrograph and EDX mapping analysis of TiB/Ti composite

(a) SEM micrograph (b) Ti (c) C

Fig. 4 SEM micrograph and EDX mapping analysis of TiC/Ti composite

TiBw

Ti

TiC

3

Page 4: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

reinforcements at the network boundary increased with increasing overall volume fraction.

Figure 6 shows the Vickers microhardnesses of the composites. The Vickers microhardnesses of both composites

increased with increasing reinforcement volume fraction. The Vickers microhardnesses of the TiB/Ti composites were

higher than those of the TiC/Ti composites. A sharp increase in the Vickers microhardnesses of the TiB/Ti composites

was observed compared with the TiC/Ti composites because the number of TiB whiskers in the Ti phases increased with

increasing TiB volume fraction. Another reason for the sharp increase in the microhardness was due tothe high oxygen

content in TiB2 compared with TiC. The oxygen contents of the TiB2 and TiC were <2.0 wt.% and <0.8 wt.%,

respectively. As the TiB volume fraction increased, the oxygen content of the Ti phases increased. Therefore, the

presence of oxygen in the Ti phases resulted in an increase in the Vickers microhardness of the composite containing the

TiB reinforcement.

(a) 15 vol.% TiB (b) 25 vol.% TiB

Fig. 5 SEM micrographs of TiB/Ti composites containing different TiB volume fractions.

Fig. 6 Variation of Vickers microhardness of the composites with reinforcement volume fraction.

3.2 Dry sliding wear behavior of the composites

Figure 7 shows how the specific wear rate of the composites depended on the reinforcement volume fraction. For

the TiC/Ti composites, the specific wear rate decreased gradually as the TiC volume fraction increased to 15 vol.%, and

then decreased sharply at 25 vol.%. On the other hand, for the TiB/Ti composite, a high specific wear rate was

maintained up to a TiB volume fraction of 15 vol.%, and then decreased sharply over 15 vol.%. For the composites

containing the reinforcement volume fraction of 15 vol.% or more, the specific wear rates of the TiC/Ti composites

exhibited smaller than those of the TiB/Ti composites.

Wear depth profiles of pure Ti and the composites are shown in Fig. 8. The wear depth of the composites decreased

with increasing reinforcement volume fraction. The wear depth of the TiB/Ti composite was deeper than that of the

TiC/Ti composite. For the composite containing 25 vol.% TiC, the composite surface was hardly worn away but was

maintained the same level as the original surface. From the results in Figs. 7 and 8, the TiC/Ti composites showed good

wear resistance compared with the TiB/Ti composites. Figure 9 shows the SEM and EDX mapping images of the wear

track of pure Ti and the composites containing TiC and TiB. For the pure Ti, surface asperities with long grooves on the

wear surface were observed. The protruding area of the wear track contained a certain amount of oxygen, as shown in

TiB whisker

TiC

200

300

400

500

600

0 10 20 30

Reinforcement volume fraction, vol.%

Vic

ker

s m

icro

hard

nes

s, H

V

TiC/Ti

TiB/Ti

4

Page 5: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

Fig. 7 Variation of specific wear rate of the composites with reinforcement volume fraction.

(a) Pure Ti and TiC/Ti composites containing 15 and 25 vol.% TiC

(b) TiB/Ti composites containing 15, 25, and 35 vol.% TiB

Fig. 8 Wear depth profiles of pure Ti and composites with different reinforcement volume fractions.

Fig. 9 (a). On the other hand, for the TiC/Ti composites, both wear tracks consisted of flat and rough surfaces. The

flat surface area increased with increasing TiC volume fraction. For 15 vol.% TiC, the flat surface regions contained

Ti, O, and Fe. For 25 vol.% TiC, they showed high O and Fe contents.

From Fig. 9 (b), the wear surface of the composite containing 15 vol.% TiB exhibited the presence of deep grooves

35 vol.% TiB

25 vol.% TiB

15 vol.% TiB

25 vol.% TiC

15 vol.% TiC

Pure Ti

TiC/Ti

TiB/Ti

0 vol.% TiC

(Pure Ti)

TiC/Ti

TiB/Ti

5

Page 6: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

Sliding direction

Sliding direction

(a) Pure Ti and composites containing 15 and 25 vol.% TiC

(b) Composites containing 15, 25, and 35 vol.% TiB

Fig. 9 SEM and EDX mapping images of wear tracks in pure Ti and composites containing TiB and TiC.

6

Page 7: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

along the sliding direction. For the 25 and 35 vol.% TiB, deep grooves were no longer seen on the wear surface, and its

topography was fairly smooth. The protruding regions on the wear track also showed a certain amount of oxygen. The

wear surface of the composite containing 25 vol.% TiB consisted of gray and white phases. The gray phase exhibited

higher oxygen and iron contents than the white phase. For 35 vol.% TiB, the wear track contained uniformly distributed

oxygen. Therefore, the wear track was covered with a Ti oxide layer.

Figure 10 shows the XRD patterns of the composite surfaces containing 15 and 25 vol.% reinforcement before and

after the wear test. As shown in Fig.10 (a), for the composite containing TiC, the diffraction peaks of TiO and FeO were

detected, and the intensities of the peaks increased with increasing the TiC volume fraction. On the other hand, from

Fig.10 (b), for the composite containing TiB, diffraction peaks of TiO and FeO were hardly observed.

3.3 Discussion

Here we discuss the wear behavior of the TiB/Ti and TiC/Ti composites. Figure 11 shows friction coefficient curves

of pure Ti and composites containing 15 and 25 vol.% reinforcements. Wear surfaces, cross-sections of the wear tracks,

and wear debris of the composites containing 15 and 25 vol.% reinforcements are shown in Figs. 12 and 13. Both

composites showed short initial low-friction regions, and then the friction coefficients increased rapidly. Finally, the

friction coefficients exhibited high values between 0.45 and 0.65. In the low-friction regions, the steel balls sliding on

the composite hardly generated any wear debris. The rapid increase of the friction coefficient was due to the generation

of wear debris including the TiC and TiB reinforcements. The composite containing TiC exhibited a shorter initial low-

friction region than the composite containing TiB. The composite containing TiC showed lower tensile strength than the

composite containing TiB ranging from 5 to 25 vol.% (Kamegawa, S., et al., 2015). For a titanium matrix composite,

at the initial stage of the plastic deformation process under the applied load, the increase in the load carried by the

reinforcement is mainly a result of the progressive strain hardening of the matrix; therefore, particle fractures occur in

the critical reinforcements, such as large particles, clusters of particles, and so forth (da Silva, A. A. M., el. at., 2006,

Segurado, J., el. at., 2003). It can be observed from Fig. 12 (b) that perfectly unsintered TiC particles, which were

(a) Composite containing 15 and 25 vol.% TiC

(b) Composite containing 15 and 25 vol.% TiB

Fig.10 XRD patterns of the composites containing 15 and 25 vol.% of TiB and TiC before and after wear test

7

Page 8: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

weakly bonded, were agglomerated. Furthermore, the interfaces between Ti and TiC particles were weakly bonded due

to the formation of Ti2C (Wanjara, P., et al., 2000). Therefore, the TiC particles were easily removed from the wear

surface. On the other hand, as mentioned earlier, the TiB2 reacted with Ti to form the TiB clusters at the interface between

the Ti particles. The TiB clusters were also fractured due to fatigue caused by the load of the sliding balls. However, the

formation of large TiB clusters made it more difficult to fracture the TiB clusters and to remove them from the wear track.

Therefore, the low-friction regions for the composites containing TiB were longer than those for the composites

containing TiC. From Fig. 11 (a), at the high friction coefficient region, there was a large variation in the friction

coefficient of the composite containing TiC compared to the composite containing TiB and pure Ti: the friction coefficient

of the composite containing TiC fluctuated between 0.45 and 0.65. As shown in Fig.9, the wear tracks of the composites

containing 15 vol.% reinforcement exhibited the abrasive wear behavior with the grooves along the sliding direction. The

abrasion grooves were formed by the reinforcement debris removed from the wear track. The groove width of the

composite containing TiB is larger than that of the composite containing TiC, because the size of the removed TiB clusters

was larger than that of the TiC particles. The groove width decreased with increasing a contact area between the balls

and the wear track. Therefore, the contact area of the composite containing TiC is larger than that of the composite

containing TiB. The sliding resistance of the balls increases with increasing the contact area. Furthermore, many TiC

particles protruded from the wear track can be observed as shown in Fig.12. The presence of the many TiC particles

protruded from the wear track results in the increase of the sliding resistance. The large variation of the friction coefficient

of the composite containing 15 vol.% TiC is due to the large contact area and the presence of the protruding TiC particles.

For 25 vol.% reinforcement, both composites also showed a short initial low-friction region due to the sliding balls

(b) Pure Ti and the composite containing 25 vol.% reinforcement

Fig.11 Friction coefficient curves of pure Ti and the composites containing 15 and 25 vol.% reinforcement.

(a) Pure Ti and the composite containing 15 vol.% reinforcement

8

Page 9: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

on the composite generating little wear debris including reinforcements, and then the friction coefficients increased

rapidly due to the generation of wear debris. For the Ti composite containing TiC, the friction coefficient reached a

maximum and decreased due to the formation of flat transfer layers of compacted wear debris. For the Ti composite

containing TiB, in the steady-state after the sudden increase of the friction coefficient, the friction coefficient of the

composite containing 25 vol.% TiB was slightly higher than that of the composite containing 15 vol.% TiB.

Though the Vickers microhardness of the composites containing TiC was lower than that of the composite

containing TiB, but The composites containing over 15 vol.% TiC exhibited good wear resistance compared with the

composites containing TiB. As motioned before, the fracturing of the agglomerated TiC particles and TiB clusters

resulted in a sudden increase in the friction coefficients. As shown in Figs. 12(c) and 13 (c), the wear debris of the

composite containing TiB was larger than that of the composite containing TiC. This means that the wear debris of

the composite containing TiB included the fractured TiB clusters. The fracture and removal of the TiB clusters resulted

in the formation of deeper ploughed grooves in the wear tracks due to the abrasive wearing. In addition, as shown in

Fig. 13 (b), the white lines indicate the wear-damaged composite surfaces. The wear-damaged surface of the composite

containing TiC was smoother than that of the composite containing TiB. The main reason for this phenomenon is the

fracturing and removal of the TiB clusters.

In the cross-section of the composite containing 15 vol.% reinforcement, as shown in Fig. 12 (b), few thin transfer

layers can be observed on the worn surface. For the composites containing 25 vol.% reinforcement, the wear surface was

partly covered with the thin transfer layers. The wear surface, cross-section and wear debris of the composites containing

35 vol.% TiB are shown in Fig. 14. The white lines also indicate the wear-damaged composite surface. As shown in

Fig. 14 (a) and (b), the wear surface with small asperities was covered with a transfer layer of compacted wear debris.

The transfer layer thickness of the composite containing TiB increased with increasing TiB volume fraction. This means

that the TiB clusters fractured and were removed from the surface of the composite containing a high volume fraction of

TiB. The wear debris of the composite containing 35 vol.% TiB, which was fractured during sliding, was smaller than

that of the composites containing 15 and 25 vol.% TiB. As shown in Fig .14 (b), many voids were observed in the TiB

clusters. Therefore, the fine wear debris was fractured and removed from the wear surface. The presence of fine wear

debris enabled the formation of a thick transfer layer. For these reasons, the specific wear rate of the composite

containing TiC was lower than that of the composite containing TiB. In other words, the composite containing TiC

exhibited superior wear resistance compared with the composite containing TiB.

(a) Wear surface (b) Cross-section (c) Wear debris

15

TiC

/Ti

15

TiB

/Ti

Fig. 12 Wear surfaces and cross-sections of wear tracks, and wear debris of the TiC/Ti and TiB/Ti composites

containing 15 vol.% reinforcement.

Transfer layer

Protruding TiC

particles

Ploughed grooves

Ti TiC

Protruding TiC particles

Fractured TiB clusters

Ti+TiBW TiB

TiB cluster

Steel ball

Steel ball

9

Page 10: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

As shown in Fig. 9, for the composite containing 25 vol.% TiC, O- and Fe-rich phases, which were formed from

the compacted wear debris, can be observed. The composite containing 25 vol.% TiC was worn only a little, as indicated

by the fact that the wear track was almost the same level as the original composite surface, as can be seen in Fig. 8 (a).

Therefore, the steel balls slid on and contacted the same surface of the wear track, and they were worn to form the O-

and Fe-rich phases on the wear track. In contrast to the composite containing 35 vol.% TiB, a small amount of Fe was

detected from the wear track. Fine wear debris was generated from its surface, with a maximum wear depth of

approximately 20 μm. Therefore, for the composite containing TiB, after the wear debris was generated, the cycle of

removal and compaction of the wear debris repeated in the wear test. Since the contact time between the steel balls and

the wear debris or the worn surface was short, the steel balls were hardly worn.

(a) Wear surface (b) Cross-section (c) Wear debris

25

TiC

/Ti

25

TiB

/Ti

Fig. 13 Wear surfaces and cross-sections of wear tracks, and wear debris of TiC/Ti and TiB/Ti composites

containing 25 vol.% reinforcement.

(a) Wear surface (b) Cross-section (c) Wear debris

Fig. 14 Wear surface and cross-section of wear track, and wear debris of TiB/Ti composite containing 35 vol.%

reinforcement.

4. Conclusion

The effect of the reinforcement material, microstructural features, and reinforcement volume fraction on the wear

and friction behavior of TiB/Ti and TiC/Ti composites was studied. The results can be summarized as follows:

Ti-O-Fe phase

O and Fe rich phase

(Transfer layer)

Ti-rich phase

Ti-O phase

(Transfer layer)

Ti+TiBw

TiB

Transfer layer

Transfer layer

Ti

TiC

Transfer layer

TiB

Ti+TiBw

10

Page 11: Mechanical Engineering Journal€¦ · Mechanical Engineering Journal Dry sliding wear behavior of TiB/Ti and TiC/Ti composites Hiroshi IZUI*, Kazuhiro TOEN*, Shoji KAMEGAWA* and

2© 2018 The Japan Society of Mechanical Engineers

Izui, Toen, Kamegawa and Komiya, Mechanical Engineering Journal, Vol.5, No.4 (2018)

[DOI: 10.1299/mej.17-00523]

1. The specific wear rate of the composites containing TiC was lower than that of the composites containing TiB. The

composites containing TiC exhibited superior wear resistance compared to the composites containing TiB. The composite

containing 25 vol.% TiC was hardly worn away. The composite containing 35 vol.% TiB showed improved wear

resistance.

2. The wear depths of the composites surface decreased with increasing reinforcement volume fraction. The wear

surface of the composite containing 25 vol.% TiC were maintained at the same level as the unworn surface.

3. The wear behavior of the matrix composites depended on the microstructural properties, such as the distribution of

reinforcement material and the nature of the reaction products between Ti and the reinforcement particles.

4. The friction coefficients of both composites increased rapidly due to the generation of wear debris. The friction

coefficient was affected by the conditions at the wear surface and the constituent elements of the transfer layers.

References

Alam, M. O., and Haseeb, A.S.M.A., Response of Ti-6Al-4V and Ti-24Al-11Nb Alloys to Dry Sliding Wear against

Hardened Steel, Tribology International, Vol.35, No. 6(2002), pp.357-362.

Alman, D.E., and Hawk, J.A., The Abrasive Wear of Sintered Titanium Matrix – Ceramic Particle Reinforced

Composites, Wear, Vol.225-229, No.1(1999), pp.629-639.

Candel, J.J., Amigo, V., Ramos, J.A., and Busquets, D., Sliding Wear Resistance of TiCp Reinforced Titanium Composite

Coating Produced by Laser Cladding, Surface & Coatings Technology, Vol.204, No.10(2010), pp.3161-3166.

Chang, J., Yu, Y., Fu, L., Li, F., Qiao, Z., Li, J., Yang, J., and Liu, W., Effect of TiB2 on Dry-sliding Tribological Properties

of TiAl Intermetallics, Tribology International, Vol.62, No.1(2013), pp.91-99.

da Silva, A.A.M, dos Santos, J. F., and Strohaecker, T. R., An investigation of the fracture behavior of diffusion-bonded

Ti6Al4V/TiC/10p, Composites Science and Technology, Vol.66, No.13(2006), pp.2063-2068.

Hu, R.H., and Lim, J.K., Hardness and Wear Resistance Improvement of Surface Composite Layer on Ti-6Al-4V

Substrate Fabricated by Powder sintering, Materials and Design, Vol.31, No.1(2010), pp.2670-2675.

Izui, H., Oota, A., Matsuura, K., and Kamegawa, S., Tensile Behavior of TiB-reinforced Ti matrix Composites with

Different Titanium Powders, Mechanical Engineering Journal, Vol.3, No. 4 (2016), pp.1-8.

Kamegawa, S., Izui, H., Komiya, Y., Kobayashi, K., and Arimori, R., Effect of matrix and reinforcement powder types

on tensile strength and wear properties of TiB/Ti and TiC/Ti composites prepared by SPS, TMS2015 Supplemental

Proceedings, (2015), pp.203-210.

Luo, S. D., Yang, Y. F., Schaffer, G. B., and Qian, M., Warm die compaction and sintering of titanium and titanium ally

powders, Journal of Materials Processing Technology, Vol. 214, No.3 (2014), pp.660-666.

Mahamood, R.M., A, E.T., S, M., and P, S., Scanning Velocity Influence on Microstructure, Microhardness and Wear

Resistance Performance of Laser Deposited Ti6Al4V/TiC composite, Materials and Design, vol.50, No.1(2013),

pp.656-666.

Rasool, G., Mridha, S., and Stack, M.M., Mapping Wear mechanisms of TiC/Ti Composite Coatings, Wear, Vol.328-329,

No.1(2015), pp.498-508.

Sanpei, Y., Izui, H., Komiya, Y., Shigimura, H., and Suzuki, C., Mechanical Properties of TiB-Reinforced Ti-6Al-4V

Using Matrix Powders with Different Particle Sizes and Morphologies, TMS 2015 144th Annual Meeting &

Exhibition, pp.219-226.

Segurado, j., SeGonzalez, C. and Llorca, J., A numerical investigation of the effect of particle clustering on the mechanical

properties of composites, Acta Materialia, Vol. 51, No.8 (2003), pp.2355-2369.

Threrujirapapong, T., Kondou, K., Umeda, J., and Imai, H., Friction and Wear Behavior of Titanium Matrix Composite

Reinforced with Carbon Nanotubes under Dry Conditions, Transactions of JWRI, Vol.37, No.2(2008), pp.51-56.

Tsang, H. T., Chao, C. G., and Ma, C. Y., In-situ Fracture Observation of a TiC/Ti MMC Produced by Combustion

Synthesis, Scripta Materialia, Vol. 35, No. 8(1996), pp. 1007-1012.

Vallauri, D., Adrian, I.C.A., and Chrysanthou, A., TiC-TiB2 Composites: A Review of Phase Relationships, Processing

and Properties, Journal of the European Ceramic Society, Vol.28, No.8 (2008), pp.1697-1713.

Wanjara, P., Drew, R. A. L., Root, J., and Yue, S., Evidence for Stable Stoichiometric Ti2C at the Interface in TiC

Particulate Reinforced Ti Alloy Composites, Acta Materialia, Vol. 48, No.7 (2000), pp. 1443-1450.

11