Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure

Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure

Accepted Manuscript Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure N.K. Fukumasu, C.F. Bernardes, M.A...

12MB Sizes 0 Downloads 18 Views

Accepted Manuscript Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure N.K. Fukumasu, C.F. Bernardes, M.A. Ramirez, V.J. Trava-Airoldi, R.M. Souza, I.F. Machado PII:

S0301-679X(18)30193-2

DOI:

10.1016/j.triboint.2018.04.006

Reference:

JTRI 5185

To appear in:

Tribology International

Received Date: 29 December 2017 Revised Date:

31 March 2018

Accepted Date: 7 April 2018

Please cite this article as: Fukumasu NK, Bernardes CF, Ramirez MA, Trava-Airoldi VJ, Souza RM, Machado IF, Local transformation of amorphous hydrogenated carbon coating induced by high contact pressure, Tribology International (2018), doi: 10.1016/j.triboint.2018.04.006. This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

ACCEPTED MANUSCRIPT

2

Local Transformation of Amorphous Hydrogenated Carbon Coating induced by High Contact Pressure

3

N.K. Fukumasua,∗, C.F. Bernardesa , M.A. Ramirezb , V.J. Trava-Airoldib , R.M. Souzaa , I.F. Machadoa

5

Abstract

SC

The development of coatings presenting adaptive characteristics based on mechanical loads would promote improved local characteristics. In this work, the surface modification of an a-C:H coating was analyzed based on the contact pressure using two tribological tests: lubricated reciprocating ball-on-plane and dry micro-scratch tests. Local transformation were evaluated based on Raman spectroscopy and indentation hardness. Results indicated regions presenting a red-shift of the G band and higher indentation hardness. High compressive stress field developed under the contacting region (numerical simulations) coupled to the red-shift of the G band and the increase on indentation hardness inside the tested regions are compatible with the nucleation of sp3 carbon bond sites derived from sp2 bonds,

EP

TE

D

indicating a possible modification of the coating.

AC C

6

Phenomena Laboratory, Polytechnic School of the University of Sao Paulo, Sao Paulo, Brazil b National Institute for Space Research, Brazil

RI PT

a Surface

4

M AN U

1

author. Tel.: +55-11-3091-9865 Email address: [email protected] (N.K. Fukumasu )

∗ Corresponding

Preprint submitted to Elsevier

April 10, 2018

ACCEPTED MANUSCRIPT

7

1. Introduction Automotive and aerospace industries present several examples where an improvement in performance is required

9

to maximize energy efficiency and durability. According to Holmberg et al. [1], on regular passenger vehicles, power

10

generation and transmission systems present total mechanical losses due to friction responsible for approximately 28%

11

of fuel consumption. In this context, several alternatives may be applied to reduce friction losses, such as the use of

12

low friction coatings and improved surface finishing, reducing surface roughness or developing surface topographies

13

to improve lubrication conditions.

RI PT

8

The development and use of low friction coatings became an attractive technology with the use of modern reactors

15

that employ Plasma Enhanced Chemical Vapor Deposition (PECVD) techniques to produce carbon based coatings on

16

regular or complex geometries [2, 3, 4], such as engine taps, helical gears and bearings. High cyclic loads and relative

17

velocities are achieved during the operation of power train components. Those conditions may lead to significant wear

18

and failure of the coatings, decreasing the overall performance of the system. The understanding of coating behavior

19

under extreme conditions is vital to prevent this scenario, and tribological tests are usually conducted before coating

20

the real component. These tests allow exploring issues such as coating/substrate interaction under tribological loads.

M AN U

SC

14

Usually, carbon-based coatings present satisfactory performance in tribological applications, in which high hard-

22

ness and low coefficient of friction (COF) are commonly required [2, 3, 5, 6, 7, 8]. Literature reports an increasing use

23

of diamond like carbon (DLC) coatings, in which tribological properties can be tuned depending on the environment.

24

Under dry sliding condition, DLC coated systems may present a reduction of friction force based on the graphitization

25

of the contacting surfaces, as observed by Liu et al. [9]. This phenomenon is related to the rearrangement of the sp3

26

and sp2 carbon bonds by energy transferred from the mechanical movement to chemical bond kinetics. Polaki et al.

27

[10] reported the influence of high contacting surfaces relative speed of the reduction of COF and wear resistance

28

by the conversion of the sp3 to sp2 bonds. This change on the chemical bonds indicates the formation of a graphitic

29

phase that presents lower COF, due to the lamellar atomic structure, and lower wear resistance associated with lower

30

hardness of this phase.

EP

TE

D

21

Local phase change promoted by localized contact pressure in ductile materials are well reported in the literature

32

[11, 15, 13] and leads to gradients of local microstructure, mechanical and fracture properties. This behavior is not

33

commonly seen for carbon brittle materials, but Gogotsi et al. [14] showed a phase transformation of diamond after

34

an indentation test, in which the deformation of the original material produced nanocrystalline and graphitic carbon

35

localized structures. In terms of carbon-based coatings, Wang et al. [15] showed the transformation of hydrogenated

36

amorphous carbon (a-C:H) coating into graphitic nanocrystallites at the surface of the coating. These nanocrystals

37

were identified as small regions presenting graphite lamellar crystal arrangement, observed using the High Resolution

38

Transmission Electron Microscopy (HRTEM) technique. The formation of these nanostructures was associated to the

39

increase of sp2 carbon bond (G band) peak intensity of the Raman spectrum, observed inside of the deformed region.

40

As presented by Gogotsi et al. [14], Wang et al. [15] and others in the literature [9, 10, 16], the formation

AC C

31

2

ACCEPTED MANUSCRIPT

41

of graphitic structure arrangements from other carbon-based structures is widely known. On the other hand, phase

42

transformation from graphite or amorphous carbon into diamond structures is rare. Usually, stable diamond structures

43

are obtained by Chemical Vapor Deposition (CVD) or thermomechanical systems with high temperature and high

44

pressure (HTHP) techniques [17, 18, 19]. The equilibrium phase diagram of pure carbon systems [20, 30] and experimental observation [21, 22] indicates

46

the possibility of formation of diamond structures for pressures above 10 GPa at room temperature. This process

47

is possible given high hydrostatic pressures, which induces a stable carbon phase with diamond-like microstructure

48

(higher sp3 bonds) and a metastable phase of graphite-like microstructure (sp2 associated bonds).

RI PT

45

For dynamic cases, in which the phase transformation of carbon occurs under continuosly increasing pressures at

50

room temperature was reported by Bai et al. [22]. In that work, para-xylene was subjected to pressures as high as 30

51

GPa at room temperature. The phase transformation from liquid to carbon crystal structures was measured by XRD,

52

Raman and Infrared spectroscopy techniques. Results showed that the increase of the load compacting the para-xylene

53

produced a crystal structure with increasing sp3 (D band) Raman peak proportional to the applied load. Those authors

54

have also indicated that the pressure induced carbon phase transformation was an irreversible process, in which the

55

unloaded sample presented XRD and Raman spectrum peaks not initially presented for the para-xylene liquid.

M AN U

SC

49

Tribological applications usually present elevated contacting loads, resulting in contact pressures on the order of

57

tenths of GPa. Then, the use of a carbon-based coating presenting a dynamic mechanical behavior would promote

58

improved wear resistance while maintaining lower coeficient of friction. In this work, Raman spectroscopy and nano-

59

indentation measurements were conducted to monitor the structure and phase transformations of the materials inside

60

the wear track of high-contact-load tribological analyses. The system under study was composed of an a-C:H coating

61

deposited onto a ductile substrate subjected to high mechanical contact pressures, on the order of 2 GPa.

62

2. Coating deposition and characterization

EP

TE

D

56

To minimize the influence of substrate on the analyses of the coating, the substrate was made of AISI H13 steel,

64

which was water quenched from austenitizing temperature of 1000o C and tempered at 200o C. Prior to the deposition,

65

substrate surface was polished up to 1 µm (20 nm average roughness).

AC C

63

66

The coating and the interlayer in this work were deposited using a pulsed Direct Current Plasma Enhanced Chem-

67

ical Vapor Deposition (DC PECVD) reactor modified to include an additional cathode. This arrangement confines the

68

plasma near the surface of the samples, improving the deposition rate, the quality of the coating and allows a lower

69

deposition temperature (under 100o C). Details of this reactor and the deposition parameters were presented on the

70

literature [2, 23, 24].

71

A silicon interlayer was deposited to improve the adhesion of the carbon-based coating and the substrate. Silane

72

was used to grow this interlayer with 3 sccm gas flow, a bias of -950 V and a working pressure of 0.3 Pa. After this

73

process, the coating was grown using acetylene as precursor with an 8 sccm gas flow, at a pressure of 0.1 Pa and a 3

ACCEPTED MANUSCRIPT

74

constant bias of -800 V. According to Lugo et al. [2], these deposition parameters resulted in a silicon interlayer and a carbon-based

76

coating with, approximately, 500 nm and 1000 nm in thickness, respectively. The composition of the coating was

77

measured by Elastic Recoil Detection Analysis (ERDA) [2] as being 25 wt% of hydrogen and 75 wt% of carbon for

78

the first 500 nm and similar deposition results were obtained by Capote et al. [23, 24].

RI PT

75

81

1800 grating per mm difraction mesh, a 50 µm slit and a 100x objective lens, allowing measurement of D and G bands

82

of the Raman spectra in regions up to 1 µm in diameter with a spectral resolution of 2 cm−1 . All acquired Raman

83

spectra were processed in the Origin v9.5 by Origin Labs Corp., using a Gaussian Non-linear curve fit with a R2

84

higher than 0.95 to deconvolute the D and G shift positions.

SC

80

The characterization of the carbon bonds and the ID /IG intensity ratio at the surface was obtained by using a R confocal Horiba Xplora OneT M Raman system. This system was configured with a 532 nm laser wavelenght, a

79

Mechanical properties of the a-C:H coating were evaluated using the Hysitron Ti950 triboindenter. Several nano-

86

indentation curves were measured to estimate both hardness and elastic modulus, using the Oliver and Pharr method

87

[25]. Load to penetration depth curves were acquired using a Berkovich tip and a maximum load of 700 µN. This

88

load assured the measurement of coating properties with no significant influence of the substrate, once the maximum

89

penetration depth was kept under 50 nm. Additionally, critical loads for coating failure were measured using a cono-

90

spherical tip with 10 µm in diameter and maximum normal load of 360 mN. These critical loads were characterized

91

as the first pop-in in the load-penetration depth indentation curves and are directly linked to the failure of the coating

92

[26]. Also, this equipment can be used in a Scanning Probe Microscopy (SPM) configuration, producing a topography

93

image of selected regions. In this configuration, the Berkovich tip was used with a normal load of 2 µN to acquire

94

images of 10 µm x 10 µm.

95

3. Tribological analyses

EP

TE

D

M AN U

85

Components subjected to pure sliding conditions and elevated contacting loads may present high wear and friction.

97

One way to prevent such critical tribological conditions is the use of protective coatings. In this work, the tribological

98

behavior of a hydrogenated amorphous carbon coating was analyzed based on two sets of experiments: macroscale

99

lubricated reciprocating tests and microscale dry scratch tests.

AC C

96

100

R Macroscale lubricated reciprocating tests were conducted on an Optimol SRV v4 device with a ball on plane

101

configuration (Figure 1), considering that both the plane and the sphere were coated. Tests were based on a step-

102

increase in load from 5 N to 150 N with 5 minutes in each step (total duration of 30 minutes), 25 Hz of reciprocating

103

frequency and a stroke of 2 mm. A droplet of additive free Poly-Alpha-Olefin (PAO) synthetic lubricant in a mixture

104

of PAO4/6 was used in the contact region. The tested spheres were made of AISI 52100 material, coated in a similar

105

process used for the substrate.

106

Microscale scratch tests were carried on a Hysitron Ti950 triboindenter using a cono-spherical tip with 10 µm 4

SC

RI PT

ACCEPTED MANUSCRIPT

Figure 1: Macroscale reciprocating test configuration analyzed in this work, in which both sphere and disk were coated.

in diameter and loads ranging from 25 mN to 250 mN. The total scratch length was 500 µm and constant load

108

configuration were used. In microscale analyses, no liquid lubricant was used (dry condition).

109

4. Finite element analysis

M AN U

107

The microscale scratch test was also analyzed using the TriboCODE numerical package, which is in development

111

at the “Laborat´orio de Fenˆomenos de Superf´ıcie” (LFS) of the University of S˜ao Paulo, Brazil. This package is based

112

on several modules, which are used to simulate and analyze microscale tribological phenomena presenting complex

113

material microstructural configurations and physical mechanisms, such as abrasion, adhesion and thermal fatigue. In

D

110

116

The numerical model was solved using an explicit time evolving algorithm in quasi-static regime, in which no time

117

derivatives nor inertial forces were considered.

EP

TE

115

this work, the scratch test was analyzed as a tridimensional numerical model using three TriboCODE modules (TC R Abrasion, TC Bulk Material Properties and TC Coatings), coupled to the ABAQUS finite element environment.

114

The geometry of the numerical model is presented in Figure 2. In this figure, the configuration consists of the

119

cono-spherical tip (dark gray), the coating (red) and the substrate (light gray). The coating thickness was 2 µm and

120

the substrate was defined as a parallelepiped of 100 µm x 50 µm x 10 µm. The scratch length was 50 µm and the

121

normal load was increased continuously from 0 to 170 mN along the track. No lubricant was considered and two

122

failure models were employed: cohesive zone model for the coating/substrate interface and brittle failure model for

123

the coating. The substrate was considered elastic-perfectly plastic with mechanical properties representative of AISI

124

H13 steel. Mechanical properties of the coating were obtained from the experimental portion of this work (Section

125

2). The failure modeling and property measurements were analyzed following the procedure presented by Fukumasu

126

et al. [27].

AC C

118

5

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

Figure 2: Numerical model configuration consisting of a cono-spherical tip used to scratch the coated (red) substrate (light gray)

a)

D

b)

Figure 3: Coefficient of friction in the reciprocating tests: a) evolution during the ramping load and b) along the stroke for the last

TE

reciprocating movement of the sphere.

5. Results and discussions

128

5.1. Macroscale reciprocating test results

EP

127

Figure 3 presents the evolution of the coefficient of friction (COF) during the reciprocating tests. The COF curves

130

are presented based on the average of 3 repetition tests. Figure 3a indicates a stabilized value after 50 s and a minimum

131

COF of of 0.091±0.002 achieved at 350 s. This stabilized value of COF is compatible with peak values observed in

132

the reversion position of the reciprocating movement, presented in Figure 3b. In this figure, both reversal points of

133

the sphere movement present a surface to surface contact condition with no lubricant due to the zero velocity. Also, a

134

COF reduction to 0.080±0.001 at the middle of the stroke is observed. This reduction is related to a better lubrication

135

condition compared to the reversal points, given by the occurrence of the maximum relative velocity between the

136

contacting surfaces during the reciprocating cycle.

AC C

129

137

Figure 4 shows a SEM image of the macroscale wear track at the middle of the stroke, after the reciprocating test.

138

This image is representative of the tests and indicates that the coating was well adhered to the substrate, except for

139

isolated coating failures as indicated inside the marked region. 6

SC

RI PT

ACCEPTED MANUSCRIPT

M AN U

Figure 4: Representative wear track after the reciprocating test. The traced red rectangle indicates the region analyzed near the failure of the coating

Raman spectroscopy analyses inside the wear track indicated spatial variations on the G band, which may be

141

related to the fraction of sp2 bonds in the carbon coating. To analyze the change in the Raman shift of the G band, an

142

intensity ratio between two spectral ranges IGr /IGb was analysed. The IGr was defined in the spectral range from 1500

143

cm−1 to 1550 cm−1 , while IGb was defined from 1500 cm−1 to 1600 cm−1 . This selection allows the indication of the G

144

peak position inside the IGr spectrum range for higher values of IGr /IGb , while lower values indicate a G peak in 1550

145

cm−1 to 1600 cm−1 range.

TE

D

140

Figure 5 presents the IGr /IGb ratio image superimposed to the image of a failed region of the coating. In this figure,

147

a red-shift of the G band (higher IGr /IGb ) was observed at the coating surface inside and at the borders of the failed

148

region. The spatial variation of the non-processed G band peak position along the wear track suggests load-induced

149

surface changes at the coating during the reciprocating test. The deconvolution of the Raman spectrum are presented

150

in this figure. For positions presenting higher IGr /IGb ratio (red regions), the G peak position was 1534.4±2.6 cm−1 and

151

a FWHMG value was 197.9±7.8 cm−1 while, for regions presenting lower IGr /IGb (blue regions), the G peak position

152

was 1552.3±1.8 cm−1 and FWHMG was 175.4±8.7 cm−1 .

AC C

EP

146

153

Nano-indentation measurements (Figure 6a), with the corresponding calculation of the hardness and reduced

154

elastic modulus (Figure 6b), were applied to analyze the coating surface in regions presenting variations on the G

155

band position. In this figure, gray color represents measured regions presenting low values of reduced elastic modulus

156

and hardness, which may be a result of coating adhesive or cohesive failure. Nevertheless, Red and Black color

157

measurements presented similar reduced elastic modulus, in the range of 200 GPa to 230 GPa. Black squares in Figure

158

6b represent the original condition (outside the wear track) of the coating, while red measurements are representative

159

of the worn surface presenting a higher IGr /IGb ratio. Comparing both results, the worn surface presented higher 7

EP

TE

D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

Figure 5: Raman shift change of G peak position by mapping the intensity ratio IGr /IGb near the failed coating region: spatial variation of the non-processed G band peak position inside the wear track. In the graphs, gray squares indicate typical spectra obtained for similar regions, while lines indicate the deconvolution of the Raman spectra into D (blue) and G (red) bands.

8

ACCEPTED MANUSCRIPT

7 0 0

In s id e

5 0 0

O u ts id e 4 0 0 3 0 0

In s id e w ith in te r f a c e f a ilu r e

2 0 0 1 0 0 0 0

5

1 0

1 5

2 0

2 5

3 0

3 5

4 0

P e n e tr a tio n D e p th ( n m )

4 5

b)

SC

a)

RI PT

I n d e n t a t i o n L o a d ( µm )

6 0 0

Figure 6: Nano-indentation measurements of the coating: a) nano-indentation curves for inside (red and gray lines) and outside (black lines) of the wear track; b) results for hardness and reduced elastic modulus of the coating for inside (circles) and out-

M AN U

side (squares) of the wear track. Red circles indicate similar reduced elastic modulus but higher hardness compared to outside measurements (black squares), while gray circles indicate a reduction on both characteristics.

160

hardness (in the range of 25 GPa to 40 GPa), suggesting a change in the mechanical behavior that could be associated

161

to the Raman red-shift of the G band. For comparisson, the measured indentation hardness of the substrate was,

162

appoximately, 6.2±0.2 GPa.

163

As reported in the next section, dry microscale scratch tests reproduced the G band peak position variation, indicating no contribution from the additive free lubricant on the modification of the worn coating surface.

165

5.2. Microscale scratch test results

TE

D

164

The groove marks, produced inside the wear track of the macroscale reciprocating test, present similarities to

167

the deformation induced by a single abrasive particle scratching the coating surface. Therefore, microscale scratch

168

tests could be able to reproduce the groove formation, inducing localized changes of hardness and G band position

169

promoted by the abrasive particle in contact with the coating surface.

EP

166

In this test, the influence of the normal load on the coefficient of friction is displayed on Figure 7. Load values up

171

to 100 mN produce similar COF of, approximately, 0.05. Increasing the load led to an increase of the COF, in which

172

the intermediary load of 150 mN produced a COF of 0.08 while the highest analyzed load (250 mN) led to a COF of

173

0.15. In all cases, the tangential force was stable along the entire scratch distance.

AC C

170

174

Figure 8 shows a representative portion of the wear track obtained with the intermediary load case (150 mN

175

case). In this figure, the coating presents three failure modes: adhesive with small cohesive failures of the coating (a),

176

complete spallation of the coating (b) and adhesive failure of the coating without visible cohesive failure (c).

177

The measured map of the IGr /IGb ratio is presented in Figure 9, superimposed to an optical image of the microscale

178

scratch track, in a region presenting no spallation of the coating (Figure 8). In this figure, the higher intensity of

179

IGr /IGb (red colored regions) can be directly associated to regions in which the coating presented some type surface 9

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

AC C

EP

TE

D

Figure 7: Coefficient of friction along the scratch length during the test for analyzed normal loads

Figure 8: Microscale scratch track presenting the local typical observed failure modes: a) adhesive and cohesive failures of the coating; b) complete spallation of the coating and c) adhesive failure of the coating/substrate interface.

10

TE

D

M AN U

SC

RI PT

ACCEPTED MANUSCRIPT

EP

Figure 9: Superimposed Raman spectroscopy map of IGr /IGb ratio on the scratch track of figure 8. Higher ratio values (red colored regions) indicate a red-shift of the G band.

modification. This localized increase on the IGr /IGb ratio indicates a red-shift of the G band position induced by the

181

deformation of the coating, subjected to high localized contact stresses.

AC C

180

182

Figure 10 shows typical Raman spectra for point measurements of the coating surface inside and outside of the

183

scratched region. Also, this figure presents the deconvolution of the Raman spectrum inside and outside the scratch

184

track. A total of 10 point measurements in each region were analyzed, leading to a deconvoluted G band peak position

185

of 1544.7 ± 1.6 cm−1 inside and 1557.1 ± 0.7 cm−1 outside the scratched region. Also, the inside region presented a

186

full width at half maximum (FWHMG ) value of 183.9 ± 4.1 cm−1 while the outside region presented a value of 171.4

187

± 1.4 cm−1 .

188

Figure 11 was acquired using the Scanning Probe Microscopy technique, in which the indentation tip scans the 11

In s id e

In te n s ity ( a .u .)

-1

M AN U

1 3 5 5 c m

1 5 5 7 c m

1 3 7 3 c m

TE

1 0 0 0

1 2 0 0

EP

8 0 0

-1

-1

D

O u ts id e

-1

SC

1 5 4 4 c m

RI PT

ACCEPTED MANUSCRIPT

1 4 0 0

R a m a n s h ift ( c m

1 6 0 0 -1

1 8 0 0

2 0 0 0

)

AC C

Figure 10: Raman spectroscopy analysis of a-C:H coatings after the scratch test. Gray squares indicate typical spectra obtained for inside and outside the scratched regions, while lines indicate the deconvolution of the Raman spectra into D (blue) and G (red) bands.

12

ACCEPTED MANUSCRIPT

1 0 0

H e ig h t ( n m )

-1 0 -2 0 -3 0

-5 0 -6 0 0

a)

b)

2

RI PT

-4 0

4

6

8

1 0

L a te r a l D is ta n c e ( µ m )

Figure 11: Scanning Probe Microscopy of the scratch track presenting cohesive failure of the coating: a) planar view indicating the indentation analyses markers (black circles) and profile measurement across the track (green dotted line); b) measurement of

M AN U

SC

surface topography across the scratch track.

D

Figure 12: Indentation hardness measurements at positions indicated on Figure 11a and outside the scratch track (original condi-

TE

tion).

surface to obtain an image of the topography of the analyzed region. Figure 11a shows the top view of the wear

190

track. In this figure, three cohesive cracks at the coating surface are observed and recognized as the curved localized

191

variation on the topography. Figure 11b shows the depth profile of the scanned region, indicated in Figure 11a by the

192

green dotted line. The maximum residual depth of the scratched region was 60 nm or 6% of a-C:H coating thickness.

193

Also, hardness measurements (Figure 12) were taken at three points between two cohesive cracks of Figure 11a, as

194

indicated by points 1, 2 and 3.

AC C

EP

189

195

Figure 12 indicates higher hardness for the three measured points on the wear track compared to the initial values

196

of the coating. The difference on hardness measured for point 1 compared to points 2 and 3 may be related to the

197

proximity to the crack, in which the stifness of the coating may be reduced by the coating failure. Nevertheless,

198

these results suggest an increase of coating hardness coupled to an increase of IGr /IGb ratio, both promoted by the

199

deformation of the system induced by the scratch test, similar to the results obtained inside the grooves of macroscale

200

reciprocating tests.

201

Literature indicates that for amorphous carbon coatings subjected to elevated temperatures (annealing or rubbing 13

a)

RI PT

ACCEPTED MANUSCRIPT

b)

indentation curve and b) first pop-in presented on the loading curve.

SC

Figure 13: Coating cohesive failure critical loads measured by nano-indentation technique: a) typical load/penetration depth

processes), the sp3 carbon bonds tend to be converted to sp2 bonds and a blue-shift of the G band is observed [28, 29].

203

In this work, the deconvolution of the Raman spectrum indicated an opposite trend, in which a red-shift of the G band

204

and an increase of the FWHMG value inside the scratched region were observed. According to Ferrari and Robertson

205

[29], this red-shift is linked to an increase of sp3 content for an a-C:H coating and an increase in hardness. Also, the

206

increase of the FWHMG indicate a decrease of the in-plane correlation length (La ) and an increase of carbon disorder,

207

which is an indicative of the nucleation of small sp3 sites derived from sp2 bonds [29].

M AN U

202

One possible cause for this coating modification would be the high contact pressure developed during the scratch

209

test. To better evaluate the contact conditions, a finite element model was developed to simulate the coating/substrate

210

mechanical behavior under these test conditions.

211

5.3. Finite element analysis of the scratch test

212

5.3.1. Mechanical and fracture properties for the numerical simulations

EP

TE

D

208

The initial condition of Young’s modulus of the coating is obtained using the reduced elastic modulus (black

214

square symbols) plotted in Figure 6. The critical load for coating cohesive failure was obtained by nano-indentation

215

technique using a cono-spherical tip. A typical load/penetration depth curve for this test is presented on Figure 13a,

216

in which pop-ins are related to the cohesive failure of the coating, as previously described by Li [26] and considered

217

by Fukumasu [27]. According to those authors, the critical load for coating failure is assumed as the load in which

218

the first pop-in occurs. Figure 13b, which presents a series of pop-ins observed during the indentation of the coated

219

sample by the cono-spherical tip, indicates that the first pop-in occurs in a range from 139 mN to 158 mN normal load,

220

presenting an average value for this coating as 148±5 mN. This value is compatible with the scratch test transitioning

221

load, in which an increase in the COF and the coating failure were observed.

AC C

213

222

Data from Figure 13b were used to evaluate mode I failure stress and fracture energy for the analyzed coating

223

using the TC Coatings module of the TriboCODE. The critical failure stress was calculated as 2.5 GPa and the fracture 14

ACCEPTED MANUSCRIPT

226

used to model cohesive failure of the coating in this work.

227

5.3.2. Numerical scratch test results

RI PT

225

energy as 0.0002 J/µm. These values coupled to the failure strain of 0.0025, which is associated to the characteristic R mesh size, are the input parameters for the brittle failure model [31] implemented in the Abaqus environment and

224

Numerical simulations of the microscale scratch test provided the contact pressure and stress distributions of the

229

coating/substrate system. Figure 14 shows the contact pressure before cohesive or adhesive failure of the coating.

230

In this figure, the maximum contact pressure was calculated as 12 GPa distributed in a half moon geometry towards

231

the direction of the movement of the indenter (Figures 14a and 14b). The numerical simulation reproduces the

232

experimental pressure distribution given the similarities to the geometry of the spalled region of the coating, indicated

233

by the failure mode b presented in Figure 8.

SC

228

Figure 14c shows the evolution of the contact pressure with the increase of the normal load. In this figure, the

235

maximum contact pressure (12 GPa) is achieved for a load of 100 mN and is kept constant for higher loads, promoted

236

by the plastic deformation of the substrate. Once the coating damage threshold is achieved, cohesive cracks nucleate

237

at the surface of the coating and the contact pressure starts to oscilate, as noted in this figure. The numerical simulation

238

predicts the nucleation of cohesive cracks that promote the oscilations of the contact pressure at a scratch distance of

239

about 40 µm or 140 mN (Figure 14c), compatible with experimental critical normal load for the failure of the coating.

240

Figure 15 presents the distribution of the minimum principal stresses in the coating/substrate system. The high

241

contact pressure developed at the coating surface promoted an elevated internal stress distribution in the coating,

242

achieving compressive values of 20 GPa near the surface and decreasing to 12 GPa at coating/substrate interface.

D

M AN U

234

As reported by Bai et al. [22], this high level of compressive stresses developed in the coating is enough to induce

244

a phase transformation of the carbon material at room temperature, in which compressive stresses on the order of 20

245

GPa or higher promotes the rearrangement of carbon bonds, becoming sp3 dominant (red-shift of the G band, higher

246

FWHMG values and higher hardness).

EP

TE

243

After the removal of the load, the induced carbon transformation inside the wear grooves and the scratch track

248

remains stable. This induced phase was observed by the red-shift of the G band (Figures 5 and 10) and by higher

249

hardness measurements (Figures 6 and 12) for both macro and microscale analyzed tracks. This phase stability was

250

also reported by Bai et al. after the removal of the load, in which a solid phase of the carbon material was observed,

251

presenting crystalline material identified by the X-Ray diffraction technique.

AC C

247

15

RI PT

ACCEPTED MANUSCRIPT

M AN U

SC

a)

b)

c)

D

Figure 14: Contact Pressure at the coating promoted by the indenter movement: a) Instantaneous spatial distribution of the contact

AC C

EP

TE

pressure; b) detail of the contact region and c) evolution of the contact pressure with the ramping load during the scratch test.

Figure 15: Minimum Principal Stresses distribution developed inside of the material during the scratch test.

16

ACCEPTED MANUSCRIPT

252

6. Conclusions Coatings that present adaptive features, such as microstructure and mechanical properties, have the ability to

254

improve the performance of components by tuning specific characteristics during the process. In this work, local

255

transformation of an amorphous hydrogenated carbon coating (a-C:H) was analyzed after high load lubricated re-

256

ciprocating ball-on-plane tests and dry micro-scratch tests using Raman spectroscopy and indentation hardness, in

257

which:

260

261

262

263

264

265

regions under high contacting loads on both test configurations;

SC

259

• the analyzed coating presented low coeficient of friction, good adherence to the substrate and small spalled

• Raman spectroscopy analyses of the macroscale wear track indicated regions presenting a red-shift of the G band compared to the coating original condition (outside the track);

• indentation hardness analyses indicated a spatial correlation of higher hardness on regions with higher IGr /IGb

M AN U

258

RI PT

253

intensity ratio;

• microscale dry scratch tests presented similar Raman red-shift compared to macroscale reciprocating tests, indicating a negligible significant contribution from the additive free lubricant on surface modification; • good agreement was obtained between experimental and numerical analysis of coating critical load for first

267

failure. This result was enhanced by the good measurement of both mechanical and fracture properties of the

268

coating, used as input to the numerical solver;

271

272

TE

270

• the experimental spalled region shape, observed on the transitional load of the scratch test, was compatible with the spatial distribution of contact stresses computed by the numerical simulation; • numerical simulation indicates high contact pressure (>12 GPa) developed at the surface and high internal

EP

269

D

266

stresses, raging from 20 GPa to 12 GPa, are developed along coating thickness; • high internal stresses, red-shift of the G band, higher values of FWHMG and the increase on indentation hardness

274

inside the scratched region are compatible with the nucleation of sp3 carbon bond sites derived from sp2 bonds,

275

according to the literature.

276

AC C

273

Acknowledgements

277

The authors recognize the National Council for Scientific and Technological Development (CNPq), the Sao Paulo

278

Research Foundation (FAPESP) and the National Bank for Economic and Social Development (BNDES) for financial

279

support.

17

ACCEPTED MANUSCRIPT

280

References [1] K. Holmberg, P. Andesson, A. Erdemir, Global Energy consumption due to friction in passenger cars. Tribol. Int., 47, 221-234 (2012).

282

[2] D.C. Lugo, P.C. Silva, M.A. Ramirez, E.J.D.M. Pillaca, C.L. Rodrigues, N.K. Fukumasu, E.J. Corat, M.H. Tabacniks, V.J. Trava-Airoldi,

283

Characterization and tribologic study in high vacuum of hydrogenated DLC films deposited using pulsed DC PECVD system for space

284

applications, Surf. Coat. Technol., 332, 135-141 (2017).

287 288 289 290 291 292 293 294

[3] J. Solis, H. Zhao, C. Wang, J.A. Verduzco, A.S. Bueno, A. Neville, Tribological performance of an H-DLC coating prepared by PECVD, Appl. Surf. Sci., 383, 222-232 (2016).

[4] D. Lusk, M. Gore, W. Boardman, T. Casserly, K. Boinapally, M. Oppus, D. Upadhyaya, A. Tudhope, M.Gupta, Y. Cao, S.Lapp, Thick DLC films deposited by PECVD on the internal surface of cylindrical substrates, Diam. Relat. Mat., 17, 1613-1621 (2008).

[5] A. Erdemir, C. Donnet, Tribology of diamond-like carbon films: recent progress and future prospects, J. Phys. D-Appl. Phys., 39, 311-327 (2006).

SC

286

[6] D. Sheeja, B.K. Tay, S.M. Krishnan, L.N. Nung, Tribological characterization of diamond-like carbon (DLC) coatings sliding against DLC coatings, Diam. Relat. Mat., 12, 1389-1395 (2003).

[7] H.A. Tasdemir, T. Tokoroyama, H. Kousaka, N. Umehara, Y. Mabuchi, Friction and wear performance of boundary-lubricated DLC/DLC contacts in synthetic base oil, Procedia Engineering, 68, 518-524 (2013).

M AN U

285

RI PT

281

295

[8] S.J. Bull, Tribology of carbon coatings: DLC, diamond and beyond, Diam. Relat. Mat., 4, 827-836 (1995).

296

[9] Y. Liu, A. Erdemir, E.I. Meletis, An investigation of the relationship between graphitization and frictional behavior of DLC coatings, Surf.

299 300 301 302 303 304 305

[10] S.R. Polaki, N. Kumar, K. Ganesan, K. Madapu, A. Bahuguna, M. Kamruddin, S. Dash, A.K. Tyagi, Tribological behavior of hydrogenated DLC film: Chemical and physical transformations at nano-scale, Wear 339-339, 105-113 (2015). [11] F.P. Marques, C. Scandian, A.C. Bozzi, N.K. Fukumasu, A.P. Tschiptschin, Formation of a nanocrystalline recrystallized layer during microabrasive wear of a cobalt-chromium based alloy (Co-30Cr-19Fe), Tribol. Int., 116, 105-112 (2017). [12] S.J. Wang, H. Wang, K. Du, W. Zhang, M.L. Sui, S.X. Mao, Deformation-induced structural transition in body-centred cubic molybdenum,

D

298

Coat. Technol., 86-87, 564-568 (1996).

Nat. Commun., 5, 3433 (2014).

[13] A. Das, Contribution of deformation-induced martensite to fracture appearance of austenitic stainless steel, Mater. Sci. Technol., 32, 1366-

TE

297

1373 (2016).

306

[14] Y.G. Gogotsi, A. Kailer, K.G. Nickel, Pressure-induced phase transformations in diamond, J. Appl. Phys., 84, 1299-1304 (1998).

307

[15] D.S. Wang, S.Y. Chang, Y.C. Huang, J.B. Wu, H.J. Lai, M.S. Leu, Nanoscopic observations of stress-induced formation of graphitic nanocrystallites at amorphous carbon surfaces, Carbon, 74, 302-311 (2014).

EP

308

[16] Y.G. Gogotsi, A. Kailer, K.G. Nickel, Transformation of diamond to graphite, Nature 401, 663-664 (1999).

310

[17] B.V. Spitsyn, L.L. Bouilov, B.V. Derjaguin, Vapor growth of diamond on diamond and other surfaces, J. Cryst. Growth, 52, 219-226 (1981).

311

[18] S. Matsumoto, Y. Sato, M. Tsutsumi, N. Setaka, Growth of diamond particles from methane-hydrogen gas, J. Mater. Sci., 17, 3106-3112

312

AC C

309

(1982).

313

[19] H. Kanda, P. Capper, Bulk Crystal Growth in Electronic, Optical and Optoelectronic Materials, John Wiley and Sons, (2005).

314

[20] Q. Jiang, J.C. Li, G. Wilde, The size dependence of the diamond-graphite transition, J. Phys.-Condes. Matter, 12, 5623-5627 (2000).

315

[21] T.B. Shiell, D.G. McCulloch, J.E. Bradby, B. Haberl, R. Boehler, D.R. McKenzie, Nanocrystalline hexagonal diamond formed from glassy

316 317 318 319 320 321 322

carbon, Sci. Rep., 6, 37232 (2016).

[22] Y. Bai, Z. Yu, R. Liu, N. Li, S. Yan, K. Yang, B. Liu, D. Wei, Lin Wang, Pressure-Induced Crystallization and Phase Transformation of Para-xylene, Sci. Rep., 7, 5321 (2017). [23] G. Capote, G.C. Mastrapa, V.J. Trava-Airoldi. Influence of acetylene precursor diluted with argon on the microstructure and the mechanical and tribological properties of a-C:H films deposited via the modified pulsed-DC PECVD method, Thin Solid Films, 589, 286-291 (2015). [24] G. Capote, M.A. Ramirez, P.C. Silva, D.C. Gonc¸alez, V.J. Trava-Airoldi. Improvement of the properties and the adherence of DLC coatings deposited using a modified pulsed-DC PECVD technique and an additional cathode. Surf. Coat. Technol., 308, 70-79 (2016).

18

ACCEPTED MANUSCRIPT

323 324

[25] W.C. Oliver, G.M. Pharr, An improved technique for determining hardness and elastic modulus using load and displacement sensing indentation experiments, J. Matter. Res., 7, 1564-1583 (1992).

325

[26] X. Li, B. Bhushan, Measurement of fracture toughness of ultra-thin amorphous carbon films, Thin Solid Films 315, 214-221 (1998).

326

[27] N.K. Fukumasu, R.M. Souza, Numerical evaluation of cohesive and adhesive failure modes during the indentation of coated systems with

327

compliant substrates, Surf. Coat. Technol. 260, 266-271 (2014). [28] J. Robertson, Mechanical properties and structure of diamond-like carbon, Diam. Relat. Mat., 1, 397-406 (1992).

329

[29] A.C. Ferrari, J. Robertson, Raman spectroscopy of amorphous, nanostructured, diamond-like carbon, and nanodiamond, Phil. Trans. R. Soc.

SC

Finite Elements, Cem. Concr. Res., 6, 773-782 (1976).

M AN U

334

(1993).

[31] A. Hillerborg, M. Modeer, P.E. Petersson, Analysis of Crack Formation and Crack Growth in Concrete by Means of Fracture Mechanics and

D

333

TE

332

EP

331

¨ ı2512 (2004). Lond. A, 362, 2477 A` ’ [30] H.O. Pierson, Handbook of carbon, graphite, diamond and fullerenes: properties, processing and applications, Noyes Publication, New Jersey

AC C

330

RI PT

328

19

ACCEPTED MANUSCRIPT

• High contact pressure tribology measurements of Diamond like Carbon coatings

336

• High correlation of higher indentation hardness on regions with increased sp3 carbon bonds inside wear track

337

• Damage calibrated model to reproduce realistic cohesive and adhesive coating failure by FEM model

338

• Contact pressures and internal coating stresses were induced in the range of equilibrium formation of diamond-

EP

TE

D

M AN U

SC

like carbon phase

AC C

339

RI PT

335

20

ACCEPTED MANUSCRIPT

• High contact pressure tribology measurements of Diamond like Carbon coatings

316

• High correlation of higher indentation hardness on regions with increased sp3 carbon bonds inside wear track

317

• Damage calibrated model to reproduce realistic cohesive and adhesive coating failure by FEM model

318

• Contact pressures and internal coating stresses were induced in the range of equilibrium formation of diamond-

EP

TE D

M AN U

SC

like carbon phase

AC C

319

RI PT

315

18