Vous êtes sur la page 1sur 8

Wear 302 (2013) 955962

Contents lists available at SciVerse ScienceDirect

Wear
journal homepage: www.elsevier.com/locate/wear

Wear resistant electrically conductive AuZnO nanocomposite


coatings synthesized by e-beam evaporation
N. Argibay n, S.V. Prasad, R.S. Goeke, M.T. Dugger, J.R. Michael
Materials Science and Engineering Center, Sandia National Laboratories, Albuquerque, NM 87123, USA

a r t i c l e i n f o

abstract

Article history:
Received 5 September 2012
Received in revised form
11 January 2013
Accepted 17 January 2013
Available online 1 February 2013

The tribological and electrical behavior of e-beam codeposited AuZnO nanocomposite lms were
investigated for compositions in the range 0.128.0 vol% of ZnO. A 2.0 vol% ZnO lm did not exhibit
measurable wear sliding against a commercial gold alloy rider. Electron backscatter diffraction (EBSD)
analysis on lm surfaces of varying composition revealed a signicant reduction in grain size for a
0.1 vol% ZnO lm contrasting with a pure Au lm. Electrical resistivity measurements of lms in the
range 0.028.0 vol% ZnO exhibited a linear increase in resistivity from 2.73 mO cm to 39.88 mO cm. The
friction, wear, and electrical contact resistance of 2.0 and 28.0 vol% ZnO lms sliding against
commercially available hardened Au riders (72Au14Cu8Pt5Ag by weight) were investigated.
Friction coefcient and electrical contact resistance data were in the range m 0.30.5 and ECR 40 mO
and 5002000 mO, respectively, for 2 mm thick lms deposited on conductive substrates in unidirectional sliding at 1 mm/s under a 185 MPa maximum Hertzian contact pressure (100 mN contact force).
The presence of ZnO transfer to the rider was observed via energy-dispersive X-ray spectroscopy (EDS).
Signicant improvements in wear resistance and friction behavior were observed for AuZnO
composite lms with volume fractions of ZnO commensurate with the metal species codeposited in
traditional hardened gold coatings.
& 2013 Elsevier B.V. All rights reserved.

Keywords:
AuZnO
Wear
Dispersion strengthened composites
Nanocomposites
Electrical resistivity

1. Introduction
The use of Au in the form of high Au content alloys in electroplated thin lms, wire bonding, and as a solder or solderable
coating is commonplace in the electronics industry [1]. A 2001
report estimated that approximately 280 metric tons of gold were
used in electronic related applications that year [2], the largest
portion attributed to electroplated coatings on connectors and
contacts [1]. In a number of applications across several size scales,
such as micro-electromechanical systems (MEMS) switches or
high-current density slip-rings in modern wind turbines, the
performance of the system often depends on the ability to
transfer electricity robustly across a separable or sliding mechanical contact, prompting the use of thin hard gold lms as a low
wear, low friction, chemically inert, and conductive alternative to
bare, non-noble, and pure metals. Hard gold, which may be
deposited in thin lms through a number of processes, the oldest
and most commercially prevalent being electroplating [3,4], refers
to alloyed Au composed of typically greater than 98 wt%
Au exhibiting the chemical inertness and high electrical conductivity of pure Au. The addition of 12 wt% of codeposited species

Corresponding author. Tel.: 1 505 284 4397.


E-mail addresses: nargiba@sandia.gov, nargibay@gmail.com (N. Argibay).

0043-1648/$ - see front matter & 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.wear.2013.01.049

such as Co or Ni signicantly improves the tribological performance [5], in contrast to gross adhesive wear (galling) [6] and
fretting wear [7] exhibited by pure Au in otherwise unlubricated
contact.
One of the challenges affecting the performance of hard gold
coated non-noble metal connectors is the long-term degradation
in electrical contact resistance attributed to the gradual low
temperature (below of about 150 1C) solid diffusion of codeposited and underlayer metal species to the surface [810] and
subsequent formation of low conductivity oxide lms [5]. The
magnitude of this phenomenon is exacerbated by the use of
increasingly thinner lms in an effort to reduce waste and cost
while optimizing performance, particularly for the case of low
duty cycle separable connectors where wear may be low. While
the use of diffusion barrier lms comprised of low miscibility, low
diffusivity metals (e.g. nickel) provide a means of mitigating
diffusion of underlayer metals exhibiting high diffusivity (e.g.
copper) [11,12]. Recent efforts to nd alternative materials solutions to mitigate this type of degradation has led to the use of
codeposited non-metal species in place of the traditional metals
[13,14]. As demonstrated by Lo et al. [15], the principal role of the
codeposited species appears to be grain boundary pinning during
deposition preventing recrystallization, leading to a ner grained
microstructure and correspondingly increased hardness attributed to the HallPetch strengthening.

956

N. Argibay et al. / Wear 302 (2013) 955962

This paper describes the synthesis of zinc oxide (ZnO) hardened gold lms via electron beam deposition, and presents
experimental measurements of friction, wear, and electrical contact resistance of the materials in unlubricated contact against a
commercially available hardened gold alloy satisfying ASTM
standard specication B541 for gold based electrical contact
alloys. Measurements of electrical resistivity and surface microstructure are presented for a variety of lm compositions.

2. Methods
2.1. Materials synthesis and characterization
Composite thin lms of gold and zinc oxide were synthesized
by co-deposition using a 10 kV Triad e-beam evaporation system.
The source materials were Au pellets with 99.999% purity and
ZnO tablets with 99.9% purity, both from Materion Advanced
Chemicals. Each material deposition rate is controlled independently using feedback from quartz crystal microbalances (QCM).
The calibration of the lm thickness and deposition rate was
established by stylus prolometry. A shutter in front of the
substrates enables the deposition rates to be established before
deposition, providing a uniform lm composition from start to
end. The source to substrate distance was 305 mm and the
depositions were carried out at room temperature. This high
vacuum system was pumped to a base pressure of less than
1.3  10  4 Pa before beginning the deposition.
Using this technique we deposited 2 mm thick AuZnO lms at
a number of selected compositions onto bare 25 mm  25 mm
sections of silicon wafer for the purpose of determining electrical
resistivity and composition. Silicon wafers previously coated with
titanium and platinum at a thickness of 2 mm each were coated
for the purpose of hardness testing. Metallic substrates of
polished Alloy 52 (nominally a 50/50 wt% alloy of FeNi) were
coated with 2 mm thick lms for the purpose of sliding electrical
contact testing; the substrates were polished to an average
surface roughness of less than 10 nm prior to coating. The
deposition rates were varied between 0.00 and 2.00 nm/s to
synthesize the desired compositions. A number of substrate
materials were coated in each deposition run to take advantage
of multiple measurement techniques.
Electron backscatter diffraction (EBSD) was carried out on the
Au-lms. The samples required no additional preparation due to
the low roughness of the surface and their inherent conductivity.
EBSD measurements were conducted in a Zeiss Supra 55VP
scanning electron microscope (SEM) using a 20 kV accelerating
voltage and a beam current of 35 nA. Orientation maps were
obtained using a square grid of points that was 400  300 pixels

with a step size of 25 nm. The resulting EBSD patterns were


analyzed using the Oxford Instrument Channel 5 software. Some
minor data correction or cleaning was used to remove individual
mis-indexed pixels and to ll in missing pixels. Electron micrographs of the worn surfaces and transfer lms were acquired
using a JEOL JSM 5900LV SEM using a 15 kV accelerating voltage.
A Thermo Scientic NORAN System SIX microanalysis system was
used to perform energy dispersive X-ray spectroscopy (EDS)
analyses of the worn surfaces and surface debris. The electrical
resistivity of the lm material was characterized using a four
point probe technique, where the lm or sheet resistance is
equivalent to the material resistivity divided by the lm thickness. For a known lm thickness, measured by scanning white
light interferometer (SWLI) prolometry at the edge of a masked
corner for each coated substrate, the material resistivity was
calculated based on two resistance measurements where the
location of the probes was varied accordingly [16]. A preliminary
measurement of hardness performed using a nano-indenter and a
1 mm radius conical diamond tip on a Au1 vol% ZnO lm gave a
value of approximately 2.6 GPa at 0.1% strain [17].
2.2. Experimental methods: sliding electrical contact testing
A custom built multifunction electrical contact tribometer was
used to measure friction coefcients and electrical contact resistance (ECR) in unidirectional linear reciprocating motion between
e-beam plated Alloy 52 coupons and metallic riders. Fig. 1
presents a photograph and accompanying diagram of the sliding
electrical contact, showing the electrical connections to the rider
and coated substrate. The riders were Neyoro G cylinders (nominally 72Au14Cu8Pt5Ag by weight, purchased from Deringer
Ney) with a spherically tipped end machined by turning to a
radius of 1.59 70.12 mm and lapped to achieve an average surface roughness of 10 nm or better. A nominal contact (normal)
force of 100 mN and sliding speed of 1 mm/s were used for all
experiments, commensurate with the conditions found in a
variety of electrical devices. Friction and ECR were measured for
one pure Au and two nanocomposite lms (Au alloyed with
2.0 and 28.0 vol% ZnO). For the pure Au and Au28.0 vol% ZnO
coated specimens 100 cycles were performed with a track length
of 5 mm. For the 2.0 vol% ZnO coated specimen a total of 1000
cycles were run, and the track length was systematically reduced
to generate regions of increasingly greater wear; the initial track
length was 5 mm for the rst 100 cycles, followed by 400
additional cycles to a length of 4 mm, and an additional 500
cycles to a length of 3 mm. The result was to achieve one
uninterrupted test where the wear track segment between
4 and 5 mm exhibited a total of 100 cycles, the segment between
3 and 4 mm a total of 500 cycles, and the segment between 0 and

20

10 mm

spherically
tipped rider

canted
rider
sense/source
high channels
source/meter unit
coated substrate

sense/source
low channels

polycarbonate insulator
unidirectional sliding

Fig. 1. (A) Photograph of the rider and substrate during a sliding contact experiment and (B) a diagram illustrating the electrical circuit.

N. Argibay et al. / Wear 302 (2013) 955962

3. Results
3.1. Electrical resistivity and surface microstructure of AuZnO
nanocomposites
The measured electrical resistivity values for these lms as a
function of composition are shown in Table 1. The volume
percentages shown are the targeted values based on the deposition rates of the individual materials as the resulting composition
has not yet been veried. The data shown in Table 1 and plotted
in Fig. 2 (in semi-log form) shows that electrical resistivity
increases linearly with ZnO concentration up to 28.0 vol%
ZnO. A least squares linear regression (shown in Fig. 2) resulted
in a high goodness-of-t, with a coefcient of determination
R2 0.992.
Fig. 3 shows results of EBSD characterization of 2 mm thick ebeam deposited pure Au and Au0.1 vol% ZnO nanocomposite
lms. Fig. 3(A) shows the orientation map with respect to the
surface normal or the deposition direction. The map is primarily
shaded in blue, indicating a strong preference of the lm to grow
parallel to the /111S direction. The color key for these orientation maps is shown as an inset. There appears to be a bi-modal
grain size distribution with many larger grains surrounded by
smaller grain size regions. Fig. 3(B) shows an in-plane orientation
map for the Au0.1 vol% ZnO lm showing a signicantly reduced
average surface grain size in contrast to the pure Au lm. Fig. 3(C)
shows the contoured /111S pole gure for the pure-Au sample,
where the /111S ber texture of the lm is clearly visible. This
implies that the lm grows with a distinct and strong /111S outof-plane texture with no in-plane texture.
Table 1
Experimental measurements of electrical resistivity as a function of composition
using the four point method with corner probes.
Au (vol%)

ZnO (vol%)

Electrical resistivity, q (lX cm)

100.0
99.9
99.5
99.0
98.0
95.0
87.5
72.0

0
0.1
0.5
1.0
2.0
5.0
12.5
28.0

2.73
2.84
3.45
4.65
5.57
10.71
16.52
39.88

3.2. E-beam pure Au


Fig. 4 shows friction coefcient and electrical contact resistance (ECR) data for a 2 mm thick e-beam deposited pure Au lm.
Here, an additional plot provided along with the friction coefcient and ECR data plots provides a measure of the quality of the
contact, based on the arbitrarily chosen criterion that any ECR
data points above 1 O were considered as a contact break and
were omitted from the calculation of average ECR for each cycle
(from 250 data points acquired per cycle). This criterion was
justied based on the observation that majority of these points
resulted in ECR measurements on the order of megaohms, rather
than a few ohms. So a fraction of ECR data above 1 O of 0.5 implies
that half of the 250 ECR data points acquired for that cycle
occurred during a contact separation event, possibly attributed
to phenomena such as stick-slip or contact with relatively
electrically insulating third-body debris particles.
For the pure Au lm the initial friction behavior was m 41,
suggestive of adhesive contact between the Au lm and primarily
Au alloy rider. An apparent transition period was observed up to
cycle 28, after which the average and standard deviation in ECR
increase steadily. The quality of electrical contact suggested by
the graph of ECR data above 1 O suggests a period of anomalously
high instability between cycles 28 and 43, where nearly 80% of
the data was indicative of contact separation. The lm was visibly
(to the naked eye) worn off during this transition period.
A scanning white light interferometer was used to measure the
step height near the edge of the wear track upon completion of
the experiment, revealing complete removal of the lm over a
width of approximately 2 mm along the entire 5 mm length of
wear track. Cracking and delamination was observed near the
wear track edges in an optical microscope. Relatively low friction
coefcients were measured beyond cycle 55, likely a result of
sliding contact between the rider and Alloy 52 substrate. The ECR
data also suggests a transition from sliding against the Au lm to
sliding against the Alloy 52 substrate. Fig. 5 shows an SEM
micrograph of debris accumulated on the surface of the rider.

3.3. AuZnO nanocomposites


Fig. 6 shows friction coefcient and ECR data for two nanocomposite lms, consisting of 2 mm thick e-beam deposited

40

electrical resistivity (-cm)

3 mm a total of 1000 cycles, enabling a multipoint determination


of specic wear rate of the lms. Data was acquired and stored at
a rate of 50 Hz for normal force, friction force, wear track position,
electrical current ow through the circuit, and voltage drop across
the contact. An Agilent Technologies B2911A precision source/
meter unit was used to provide electrical current through the
contact in a voltage-regulated remote-sensing mode. The voltage
was varied for each lm to achieve approximate direct currents
(DC) of 100 mA, experimentally determined by running a voltage
sweep upon rst contact and prior to initiating a sliding experiment. Voltage-regulation was preferred to current-regulation so
as to prevent arcing due to periodic contact separation during
sliding, where the internal voltage regulating algorithm of the
SMU may generate sufciently elevated voltage spikes while
chasing rapidly changing ECR. The rider was canted in its holder
so that multiple experiments could be run with a single pin by
rotating it to an unworn position on the spherical tip. Sliding
experiments were performed exposed to laboratory air maintained at 71 1C, and with a relative humidity of 10%.

957

Au ZnO

= 1.29 . (vol.%ZnO) + 2.869 -cm

R 0.992

30

20

10

0
0.1

10

100

vol. % ZnO
Fig. 2. Electrical resistivity of electron beam codeposited AuZnO as a function of
composition; resistivity measurements were made using a four point probe
method (i.e. the van der Pauw method [16]) for 2 mm thick lms of varying
composition deposited directly onto bare square cross-section silicon wafer
sections with a 400 nm thick SiO2 layer insulating the substrate from the lms.
A least squares linear regression with a coefcient of determination R2 0.992
is shown.

958

N. Argibay et al. / Wear 302 (2013) 955962

{001}

{110}

111

2 m

2 m

{111}
101

001

{112}

2 m

2 m

Fig. 3. Electron backscatter diffraction (EBSD) maps of 2 mm thick e-beam deposited lms of (A and B) pure Au and (C and D) Au0.1% ZnO nanocomposite in the surface
normal (A and C) and in-plane orientations (B and D), and pole gures (E) corresponding to the pure Au lm. Both lms exhibited a preferential /111S out-of-plane
texture (prevalence of blue in maps A and C). The surface grain size was signicantly smaller for the nanocomposite lm; black pixels represent regions of grain size below
the resolution of the instrument, approximately 100 nm. (For interpretation of the references to color in this gure legend, the reader is referred to the web version of this
article.)

1.5

high adhesion

friction coefficient

1.3

onset of
high wear

1.1

chatter
due to
debris
pile-up

0.9

sli

solid line: avg


shaded area: avg

+/-

transition to sliding on substrate

din

dir
ec

tio

0.7
0.5
0.3

fraction of ECR
data above 1

contact
resistance ()

0.1
0.5
0.4
0.3
0.2
0.1
0.0

100 m

1.0
0.5
0.0
0

10

20

30

40

50
cycle

60

70

80

90

100

Fig. 4. Friction coefcient and ECR data per sliding cycle for a 2 mm thick lm of
e-beam deposited pure Au on an Alloy 52 conductive substrate sliding against a
spherically tipped Neyoro-G rider (72Au14Cu8Pt5Ag by weight); based on the
250 data points acquired over each sliding cycle, the shaded areas correspond to
the average values plus/minus one standard deviation. The graph of fraction per
cycle of ECR data above 1 O indicates the fraction of 250 data points per cycle that
were above an arbitrary threshold of 1 O, attributed to contact separation due to
debris. These points were omitted from the plotted ECR average and standard
deviation calculations.

2.0 vol% and 28.0 vol% ZnO lms. In contrast with the pure Au lm
data, the two lms exhibit average friction coefcients m o0.5
from the rst cycle, though the standard deviation in friction
coefcient data per cycle is initially signicantly higher for the
2.0% ZnO lm. The 28.0% ZnO lm exhibited a 4 order of
magnitude greater ECR in contrast to the 2.0% ZnO lm, which
remained consistently low and on the same order as that
exhibited by the pure Au lm prior to gross delamination for
the total 100 cycles. However, unlike the initial behavior of the

Fig. 5. SEM micrograph of wear debris on the rider surface using a shadow
backscatter electron (BSE) detector. The observed dark contrast areas on the edges
of debris particles are due to shadowing of the off-axis BSE detector; the curvature
of the pin causes the dark contrast near the lower right of the image.

pure Au lm, the 2.0 vol% ZnO lm exhibited contact separation


(based on the criterion of ECR greater than 1 O) in the range of
1050% of the 250 data points acquired per cycle, over the length
of the 100 cycle test. Fig. 7 shows the continuation of the sliding
experiment for the 2.0 vol% lm up to a total of 1000 cycles, with
periodically varying track lengths as discussed in the previous
section.
In contrast with the pure Au data, the relatively low average
and standard deviation in ECR data for the 2.0 vol% ZnO lm
coupled with the signicant number of points above 1 O may be
explained by the presence of non-conductive ZnO debris present
around the rider wear scar. Fig. 8 shows SEM/EDS characterization of the rider wear scar as well as surface topographical scans
acquired with a scanning white light interferometer (SWLI). Fig. 9
shows SWLI topographical maps and SEM micrographs of the
2.0 vol% lm wear track in a region of the wear track which
observed 1000 cycles of sliding. It is apparent from these images

N. Argibay et al. / Wear 302 (2013) 955962

1.5
friction coefficient

solid line: avg


shaded area: avg +/-

2.0 vol. % ZnO-Au


28.0 vol. % ZnO-Au

1.3
1.1

959

(Fig. 9) indicate primarily material transfer from the rider to the


lm, in agreement with the signicant wear removal observed on
the rider (Fig. 8), after 1000 cycles of sliding (with progressively
shorter cycle lengths, as described before).

0.9
0.7

4. Discussion

0.5
0.3

4.1. Materials characterization and electrical resistivity

0.1

fraction of ECR
data above 1

contact resistance ()

2,000
1,000
500
0.5
0.4
0.3
0.2
0.1
0.0
1.0

* for the 28.0 % ZnO-Au nanocomposite the1threshold was ignored and all the data
acquired for each cycle was used to calculate the average and standard deviation

0.5
0.0
0

10

20

30

40

50

60

70

80

90

100

cycle
Fig. 6. Friction coefcient and ECR data per sliding cycle for 2 mm thick e-beam
deposited nanocomposite lms 2.0 vol% and 28.0 vol% ZnO on conductive Alloy
52 substrates sliding against a spherically tipped Neyoro-G rider (72Au14Cu
8Pt5Ag by weight); based on the 250 data points acquired over each sliding cycle,
the shaded areas correspond to the average values plus/minus one standard
deviation. The graph of fraction per cycle of ECR data above 1 O indicates the
fraction of 250 data points per cycle that were above an arbitrary threshold of 1 O,
attributed to contact separation due to debris. These points were omitted from the
plotted ECR average and standard deviation calculations.

friction coefficient

1.0
solid line: avg

continuation of Figure 6
0.8

shaded area: avg +/-

0.6
0.4
0.2

fraction of ECR
data above 1

ECR ()

0.0
0.3
0.2
0.1
0.0

data loss

data loss

1.0
0.5
0.0
0

100

200

300

400

500

600

700

800

900

1000

cycle

Fig. 7. Friction coefcient and ECR data per sliding cycle for 2 mm thick e-beam
deposited Au2.0 vol% ZnO on a conductive Alloy 52 substrate sliding against a
spherically tipped Neyoro-G rider (72Au14Cu8Pt5Ag by weight); based on
the 250 data points acquired over each sliding cycle, the shaded areas correspond
to the average value plus/minus one standard deviation. The graph of fraction per
cycle of ECR data above 1 O indicates the fraction of 250 data points per cycle that
were above an arbitrary threshold of 1 O, attributed to contact separation due to
debris. These points were omitted from the plotted ECR average and standard
deviation calculations.

Preliminary measurements of friction and electrical contact resistance suggest that ZnO may be a suitable material for codeposition
with Au to improve the tribological behavior of pure Au while
retaining high electrical conductivity. The reduction in grain size
observed with the addition of 0.1 vol% ZnO suggests the signicant
HallPetch strengthening, commensurate with the well-documented
strengthening mechanism of traditional hard Au [15]. Grain renement and associated Zener drag strengthening [18] are expected to
continue for ZnO loadings higher than 0.1 vol%, a hypothesis that is
indirectly supported by the trends in resistivity data. It is proposed
that with increasing ZnO loading the lms exhibit higher resistivity
principally as a result of a higher incidence of electron scattering
events attributed to grain renement, as described by the Mayadas
Shatzkes model [19]. This type of behavior has been reported
previously by Bannuru et al. [14] for AuV2O5 lms with relatively
low loadings of V2O5 in the range 05 at%, and attributed to increased
electron scattering associated with grain renement. An attempt was
made to measure grain size for a 2 vol% ZnO lm via EBSD, though
the grain sizes were below the resolution of the instrument (below
about 50 nm). This is also an indirect evidence that grain renement
continued beyond the 0.1 vol% ZnO loading. Over the range of
investigated ZnO compositions it is not expected that the percolation
theory should play a prominent role on the resistivity, as the
percolation thresholds tend to lie at signicantly higher loadings for
dispersed metal-insulator composite materials [20]. Since the writing
of this manuscript began further investigation of the resistivity of Au
ZnO composites over the entire range of compositions has been
carried out, the results of which support this hypothesis. This data
will be presented in a follow-on publication.
The observed increase in electrical resistivity does not account
for the 4 order of magnitude increase in contact resistance
measured for the 2 and 28 vol% ZnO lms. The order of magnitude
increase in resistivity between these compositions would not
contribute a 4 order of magnitude increase to the ECR of a sliding
contact, as the thin lms were on a conductive substrate and
conduction through the lms was relatively brief. It is evident
from the ECR measurements and SEM/EDS analysis that for
the 28 vol% ZnO lm there is sufcient ZnO at and near the
surface to prevent clean contact between the rider and lm from
the very rst sliding cycle. Further investigation and modeling are
required to validate this hypothesis.
Currently it is unknown whether the assumed volumetric
compositions accurately reect the relative amounts of ZnO and
Au in the lms. The cited volumetric percentage compositions are
based on calibrated rates of mass ux during the process of ebeam deposition; these calibrations are based on the linear
growth rates of lms of each material measured independently
by dual quartz crystal microbalances, each in line-of-sight of one
target. Electron microprobe analysis and cross-sectional transmission electron microscopy, for example, may be used to determine
the volume fraction and distribution of the codeposited ZnO.
4.2. Contact mechanics, wear, and transfer lms

that the rider observed the majority of wear, with more material
transfer occurring from the rider to the lm than the opposite.
Note that the wear track topographical images of the 2.0 vol% lm

The wear rate of the composite material in the applied conditions appears quite low. In unpublished work by the authors, wear

960

N. Argibay et al. / Wear 302 (2013) 955962

trailing edge

ge

sliding direction

sl

ilin

id
in

tra

ed

di
c
re

Zn rich
regions

tio

leading edge

20 m
nm

8.0

8.5

9.0

counts

Au
Zn

area 2

500
400
300
200
100
0

9.5 10.0

Au
Cu

7.5

8.0

keV

counts

-50

Zn Cu

7.5

area 1

Cu

500
400
300
200
100
0

500
400
300
200
100
0

-100

area 3

Cu

Au
Zn Cu

7.5

8.0

8.5

9.0

8.5

9.0

9.5 10.0

keV

9.5 10.0

counts

50

counts

100

Zn

500
400
300
200
100
0

area 4

Cu

7.5

8.0

keV

Au

8.5

9.0

9.5 10.0

keV

Fig. 8. (A and B) SEM backscatter micrographs of the wear scar on the riders sliding against Au2.0 vol% ZnO lm and 28.0 vol% ZnO, (C) a SWLI surface topographical map
of the rider shown in (A), and (D) EDS results for the four areas labeled in (A and B). A correction was applied to the topographical map of the rider (C) to subtract curvature.
The dark contrast areas correspond to Zn rich regions, indicating material transfer from the lms. The images showing wear and debris transfer for the rider sliding against
the 2.0 vol% lm (A and C) correspond to a total of 1000 sliding cycles, or 3.6 m of sliding after accounting for the systematic reduction in cycle length, while the damage
shown in (B) corresponds for the rider sliding against the 28.0% lm corresponds to 100 sliding cycles, or 0.5 m.

start of wear track

region of wear track after 1,000cyles

region of wear track after 1,000 cycles

Zn rich zone

line of debris along


wear track edge

25

after 100 cycles

25

after 1,000 cycles

rider sliding direction

rider sliding direction

debris are actually craters


100 m
nm

-25

100 m

to debris on rider
nm

-25

Fig. 9. (A and B) SEM backscatter images and (C and D) SWLI topographical maps of wear track regions on the Au2.0 vol% ZnO lm after 100 and 1000 sliding cycles. What
appear as dark contrast regions of debris along the edge of the wear track, similar to debris observed on the riders, are in fact impressions (craters). EDS performed on the
initial contact region of the wear track (shown in (A)) revealed a Zn rich composition.

N. Argibay et al. / Wear 302 (2013) 955962

150
100
50
0
-50
-100
20 m
-150
nm
Fig. 10. SWLI topographical maps of a Neyoro G rider wear scar after 100 cycles
sliding against a 2 mm thick Ni hardened Au electroplated lm under identical
experimental conditions to those described in the results. Signicant material
buildup was observed on the rider, in contrast to similar measurements exhibiting
rider wear while sliding against a Au2.0 vol% ZnO lm.

experiments under identical conditions implementing the same


Neyoro G riders sliding against electroplated 2 mm thick lms of
99.7 wt% Au0.3 wt% Ni hard Au (ASTM type I) on polished Alloy
52 substrates with a 5 mm Ni strike layer resulted in effectively all
of the wear occurring on the hard gold plated lms, in contrast to
the signicant rider wear observed after sliding against Au
2.0 vol% ZnO lms. Fig. 10 shows a SWLI topographical map of
the rider wear scar, exhibiting signicant material transfer to
the rider.
A classical Hertzian contact stress calculation was performed
for a 3.175 mm radius of rider tip curvature, elastic moduli for
Neyoro G (110 GPa, available from DeringerNey) and pure Au
(79 GPa), assuming an approximate Poisson ratio for both materials of 0.44, and a contact force of 100 mN. The maximum Hertzian
contact stress was approximately 185 MPa. Although a wear rate
could not be calculated for either the 2.0 vol% or 28.0 vol% ZnO
composite lms, the specic wear rate may be calculated for the
rider based on the missing volume assuming it is represented by a
spherical cap. The radius of the wear scar was approximated from
the SWLI image shown in Fig. 8 to be 15.8 mm. Assuming
uncertainties in tip radius of curvature of 0.1 mm and in wear
scar radius of 1.5 mm, the volume of missing material and
corresponding uncertainty following the method described by
Burris and Sawyer [21] for a rider tip radius of 3.175 mm and
wear scar radius of 15.8 mm are V1.5  10  8 mm3 and
u(V)0.59  10  8 mm3. The specic wear rate is then the ratio
of volume loss to the product of normal force and sliding distance.
For a normal force of 100 mN and a total sliding distance of 3.6 m
(corresponding to the rider shown in Fig. 8 and data in Fig. 7,
sliding against the 2.0 vol% ZnO composite), the specic wear rate
of the rider was K 4.3  10  8 mm3/N m. The signicance is that,
for these sliding conditions, the wear rate of the composite lms
against a relatively low wearing Au alloy rider was less than the
wear rate calculated for the rider. However, at 2.0 vol% ZnO the
ceramic loading percent is at the high end for typical metal
hardened Au alloys, typically on the order of o1 wt% (ASTM
B488-11) which may result in signicant differences in hardness
and yield strength between the two materials.

5. Summary
Electron beam deposition was used to coat conductive and
non-conductive substrates with AuZnO lms with varying concentrations of ZnO ranging from pure Au to 28.0 vol%. EBSD of

961

lm surfaces revealed a signicant reduction in grain size with


the addition of 0.1 vol% ZnO. Electrical resistivity measurements
for pure Au and composite lms with 0.128.0 vol% ZnO revealed
a linear relationship between electrical resistivity and ZnO
volume concentration, increasing from 2.73 mO cm for pure
e-beam deposited Au to 39.88 mO cm for Au28.0 vol% ZnO
composite lms. Friction coefcient values for 2.0 and 28.0 vol%
ZnO lms were in the range m 0.30.5 from the onset of sliding
(no noticeable run-in period, in contrast to the pure Au lms).
Electrical contact resistance measurements during sliding were
on the order of 40 mO, though about 1030% of the data points
acquired throughout a given cycle were suggestive of contact
separation; the standard deviation in ECR data omitting these
points of contact separation (ECR data greater than 1 O) was low
at aboutsECR 22 mO. The combined low standard deviation and
relatively high number of periodically high contact resistance
values were interpreted as the result of periodic contact separation, attributed to the presence of low conductivity ZnO debris.
SEM/EDS was used to verify the presence of Zn rich material on
the rider surface, while SWLI topographical maps showed effectively un-measurable wear of the lms and signicant transfer of
rider material to the lm surfaces, despite the presence of Zn rich
particles on the rider surface.

Acknowledgments
We would like to thank Rand Gareld for signicant contributions to the design, construction, and operation of the multifunction electrical contact tribometer used to acquire friction
coefcient and ECR data, as well as rider manufacturing and
preparation, and Richard P. Grant for SEM analysis. Sandia
National Laboratories is a multi-program laboratory managed
and operated by Sandia Corporation, a wholly owned subsidiary
of Lockheed Martin Corporation, for the US Department of
Energys National Nuclear Security Administration under contract
DE-AC04-94AL85000.

References
[1] P. Goodman, Current and future uses of gold in electronics, Gold Bulletin
35 (2002) 2122.
[2] Gold Survey, Gold Fields Mineral Services Ltd., 2001.
[3] Y. Okinaka, M. Hoshino, Some recent topics in gold plating for electronics
applications, Gold Bulletin 31 (1998) 313.
[4] M. Kato, Y. Okinaka, Some recent developments in non-cyanide gold plating
for electronics applications, Gold Bulletin 37 (2004) 3744.
[5] M. Antler, Gold-plated contacts: effect of thermal aging on contact resistance,
Plating and Surface Finishing 85 (1998) 8590.
[6] M. Antler, Tribology of metal coatings for electrical contacts, Thin Solid Films
84 (1981) 245256.
[7] M. Antler, M.H. Drozdowicz, Fretting corrosion of gold-plated connector
contacts, Wear 74 (1981) 2750.
[8] H.G. Tompkins, M.R. Pinnel, Relative rates of nickel diffusion and copper
diffusion through gold, Journal of Applied Physics 48 (1977) 31443146.
[9] H.G. Tompkins, M.R. Pinnel, Low temperature diffusion of copper through
gold, Journal of Applied Physics 47 (1976) 38043812.
[10] H.G. Tompkins, Diffusion of cobalt out of cobalt-hardened gold measured
with auger electron spectroscopy, Journal of the Electrochemical Society 122
(1975) 983987.
[11] M. Pinnel, J. Bennett, Qualitative observations on the diffusion of copper and
gold through a nickel barrier, Metallurgical and Materials Transactions A 7
(1976) 629635.
[12] M. Pinnel, Diffusion-related behaviour of gold in thin lm systems, Gold
Bulletin 12 (1979) 6271.
[13] M.-T. Lin, R.R. Chromik, N. Barbosa Iii, P. El-Deiry, S. Hyun, W.L. Brown,
R.P. Vinci, T.J. Delph, The inuence of vanadium alloying on the elevatedtemperature mechanical properties of thin gold lms, Thin Solid Films
515 (2007) 79197925.
[14] T. Bannuru, W.L. Brown, S. Narksitipan, R.P. Vinci, The electrical and mechanical
properties of AuV and AuV2O5 thin lms for wear-resistant RF MEMS
switches, Journal of Applied Physics 103 (2008) 083522-1083522-6.

962

N. Argibay et al. / Wear 302 (2013) 955962

[15] C.C. Lo, J.A. Augis, M.R. Pinnel, Hardening mechanisms of hard gold, Journal of
Applied Physics 50 (1979) 68876891.
[16] L.J. Van Der Pauw, A method of measuring the resistivity and Hall coefcient
on lamellae of arbitrary shape, Philips Technical Review 20 (1958) 220224.
[17] N. Moody, Personal Communication, Sandia National Laboratories, Livermore,
CA, 2012.
[18] E. Nes, N. Ryum, O. Hunderi, On the Zener drag, Acta Metallurgica 33 (1985)
1122.

[19] A.F. Mayadas, M. Shatzkes, Electrical-resistivity model for polycrystalline


lms: the case of arbitrary reection at external surfaces, Physical Review B 1
(1970) 13821389.
[20] D.S. McLachlan, Equations for the conductivity of macroscopic mixtures,
Journal of Physics C: Solid State Physics 19 (1986) 1339.
[21] D. Burris, W. Sawyer, Measurement uncertainties in wear rates, Tribology
Letters 36 (2009) 8187.

Vous aimerez peut-être aussi