You are on page 1of 8

Surface & Coatings Technology 318 (2017) 82–89

Contents lists available at ScienceDirect

Surface & Coatings Technology

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

Possibility of spraying of copper coatings on polyamide 6 with low


pressure cold spray method
Aleksandra Małachowska a,b,⁎, Marcin Winnicki b, Łukasz Konat b, Tomasz Piwowarczyk b, Lech Pawłowski a,
Andrzej Ambroziak b, Mateusz Stachowicz b
a
Science des Procédés Céramiques et Traitements de Surface UMR 7315, University of Limoges, CNRS 12, rue Atlantis, 87068 Limoges, France
b
Wroclaw University of Technology, ul. Lukasiewicza 5, 50-371 Wroclaw, Poland

a r t i c l e i n f o a b s t r a c t

Article history: This paper discusses metallization of polymers using a low-pressure cold spray (Dymet 413). Three commercial
Received 31 July 2016 copper powders: i) spherical and ii) dendritic were deposited on thermoplastic polymer - polyamide 6 (PA6). It
Received in revised form 30 January 2017 was difficult to successfully apply a copper coating directly on the polymer substrate, therefore interlayers were
Accepted 1 February 2017
applied. Additionally, the copper powder was pre-treated in hydrogen atmosphere to remove the oxide layer and
Available online 3 February 2017
reduce its critical velocity. Finally, the adhesion strength, electrical conductivity, oxygen content and microstruc-
Keywords:
ture of resulting coatings were determined. Coatings were characterized by one order of magnitude of lower con-
Cold spray ductivity than the bulk material and bond strength of 3.6 MPa. The powder shape turned out to have a decisive
Polymer metallization effect on the possibility on coatings formation.
Electrical conductivity © 2017 Elsevier B.V. All rights reserved.

1. Introduction equation proposed by Assadi et al. [11] and then developed by Schmidt
et al. [10]. It takes the following form:
Deposition of metal coatings onto polymers has been widely studied vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

u  
in order to improve thermal and electrical properties as well as to re- u T −T R
u F 1 ∙4∙σ TS ∙ 1− i
duce surface degradation. Some well-established processes were used t T m −T R
vcr ¼ þ F 2 ∙C pp ∙ðT m −T i Þ ð1Þ
include physical vapor deposition (PVD) [1–4] or electroless deposition ρ
[5], but the thickness of the obtained layer was limited and deposition
rates were low. Therefore, thermal spraying processes [6,7] are some- where: ρ – density, σTS – tensile strength, Tm – melting point, Ti – impact
times applied, which allow for the deposition of a thick coating on var- temperature, TR – reference temperature (293 K), cp – specific heat of
ious substrate geometries, and the recoating of damaged elements. The particle, F1 – mechanical calibration (for cold spray 1.2), F2 – thermal
main challenge is temperature-sensitivity of polymers. Recently, cold calibration (for cold spray 0.3).
spray process was tested a potential solution. In this process, material The model takes into account the specific heat, tensile strength, me-
is deposited in a solid state and therefore the temperature impact is chanical and thermal calibration, but particles size is not included [10]:
lower in comparison to traditional thermal spraying methods. The pow- with decrease in particle size critical velocity increase. The possible rea-
der particles are accelerated in stream of heated and pressurized gas son for higher critical velocity of small particles may be the higher con-
and projected towards the substrate. Critical velocity is a key concept tent of oxides or adsorbents hindering the bonding. Usually, powder
in the cold spray method [7]. It is defined as velocity that an individual contains a mixture of particles of varying diameters. In such cases, crit-
particle of powder must attain in order to be deposited after impact ical velocity is calculated for larger particles due to fact that smaller par-
with the substrate [8]. This definition is valid for ductile materials as ticles achieve a higher velocity [10].
brittle materials will cause erosion for any velocity at temperatures Lupoi and O'Neill [13] pointed out that polymers might be coated
below their melting temperature [9]. Critical velocity depends mostly when the particle impact energy calculated from the critical velocity
on the sprayed material's mechanical properties but varies also with of a given material and particle mass is sufficiently low. According to
particle size, particle morphology, particle impact temperature or pow- this formula, tin and lead will be easy to deposit, for aluminium and ti-
der oxidation [10–12]. The critical value may be calculated with the tanium deposition and erosion process will be coexistent, and for cop-
per, the erosion process will be prevalent [13]. This assumption was
⁎ Corresponding author at: Wroclaw University of Technology, ul. Lukasiewicza 5, 50-
confirmed for tin, which was deposited on various substrates PC/ABS,
371 Wroclaw, Poland. polypropylene, polystyrene and polyamide-6 [13]. The aluminium was
E-mail address: aleksandra.malachowska@pwr.wroc.pl (A. Małachowska). deposited on PEEK [14] and Lexan [15], however in case of Lexan, the

http://dx.doi.org/10.1016/j.surfcoat.2017.02.001
0257-8972/© 2017 Elsevier B.V. All rights reserved.
A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89 83

Table 1
Spray conditions for coatings deposition.

Powders Gas temperature Gas pressure Standoff distance Gun traverse speed, [mm/min] No. of spray gun passes Powder feeding rate, [g/min]
[°C] [MPa] [mm]

Cu-S, 200; 300 0.9 10 2000 1; 3 30


Cu-D

Cu-S – spherical copper powder; Cu-D – dendritic copper powder.

deposition process was possible only for high powder feed rates [15]. D10-11.3 μm, D50-22.3 μm, D90-40.6 μm (Fig. 1e), was used for
Additional spraying passes turned out to densify the aluminium coat- interlayers.
ings [16]. Thin copper coating was achieved on PA66 [17]. After deposi- Semicrystalline polyamide 6 (PA6) (Plastics Group, Poland) with
tion of the first layer, erosion occurred and further deposition was glass transition temperature ~ 50 °C and melting temperature ~220 °C
scanty. This was probably caused by a change of substrate surface to a was used as a substrate material. The substrates were in shape of a
metallic one and hence a change of parameters necessary for depositing plate with dimensions 3 × 50 × 100 mm for metallographic examina-
a coating. Ganesan et al. [18] proved that the erosion process might be tion, in the shape of a disc with Ø 40 mm in diameter and a thickness
limited by taking advantage of different shapes of copper powders and of 10 mm for bond strength test, and in the shape of a rectangle with
using tuned interlayers. Thick copper coatings were achieved on PVC 5 mm width and 80 mm length for electrical resistivity measurements.
and epoxy [18]. They observed also that metal deposition on thermoset Before spraying, the substrate was cleaned with acetone.
is more difficult because of their greater brittleness. The impact of pow-
der particles causes erosion of the substrate instead of its plastic
deformation. 2.3. Temperature-programmed reduction measurement
The deposition efficiency on polyamide 6 depends probably also on
the shape and oxidation level of the powder. It was already proved The TPR measurements were done with a 15 mg sample placed into
that some stresses remain in powder after the manufacturing process a quartz microreactor. It was made with a mixture of 5 vol% H2 in Ar, and
[19] which, together with a thin oxide layer, increases the critical veloc- the temperature was increased linearly to 950 °C at a rate of 10 °C/min.
ity. These stresses might be reduced through heat treatment which re- The hydrogen consumption was monitored by TCD detector. The
sults in lower critical velocity and hence better deposition properties. highest hydrogen consumption in case of a dendritic powder was
Therefore, in this paper the influence of heat treatment on copper depo- ±300 °C and in the case of spherical powder ±450 °C (Fig. 2). The ob-
sition on polyamide 6 is studied. tained values indicate that dendritic powders have higher oxygen con-
tent, which is caused by a developed surface. At the same time this
2. Experimental procedure oxide starts to reduce at a lower temperature.

2.1. Spraying conditions


2.4. Reduction of powder oxides
The coatings were deposited using a low-pressure cold spray device
Dymet 413 with a standard de Laval nozzle. Air was applied as working Data obtained from TPR measurements was used to reduce the ox-
gas. The spraying parameters for coating are listed in Table 1 and for in- ides present in powders. The reduction was made with hydrogen (com-
terlayers in Table 2. Two interlayers were selected based on literature: i) mercial purity 99.995%) using a self-made set-up consisting of a steel
an interlayer from spherical copper particle and ii) tin with alumina ad- tube and furnace FCF7SM (Czylok, Poland), working in the temperature
mixture. Spherical copper particles have high inertia and therefore are range 20–1150 °C with precision ±3 °C. The powder was placed in the
stable in contact with subsequent sprayed particles [18]. Tin is an easy tube through which hydrogen was flowing. The temperature was mea-
deformable material, with low critical velocity which might be deposit- sured using a thermocouple placed in the tube. The dendritic powder
ed easily on various substrates [10,13]. These two interlayers should fa- was reduced for 1 h in temperature ±300 °C and the spherical one for
cilitate copper deposition. The final coatings were sprayed with 1 h at a temperature ± 450 °C. After heat treatment the powder was
dendritic copper powder on an Sn + Al2O3 interlayer in three spraying milled and sieved to get the particle size below 40 μm. The powders
passes with gas temperature equal to 200 °C and gas pressure equal to size after reduction of powder oxides amounted to: i) spherical Cu
0.9 MPa. (Libra, Poland), D10-8.24 μm, D50-18.97 μm, D90-34.86 μm and (ii)
dendritic Cu (Libra, Poland), D10-11.18 μm, D50-34.40 μm, D90-
2.2. Spraying materials 98.51 μm. The coatings were sprayed one day after preparing the
powder.
Three commercially available copper powders were used for
spraying (Fig. 1): (i) spherical Cu (Libra, Poland), size of D10-5.94 μm,
D50-13.98 μm, D90-25.41 μm (Fig. 1a), and (ii) dendritic Cu (Libra, Po- 2.5. Oxygen content measurements
land), size of D10-10.21 μm, D50-30.67 μm, D90-80.77 μm (Fig. 1c).
Coatings were sprayed using Cu powders in delivery state and after The oxygen content in the used powder and resulted coatings was
heat treatment. Additionally, powder mixture of Sn + 50 wt% Al2O3 analysed using an O2 and N2 analyser - Leco TC436 (Leco, USA) with
(Obninsk Centre for Powder Spraying, Russia), with particle size of the inert gas fusion method.

Table 2
Spray conditions for interlayers' deposition.

Powders Gas temperature Gas pressure Standoff distance Gun traverse speed, [mm/min] No. of spray gun passes Powder feeding rate, [g/min]
[K] [MPa] [mm]

Sn + Al2O3 200 0.9 10 2000 1 30


Cu-S 200 0.9 10 2000 1 30
84 A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89

Fig. 1. Morphology of powders used for deposition process a) spherical copper particles, b) spherical copper particles after reduction of powder oxides c) dendritic copper particles d)
dendritic copper particles after reduction of powder oxides e) mixture of Sn + Al2O3.

2.6. Electrical resistivity measurement distance between the voltage-sensing pins was 20 mm. The measure-
ment system consisted of an Agilent E3632A power supply and Keithley
The resistance of the samples was measured using a custom-built 2000 and 2001 multimeters. The multimeters did not provide a test cur-
four-terminal sensing probe with spring-loaded pogo-pins. The rent big enough (less than 10 mA) to measure the resistance of the sam-
ples with satisfactory precision, therefore the power supply and one
multimeter were used to provide and measure the 100 mA test current
while the second multimeter was used to measure the voltage. Low re-
sistance of the samples (less than 100 mΩ) resulted in the power dissi-
pated in the sample being less than 1 mW. The electrical resistivity was
calculated according to the formula:

Sr U
r¼ ∙ ð2Þ
l I

where: ρ - resistivity (Ω m), U – measured voltage (V), I – measured


current (A), l - distance between measuring probe (m) Sr - the cross-
section of the conductive layer = a ∗ g, where a - width of the conduc-
tive layer (m), g – thickness of the conductive layer (m).
The thickness of the coatings for electrical resistivity amounted to
221 μm ± 22 (the thickness is given including Sn + Al2O3 interlayer).
The measured coatings were deposited in one spraying pass.

2.7. Bong strength

Fig. 2. Thermal programmed reduction of copper powder with hydrogen (TDC - thermal Bong strength of the coating to the substrate was determined from
conductivity detector). five samples in a tensile pull-off test in accordance with PN-EN 582
A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89 85

Fig. 3. Coating sprayed with spherical powder in delivery state with temperature 200 °C in a) 1 and b) 3 passes.

standard “Thermal spraying - Determination of Tensile Adhesive measurement was done on cross-sections in the middle of the coatings.
Strength”. The bond strength value was determined as the ratio of the Microhardness was given as an average of five measurements.
maximum load to the cross-sectional area of the sample. The coating
was bonded with the counter-sample using the epoxy resin adhesive
Distal. 2.9. Metallographic examination

The coatings' microstructure was examined on prepared cross-sec-


2.8. Microhardness measurements tions with electron scanning microscopes: Hitachi TM 3000 (Hitachi,
Japan) and Philips SEM XL 30 (Philips, The Netherlands).
Microhardness was measured with the use of a digital micro hard-
ness tester MMT-X7 MATSUZAWA CO., LTD (Akita, Japan) with a
Vickers indenter according to the standard PN-EN ISO 6507-3 of 2007. 3. Results and discussion
The load of 0.9807 N were chosen for each type of coating. The
3.1. Spraying with powder in delivery state

The preliminary test indicates that copper powder does not deform
sufficiently in contact with a polymer surface to form coating when
using low pressure cold spray equipment. In the case of spherical pow-
der, the particles just stuck to the substrate independent on the number
of passes (Fig. 3). Probably, the hardness and stiffness of the substrate
was too low to cause deformation or the particles velocity particles
was well below critical velocity. Moreover, in the literature it was
found that bigger particles have lower critical velocity [10], but in the
case of soft substrate it seems that only very small particles can deform.
Additionally, although there is a high difference in tensile strength and
Young's modulus between the polyamide 6 and the copper, the copper
particles are not driven deep into the substrate and rather remain on the
surface.
Similar phenomena take place for higher gas inlet temperatures. The
continuous layer has not been formed (Fig. 4). However, the particles
are driven deeper into the substrate are thicker in comparison to the
lower inlet temperature.
Continuous coating was also not formed in case of spraying with
Fig. 4. Coating sprayed with spherical powder in delivery state with temperature 300 °C in dendritic powder, but the powder particles are deformed more inten-
three passes. sively in compare to spherical particles (Fig. 5).

Fig. 5. Coating sprayed with dendritic powder in delivery state with temperature 200 °C in a) one and b) three passes.
86 A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89

3.2. Spraying with powder in delivery state on interlayers aluminium particles were deposited on the substrate, but whether they
formed a bond or were just embedded in soft tin remains unclear.
The solution proposed by Ganesan et al. [18] to use spherical parti- Deposited copper coatings were well-bonded with the substrate.
cles as interlayers was tested. The dendritic particles were sprayed on Their morphology depends mostly on the number of spraying passes;
the previous deposited spherical ones but without success. It can be coatings deposited in several passes were densified by a hammering ef-
seen that dendritic particles fragmented and partially deformed, but fect. It is especially noticeable in coating deposited in three passes, in
no bonding may be found (Fig. 6a). The interlayer was probably too which the main porosity area is situated in the top region, which is typ-
loose and the velocity of the dendritic particles too low. As a conse- ical for the cold sprayed method [22,23] (Fig. 8). In coating deposited in
quence, the spherical particles were displaced by subsequent dendritic one pass, porosity near the substrate might be observed which unusual
particles. There are only a few places in which the spherical particles for a cold spray process [23]. Also, morphology of the powder probably
remained. contributed to the porosity as impact energy was not accumulated in
Tin was tested as the interlayer because it poses the lowest critical one contact point.
velocity among usual cold spray feedstock material. As a variation,
Sn + A2O3 was also applied. The addition of A2O3 is commonly in the 3.4. Oxidation measurements
low pressure cold spray method: it hammers additionally already de-
posited metal particles, sandblasts the substrate, and prevents nozzle Among powders, oxygen content was lowest for the spherical one -
clogging [20]. Dendritic and spherical copper powders were used as 0.042 wt% and increased with irregularity of the powder morphology to
feedstock material. A continuous layer has been not achieved; upcom- 0.065 wt% (Fig. 9) for the dendritic powder. Powders after heat treat-
ing copper particles removed the interlayer from the substrate, leaving ment in hydrogen and storing for one day exhibit about 1/3 lower oxi-
only small amount of tin (Fig. 6b,c). dation than the initial one. The oxygen value for dendritic copper
amounted to 0.043 wt%. The deposition efficiency measured for dendrit-
ic copper after deoxidation amounted to 30.95 ± 4%. The deposition ef-
3.3. Spraying with powder after heat treatment ficiency for rest of the powders oscillated around the level of 0%.
An initial oxide layer, even at room temperature, forms rapidly on
When using powders after deoxidation higher deformation was copper. It was reported that after half an hour an initial oxide layer of
clearly observed, but there is still no bonding between adjoining parti- 1 nm is formed, then oxide growth speed slows down and after
cles (Fig. 7). The highest deformation grade was observed for dendritic 11 days the layer has a thickness of about 4.5 nm [24]. The changing
powder especially for higher number of spraying passes, which enables of oxidation rates is prescribed to the coalescence of initial oxide islands,
further for deposition adherent coating using Sn + Al2O3 interlayer which block surface diffusion routes; further oxide growth requires
(Fig. 8). Despite using of Sn + Al2O3 interlayer it was not possible to ob- much slower bulk diffusion [25]. Initial oxide layer thickness equal to
tain continuous coating with spherical powder (Fig. 7f). 1 nm might be recalculated for 20 μm spherical copper particle to
Dendritic copper was successfully deposited on PA6 with the 0.010 wt% (taking copper oxide density ~6.14 g/cm3). This oxygen con-
Sn + A2O3 interlayer. The influence of the substrate nature was tent appears after half an hour contact with air so it can be hardly
highlighted by Zhang [21]. He observed scant deposition of aluminium avoided. Taking into account necessary milling and sieving of powder
particles on the tin substrate. This was attributed to a low melting tem- after heat treatment, obtained values of oxygen content seem to be cor-
perature point; during impact, tin melted, losing its strength, and there- rect. The oxygen content derived from initial oxidation decreases with
fore aluminium particles rebounded easily. Only small amounts of small increasing the size of the particle. Due to developed surface area, copper

Fig. 6. Coating sprayed with dendritic powder in delivery state on: a) spherical copper particle, b) Sn + Al2O3 interlayer (gas temperature 200 °C) and c) spherical powder in delivery state
on Sn + Al2O3 interlayer (gas temperature 200 °C).
A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89 87

Fig. 7. Coating sprayed with: spherical copper particles after deoxidation in a) one and b) three passes and f) on Sn + Al2O3 interlayer; dendritic copper particles after deoxidation in c) one
and d) three passes and e) on spherical particles interlayer.

amounts in dendritic or globular powder are higher, which is also noted spray coatings is at the same level or even lower than feedstock material
in the measurement. Powder properties will change slowly after initial due to separation of the oxide layer during impact. This crushed oxide is
oxide formation, following logarithmic law [26]. Such behaviour was partially removed and partially remains in formed coating [30]. Howev-
observed for dendritic copper particles by Ko et al. [27]. Powder er, in this study oxygen amount in the coating was roughly two times
annealed in a vacuum reached 60% spraying efficiency, which decreased higher than for powder material in the case of copper. This additional
to 40% for two weeks' air exposure and after which it remained a six- oxidation occurs partially during the spraying process and partially dur-
week period. Oxygen content after annealing was 0.066 wt% [27]. ing cooling of the sample, so it will also affect the deposition process. As
Particle oxidation has notably influence on its deposition behaviour. it was shown for Ti particles only during their presence in the nozzle
Increase of critical velocity from 310 m/s to 550 m/s for oxygen content will the oxide growth be ~ 0.3 nm for a 10 μm particle [31]. A high
0.02 to 0.14 wt% was reported by Li et al. [28]. This might provide an ex- amount of oxides in coatings might also be the result of low velocity
planation for coating formation when spraying with powder after first which was insufficient to break them. A similar percent of oxygen con-
deoxidizing the powder with hydrogen, but the difference of 0.02 wt% tent in all coating was reported by Li [12], however a much higher
in oxygen content for heat-treated and untreated powder is small. spraying temperature was used at 520 °C.
Kang et al. [29] claimed that the relation between oxygen content and Heat treatment of powder affects not only the oxide layer but also
the increase in critical velocity is linear for aluminium powder. The low- the hardness of the powder [19]. The ultimate tensile strength of bulk
est critical velocity values were estimated to be 721 m/s for 0.001 wt%, copper material amounted to 300 MPa and increased to 360 MPa for
and increased to 867 m/s for 0.045 wt% oxygen content and are much copper in the form of powder [19]. Therefore, even heat treatment in
higher than given by Schmidt [9]. air, which causing additional oxidizing of the feedstock material, in-
Taking into account the linear relation for data given by Li [28], the creases spraying efficiency to a certain degree [32]. The increase of de-
increase of 0.02 wt% in oxygen content, which is visible between pow- position efficiency was, however, higher for powder treated in a
der in delivery state and after heat treatment, will cause critical velocity vacuum [32] The benefits of using heat treatment were extremely im-
to increase at the level of 40 m/s. Additionally, it must be noted that in portant for spraying copper coatings of polymers as it requires a very
the case of a low-pressure system, air is used as the propellant gas. narrow window of deposition and it was not possible to achieve the
This results in additional oxidation of powder particles despite the low coating without the heat treatment of powders. The oxygen content re-
temperature of the process. It was claimed that oxygen content in cold corded for coatings should have limited influence on electrical
88 A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89

Fig. 8. Coating sprayed with dendritic powder on an Sn + Al2O3 interlayer: a) one spraying pass, b) three spraying passes.

conductivity, i.e. for copper, an oxygen content of ~0.20% decreased the effect of gas pressure on deposition efficiency was observed [18]. It
electrical conductivity value to 99% [33]. could possibly be caused by lower oxidation of powder particles while
spraying (due to use of a nitrogen instead of air), which facilitates con-
3.5. Electrical resistivity measurements tact to clean metallic surfaces. As a consequence, a greater number of
metallic bonds a created.
The electrical resistivity of copper coating amounted to ~26 μΩ cm In the literature it is claimed that cold-sprayed coatings may reach
(approximately one order of magnitude higher than bulk material). even 90% of conductivity of bulk copper [37]. On this background, ob-
The low electrical conductivity was probably caused by presence of po- tained results are limited in number. One on the ways to improve the
rosity, which indicates small areas of metallic boding due to insufficient coatings' conductivity is heat treatment of coatings after spraying.
velocity. The resistivity values for all copper coatings are in the range Koivuluoto et al. [38] reported the conductivity values to be 79% IACS
given by Sudharshan [34] (24.39 μΩ cm) for following spraying param- for HPCS, and 46% IACS for LPCS-sprayed coatings; they increased to
eters gas pressure: 1.4 MPa and a temperature of 300 °C, but two times 90% IACS for HPCS and 69% IACS after heat treatment (400 °C, 2 h).
higher than this given by Ganesan [35] for copper coating on PVC at The coatings were sprayed with spherical powders. However, heat
200 °C [36] at a pressure of 2 MPa [18,35]. The difference could only treatment was not applied due to the nature of the substrate.
be partially explained through pressure difference because almost no

Fig. 9. Oxygen content in feedstock material and coatings. Fig. 10. Fracture after bond strength test.
A. Małachowska et al. / Surface & Coatings Technology 318 (2017) 82–89 89

3.6. Mechanical properties of coatings More studies are needed to further assess the influence of initial
feedstock properties on the deposition process on the polyamide 6 sub-
The bond strength of copper coatings amounted to 3.6 ± 0.2 MPa. strate. Future plans, including measurement of microhardness of pow-
The fractures occurred partially in the interlayer and partially in the der before and after spraying as a function of the oxidation level. It is
coatings (Fig. 10). Koivuluoto et al. [20] reported bond strength of den- also planned to develop a numerical model for investigation of the con-
dritic copper sprayed with a low-pressure cold spray (pressure 0.6 MPa tact mechanism of different powder particles.
and temperature 540 °C) to be ~7.5 MPa. This increased significantly to
~17.5 MPa with the addition of 50% Al2O3. Obtained bond strength value
amounts ~50% of it. Acknowledgements
The bonding mechanism and its strength is usually attributed to two
main mechanisms: formation of shear instabilities in particle/substrate This research was supported by NCN (Narodowe Centrum Nauki),
and particle/particles interface caused by extensive plastic deformation (grant no. UMO-2014/15/N/ST8/02661).
during impact [11,39] and mechanical interlocking [40]. However, in
the case of polymer/metal bonding, only the second one may appear References
due to the different nature of the coating and substrate material. Addi-
[1] C. Maurer, U. Schulz, Wear 317 (2014) 246–253.
tionally “interlocking might not be as strong as the one observed on [2] U. Schulz, Appl. Opt. 45 (2006) 1608–1618.
the metals substrate due to the soft nature of the polymer” [35]. The [3] J.-H. Yang, Y.-S. Han, J.-H. Choy, Thin Solid Films 495 (2006) 266–271.
other problem [35,41] when creating bonding was attributed to the [4] R. Suchentrunk, Kunststoff-Metallisierung: Mit 35 Tabellen, 3. Aufl, Schriftenreihe
Galvanotechnik und Oberflächenbehandlung, 37, Leuze, Bad Saulgau, 2007.
squeezed out polymer during particle impact which might partially en- [5] C.K. Lee, Mater. Chem. Phys. 114 (2009) 125–133.
close the metal particles and acts as a barrier against bonding formation. [6] R. Knight, M. Ivosevic, S. Kalinidi, G. Palmese, Proceedings of the ITSC 2003, 2003.
This was claimed to happen when using spherical copper as an interlay- [7] D.L. Gilmore, R.C. Dykhuizen, R.A. Neiser, M.F. Smith, T.J. Roemer, J. Therm. Spray
Technol. 8 (1999) 576–582.
er [35]. Tin turned out to have a more beneficial influence on adhesion [8] V.K. Champagne, The Cold Spray Materials Deposition Process: Fundamentals and
strength [18,35]. Ganesan et al. attributed this to the low mechanical Applications, Woodhead; CRC Press, Cambridge, Boca Raton, 2007 (http%3A//
properties of tin particles which were able to undergo the deformation www.worldcat.org/oclc/191557639).
[9] T. Schmidt, H. Assadi, F. Gärtner, H. Richter, T. Stoltenhoff, H. Kreye, T. Klassen, J.
instead of damaging the surface. This initial interlayer then enabled
Therm. Spray Technol. 18 (2009) 794–808.
bonding or interlocking for subsequent layers with higher mechanical [10] T. Schmidt, F. Gärtner, H. Assadi, H. Kreye, Acta Mater. 54 (2006) 729–742.
properties i.e. copper [18,35]. In the case of spherical copper particles, [11] H. Assadi, F. Gärtner, T. Stoltenhoff, H. Kreye, Acta Mater. 51 (2003) 4379–4394.
the shear strength of the copper coating was 1.93 ± 0.7 MPa in compar- [12] W.-Y. Li, C. Zhang, H.-T. Wang, X.P. Guo, H.L. Liao, C.-J. Li, C. Coddet, Appl. Surf. Sci.
253 (2007) 3557–3562.
ison to 5.4 ± 0.9 MPa for the tin interlayer [35]. Obtained bond strength [13] R. Lupoi, W. O'Neill, Surf. Coat. Technol. 205 (2010) 2167–2173.
lies in the middle of these two values. [14] X.L. Zhou, A.F. Chen, J.C. Liu, X.K. Wu, J.S. Zhang, Surf. Coat. Technol. 206 (2011)
Deposited copper coatings reached a microhardness of 125HV0.1 ± 132–136.
[15] H. Ye, J. Wang, Mater. Lett. 137 (2014) 21–24.
12 on PA6. This value is much lower than that given by Sudharshan - [16] D. Giraud, F. Borit, V. Guipont, M. Jeandin, J.M. Malhaire, Proceedings of the ITSC,
300 or even 450HV0.1 for copper coatings on aluminium [34] but simi- 2012 265–270 (Houston, Texas, USA).
lar to values given by [20,35]. However, due to different loads used for [17] M. Barletta, A. Gisario, V. Tagliaferri, Surf. Coat. Technol. 201 (2006) 296–308.
[18] A. Ganesan, M. Yamada, M. Fukumoto, J. Therm. Spray Technol. 22 (2013)
measurement, the values can be not be directly compared. The low 1275–1282.
values of copper microhardness might be prescribed to low a deforma- [19] H. Assadi, I. Irkhin, H. Gutzmann, F. Gärtner, M. Schulze, M. Villa Vidaller, T. Klassen,
tion grade of feedstock material. Additionally, lower microhardness of Adv. Powder Technol. 26 (2015) 1544–1554.
[20] H. Koivuluoto, P. Vuoristo, J. Therm. Spray Technol. 19 (2010) 1081–1092.
feedstock after heat treatment results in lower values of coatings micro- [21] D. Zhang, P.H. Shipway, D.G. McCartney, J. Therm. Spray Technol. 14 (2005)
hardness [32]. 109–116.
[22] L. Ajdelsztajn, J. Schoenung, B. Jodoin, G. Kim, Metall. Mater. Trans. A 36 (2005)
657–666.
3.7. Conclusions
[23] T.H. van Steenkiste, J.R. Smith, R.E. Teets, Surf. Coat. Technol. 154 (2002) 237–252.
[24] S. Suzuki, Y. Ishikawa, M. Isshiki, Y. Waseda, Mater. Trans. JIM 38 (1997) 1004–1009.
Copper coatings were successfully deposited on polyamide 6 with [25] J.C. Yang, B. Kolasa, J.M. Gibson, M. Yeadon, Appl. Phys. Lett. 73 (1998) 2841.
low-pressure cold spray using an Sn + Al2O3 interlayer and dendritic [26] J. Iijima, J.-W. Lim, S.-H. Hong, S. Suzuki, K. Mimura, M. Isshiki, Appl. Surf. Sci. 253
(2006) 2825–2829.
copper powder. The following specific conclusions have been drawn [27] K.H. Ko, J.O. Choi, H. Lee, B.J. Lee, Powder Technol. 218 (2012) 119–123.
based on the research: [28] C.-J. Li, H.-T. Wang, Q. Zhang, G.-J. Yang, W.-Y. Li, H.L. Liao, J. Therm. Spray Technol.
19 (2010) 95–101.
- The possibility of copper coating formation on the polymer is related [29] K. Kang, S. Yoon, Y. Ji, C. Lee, Mater. Sci. Eng. A 486 (2008) 300–307.
to powder morphology and oxygen content in the powder. Dendritic [30] W.-Y. Li, H. Liao, C.-J. Li, H.-S. Bang, C. Coddet, Appl. Surf. Sci. 253 (2007) 5084–5091.
[31] A. Małachowska, M. Winnicki, L. Pawłowski, B. Pateyron, A. Ambroziak, P.
powders, owing to having several contact places with the substrate Sokołowski, Proceedings of COMSOL Conference, Rotterdam, October 23–25, 2013.
surface, limit erosion processes during impact of the particles and [32] K.H. Ko, J.O. Choi, H. Lee, J. Mater. Process. Technol. 214 (2014) 1530–1535.
enable deposition of copper coatings on the polymer. On the other [33] J.R. Davis, A. Committee, Copper and Copper Alloys, ASM International, 2001
(https://books.google.pl/books?id=sxkPJzmkhnUC).
hand, this phenomenon causes porosity in the obtained coatings. [34] P. Sudharshan Phani, D. Srinivasa Rao, S.V. Joshi, G. Sundararajan, J. Therm. Spray
- Velocities obtained using a low-pressure cold spray are too small to Technol. 16 (2007) 425–434.
obtain porous-free copper coatings on polymers. This unfavourable [35] A. Ganesan, J. Affi, M. Yamada, M. Fukumoto, Surf. Coat. Technol. 207 (2012)
262–269.
effect is additionally strengthened by low substrate and interlayer
[36] W.B. Choi, L. Li, V. Luzin, R. Neiser, T. Gnäupel-Herold, H.J. Prask, S. Sampath, A.
hardness. Gouldstone, Acta Mater. 55 (2007) 857–866.
- Electrical conductivity amounted to 26 μΩ cm, which is one order of [37] T. Stoltenhoff, H. Kreye, H.J. Richter, J. Therm. Spray Technol. 11 (2002) 542–550.
[38] H. Koivuluoto, A. Coleman, K. Murray, M. Kearns, P. Vuoristo, J. Therm. Spray
magnitude higher than the bulk material. This low electrical conduc-
Technol. 21 (2012) 1065–1075.
tivity is caused by insufficient metallic bonding. It is probable that [39] M. Grujicic, C. Zhao, W. DeRosset, D. Helfritch, Mater. Des. 25 (2004) 681–688.
using nitrogen as the propellant gas could improve conductivity [40] T. Hussain, D.G. McCartney, P.H. Shipway, D. Zhang, J. Therm. Spray Technol. 18
owing to the reduction of oxygen content, and hence easier forma- (2009) 364–379.
[41] I. Burlacov, J. Jirkovský, L. Kavan, R. Ballhorn, R.B. Heimann, J. Photochem. Photobiol.
tion of metallic bonding. A Chem. 187 (2007) 285–292.

You might also like