Вы находитесь на странице: 1из 4

Wear 374-375 (2017) 1–4

Contents lists available at ScienceDirect

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

Case Study

Wear stability of superhydrophobic nano Ni-PTFE electrodeposits


n
Jason Tam, Zhonghui Jiao, Jonathan Chun Fung Lau, Uwe Erb
Department of Materials Science and Engineering, University of Toronto, 184 College Street, Toronto, Ontario, Canada M5S 3E4

a r t i cle in f o a b s t r a c t

Article history: Commercialization of superhydrophobic surfaces remains rather limited due to the high costs of many
Received 12 October 2016 fabrication processes and the poor understanding of the long term stability of their non-wetting prop-
Received in revised form erties. In this study, degradation of a recently developed superhydrophobic nanocrystalline Ni-PTFE
14 December 2016
composite made by electrodeposition was evaluated by a linear abrasion test on SiC abrasive media. For
Accepted 16 December 2016
comparison, the same abrasion test was also conducted on a commercially available superhydrophobic
Available online 27 December 2016
spray-on treatment to benchmark the wear performance of the Ni-PTFE composite. Due to the micro-
Keywords: structure of the Ni-PTFE composite coating, remarkable non-wetting properties were observed after
Surface engineering
extensive abrasive wear.
Composite coating
& 2016 Elsevier B.V. All rights reserved.
Electrodeposition
Superhydrophobicity
Wettability
Abrasive wear

1. Introduction blades [19], the sand impact test [20,21], and the water jet test
[22]. Progress in the current understanding of the wear properties
Over the past decade, there have been many studies on re- of such surfaces has recently been reviewed for various wear
producing the superhydrophobic properties of some biological testing techniques [12]. Among these techniques, the most com-
surfaces such as the lotus leaf on engineering materials. Super- mon and well-accepted wear test is linear abrasion.
hydrophobic surfaces have a high water contact angle (WCA), In this study, the wear resistance of recently developed elec-
usually greater than 150° and a low water sliding angle (WSA) [e.g. trodeposited nanocrystalline nickel-polytetrafluoroethylene (Ni-
1]. Owing to the high mobility of water droplets on a super-
PTFE) composite coatings [10] was evaluated by a simple linear
hydrophobic surface, there are many potential applications such as
abrasion test. The composite coatings consist of hydrophobic PTFE
self-cleaning [2], anti-fouling [3], anti-corrosion [4] and anti-icing
particles with a bimodal particle size distribution embedded
surfaces [5]. Currently, superhydrophobic surfaces can be fabri-
cated using a wide range of techniques including but not limited to throughout the entire thickness of a very hard nanocrystalline Ni
lithography [6,7], templating [8,9], and electrodeposition [4,10]. In matrix coating [10].
terms of industrial applicability, electrodeposition is one of the
most promising techniques as it is a relatively simple, scalable, and
inexpensive process that can be applied to a broad range of ma-
terials including metals, polymers, and composites [11]. Although 2. Materials and methods
there has been significant progress in the development of super-
hydrophobic surfaces, the long term stability of these surfaces in 2.1. Specimen preparation
the context of the effect of surface wear on the non-wetting
properties is not well researched. Currently, there are no standards Ni-PTFE composite coatings were electrodeposited to a thick-
to evaluate the durability of superhydrophobic surfaces as they are ness of 60 μm onto polished copper substrates (2 cm 2 cm) ac-
exposed to wear [12]. However, many different testing techniques cording to a co-deposition technique established in a previous
and approaches have been developed by various researchers, for study [10]. In addition, a commercial superhydrophobic surface
instance, linear abrasion of superhydrophobic surfaces with dif- treatment (CSHST), spray-on NeverWet™, was applied on polished
ferent grits of sandpapers [4,13–18], a scratch test with sharp copper substrates (2 cm 2 cm) following the manufacturer's in-
structions. The as-sprayed CSHST coating was about 7 μm. CSHST
n coated copper specimens were used as a comparison to the Ni-
Corresponding author.
E-mail address: uwe.erb@utoronto.ca (U. Erb). PTFE composites.

http://dx.doi.org/10.1016/j.wear.2016.12.023
0043-1648/& 2016 Elsevier B.V. All rights reserved.
2 J. Tam et al. / Wear 374-375 (2017) 1–4

placing 5 mL droplets of deionized (DI) water onto the surface. To


study the adhesion of water droplets on the surface, the water
sliding angles (WSA) of 25 mL DI water droplets were determined
by placing the specimen on a tilting stage.
Fig. 1. Schematic diagram of simple abrasion wear test apparatus.

2.2. Abrasion wear testing 3. Results and discussion

In this study, a simple abrasion wear testing apparatus was 3.1. Surface morphology and wetting properties of as-prepared
developed, as depicted in Fig. 1. The sample was pushed in a re- specimens
ciprocating motion at a rate of 10 cm/s across the surface of SiC
paper (400 and 800 grit) under a constant applied pressure of Scanning electron micrographs of the as-prepared Ni-PTFE
2000 Pa perpendicular to the surface. The abrasion stroke length composite and CSHST coatings are shown in Fig. 2a, d. For the Ni-
was 15 cm. For each sample, a fresh SiC paper was used in order to PTFE composite, dual scale surface roughness with lotus leaf-like
compare the results with previous studies [4,13,14]. The chosen morphology was achieved by co-depositing PTFE powder with
parameters, including applied pressure and grit size of SiC papers bimodal particle size distribution (average particle size diameters
were similar to several past studies on the abrasive wear of elec- of 8 mm and 0.6 mm, respectively). The high magnification image of
trodeposited superhydrophobic surfaces [4,13,14]. as-prepared Ni-PTFE composite (Fig. 2a, inset) shows smaller PTFE
particles embedded in the fine grain structure of the nanocrys-
2.3. Coating adhesion testing talline Ni matrix. Using EDS, the composite coatings were de-
termined to contain 70 vol% PTFE. On the other hand, nano-
A robust coating for practical application also requires strong particles bonded together forming a porous structure were ob-
adhesion to the underlying substrate. In addition to abrasive wear served on the CSHST coating (Fig. 2d). According to the material
testing, the adhesion of the superhydrophobic coatings on the safety data sheet, the CSHST coatings contain silicone and silica
substrate was also evaluated by a cross-cut tape test per ASTM [24]. Based on this finding, it can be speculated that the nano-
D3359-09 standard using square grid cross-cuts all the way down particles observed in the micrographs are silica nanoparticles that
to the substrate at 1 mm spacings [23]. An optical microscope was are held in clusters by hydrophobic silicone. Both of the as-pre-
employed to analyze the coatings before and after the adhesion pared specimens demonstrated superhydrophobic properties with
test. WCA 4 150° and a low WSA, less than 5° (Fig. 3).

2.4. Microstructure and wetting characterization 3.2. Coating adhesion

Surface morphology and composition of the specimens were Both Ni-PTFE composite and CSHST coatings exhibited excellent
characterized using a Hitachi SU8230 scanning electron micro- adhesion to the substrate. All lattice squares remained intact and
scope (SEM) equipped with an energy dispersive X-ray spectro- no noticeable removal of the coating from the substrate was ob-
meter (EDS). Water contact angles (WCA) were measured by served after the cross-cut tape test. As per ASTM D3359 standard,

Fig. 2. Electron micrographs of superhydrophobic coatings. (a) As-prepared Ni-PTFE, showing dual scale roughness. Inset shows submicron PTFE particles embedded in the
nanocrystalline Ni matrix, scale bar: 500 nm. (b) Ni-PTFE after 3 m of wear on 400 grit SiC paper. Wear grooves generated from the SiC abrasive can be observed. (c) Ni-PTFE
after 24 m on 800 grit, showing finer wear grooves and deformed PTFE particles. (d) As-prepared CSHST with porous structure. Inset shows clusters of nanoparticles forming
the porous structure, scale bar: 250 nm. (e) CSHST after 3 m on 400 grit. Patches of delaminated coating can be observed. (f) CSHST after 24 m on 800 grit, showing fine wear
grooves and smaller patches of exposed Cu substrate.
J. Tam et al. / Wear 374-375 (2017) 1–4 3

Fig. 3. Effect of abrasive wear on WCA (a, b) and WSA (c,d) of Ni-PTFE composite coatings and CSHST coatings. (a, c): 400 grit SiC. (b, d): 800 grit SiC. Error bars show the
standard deviation of the measurements.

both types of superhydrophobic coatings achieved the highest (Cu WCA: 52°) and the superhydrophobic CSHST coating.
classification of 5B with 0% of the coating area removed. On 800 grit medium, extended stability of superhydrophobic
properties was observed for Ni-PTFE composite; WCA remained
3.3. Wear stability on 400 and 800 grit SiC papers above 150° after 48 m of abrasion (Fig. 3b). In contrast, durability
of the CSHST was limited as the WCA dropped to 141° after only
As shown in Fig. 3a, WCA of Ni-PTFE composites remained 3 m of abrasive wear. Moreover, water droplets did no longer slide
above 150° and WSA was below 15° for about 3 m of abrasion off the surface after 24 m of abrasion (Fig. 3d). At the conclusion of
length on 400 grit SiC. The WCA gradually decreased to about 130° the wear test, CSHST specimens were barely hydrophobic as the
and WSA converged to 50° after 18 m of abrasion length. On the WCA dropped to 92°.
other hand, the CSHST specimens were no longer super- Micrographs of both specimens after 24 m of abrasive wear are
hydrophobic when the surface was abraded for only 0.75 m and shown in Fig. 2c, f. Fine wear groves and deformed micron size
the mobility of water droplets was strongly reduced as shown by protrusions were observed on the Ni-PTFE composite (Fig. 2c).
the rapid rise in WSA from 3° to 70° (Fig. 3a, c). Furthermore, Although top portions of the large scale roughness were damaged,
water droplets did not slide off the surface even when the samples superhydrophobic properties were retained and trapped air
were turned upside down for abrasion lengths greater than 3 m. pockets can still form within the submicron-scale roughness that
Micrographs of the specimens after 3 m of abrasive wear are was protected from abrasion by the larger PTFE particles and the
shown in Fig. 2b, e. Wear tracks and deformed micron size PTFE hard nanocrystalline Ni matrix of the coating [10]. Similar to the
particles caused by the abrasion can be observed for the Ni-PTFE observation made with the coarser SiC abrasive medium, patches
composite (Fig. 2b). Due to the hierarchical surface roughness, of exposed hydrophilic copper substrate were observed on the
submicron size PTFE particles in the recessed areas were un- CSHST following 24 m of wear (Fig. 2f). These hydrophilic spots act
damaged which allowed for a stable Cassie-Baxter wetting state to as pinning points for water droplets, which led to the high WSA
be maintained even with some larger scale surface damage. In (Fig. 3d).
contrast, delamination wear was observed for the CSHST speci- Furthermore, the wear stability of electrodeposited Ni-PTFE on
mens (Fig. 2e); portions of the coating were stripped off, exposing 800 grit SiC was compared to other superhydrophobic electro-
the underlying copper substrate. As a result of the surface in- deposits that were previously tested under similar conditions
homogeneity, large standard deviations of WCA and WSA mea- [4,13,14]. As shown in Table 1, the total abrasion length before the
surements were observed after abrasive wear (Fig. 3) as the water loss of superhydrophobicity for the Ni-PTFE composite is sig-
droplets come in contact with both the hydrophilic Cu substrate nificantly higher than for the other electrodeposited coatings
4 J. Tam et al. / Wear 374-375 (2017) 1–4
Table 1 roughness in antifouling, Biofouling 25 (2009) 757–767, http://dx.doi.org/
Abrasion lengths leading to the loss of superhydrophobic properties. Abrasive 10.1080/08927010903165936.
medium: 800 grit SiC. [4] F. Su, K. Yao, Facile fabrication of superhydrophobic surface with excellent
mechanical abrasion and corrosion resistance on copper substrate by a novel
Material Pressure (Pa) Abrasion Initial Final method, ACS Appl. Mater. Interfaces 6 (2014) 8762–8770, http://dx.doi.org/
Length (m) WCA (°) WCA (°) 10.1021/am501539b.
[5] L. Cao, A.K. Jones, V.K. Sikka, J. Wu, D. Gao, Anti-Icing superhydrophobic
coatings, Langmuir 25 (2009) 12444–12448, http://dx.doi.org/10.1021/
Ni-PTFE composite 2000 50 156 150
la902882b.
[this study] [6] B. Bhushan, K. Koch, Y.C. Jung, Fabrication and characterization of the hier-
CSHST [this study] 2000 3 155 141 archical structure for superhydrophobicity and self-cleaning, Ultramicroscopy.
Fluorinated Ni [4] 6000 1 162 148 109 (2009) 1029–1034, http://dx.doi.org/10.1016/j.ultramic.2009.03.030.
Co with stearic acid 1500 1.1 156 148 [7] M. Im, H. Im, J.-H. Lee, J.-B. Yoon, Y.-K. Choi, A robust superhydrophobic and
[13] superoleophobic surface with inverse-trapezoidal microstructures on a large
CuO with stearic acid 1200 0.7 163 140 transparent flexible substrate, Soft Matter 6 (2010) 1401, http://dx.doi.org/
[14] 10.1039/b925970h.
[8] J.J. Victor, D. Facchini, U. Erb, A low-cost method to produce superhydrophobic
polymer surfaces, J. Mater. Sci. 47 (2012) 3690–3697, http://dx.doi.org/
studied to date owing to the hierarchical surface roughness and 10.1007/s10853-011-6217-x.
[9] X. Sheng, J. Zhang, Superhydrophobic behaviors of polymeric surfaces with
the continual exposure of new hydrophobic PTFE particles as the aligned nanofibers, Langmuir 25 (2009) 6916–6922, http://dx.doi.org/10.1021/
surface is exposed to wear. la9002077.
[10] D. Iacovetta, J. Tam, U. Erb, Synthesis, structure, and properties of super-
hydrophobic nickel–PTFE nanocomposite coatings made by electrodeposition,
Surf. Coat. Technol. 279 (2015) 134–141, http://dx.doi.org/10.1016/j.
4. Conclusions surfcoat.2015.08.022.
[11] J. Tam, G. Palumbo, U. Erb, Recent advances in superhydrophobic electro-
deposits, Mater. (Basel) 9 (2016) 151, http://dx.doi.org/10.3390/ma9030151.
Robustness of the non-wetting properties of electrodeposited [12] A. Milionis, E. Loth, I.S. Bayer, Recent advances in the mechanical durability of
nanocrystalline Ni-PTFE composite coatings was evaluated by a superhydrophobic materials, Adv. Colloid Interface Sci. 229 (2016) 57–79, http:
simple linear abrasion test. Owing to the microstructure of the Ni- //dx.doi.org/10.1016/j.cis.2015.12.007.
[13] W. Li, Z. Kang, Fabrication of corrosion resistant superhydrophobic surface
PTFE composite, namely the dual-scale surface roughness formed
with self-cleaning property on magnesium alloy and its mechanical stability,
by the embedded PTFE particles in the hard nanocrystalline Ni Surf, Coat. Technol. 253 (2014) 205–213, http://dx.doi.org/10.1016/j.
matrix, the composite coatings demonstrated excellent resistance surfcoat.2014.05.038.
[14] C. Tan, Q. Li, P. Cai, N. Yang, Z. Xi, Fabrication of color-controllable super-
to the degradation of non-wetting properties as the surface is
hydrophobic copper compound coating with decoration performance, Appl.
subjected to abrasive wear. Compared to a commercially available Surf. Sci. 328 (2015) 623–631, http://dx.doi.org/10.1016/j.apsusc.2014.12.025.
superhydrophobic spray coating and other electrodeposited su- [15] K. Chen, S. Zhou, L. Wu, Facile fabrication of self-repairing superhydrophobic
perhydrophobic surfaces, electrodeposited Ni-PTFE maintained coatings, Chem. Commun. 50 (2014) 11891–11894, http://dx.doi.org/10.1039/
c3cc49251f.
superhydrophobicity for significantly longer abrasion lengths. The [16] Y.-Y. Zhang, Q. Ge, L.-L. Yang, X.-J. Shi, J.-J. Li, D.-Q. Yang, et al., Durable su-
results from this study demonstrate the feasibility of applying this perhydrophobic PTFE films through the introduction of micro- and nanos-
type of superhydrophobic coating in practical applications that tructured pores, Appl. Surf. Sci. 339 (2015) 151–157, http://dx.doi.org/10.1016/
j.apsusc.2015.02.143.
require excellent wear stability. [17] X. Zhu, Z. Zhang, J. Yang, X. Xu, X. Men, X. Zhou, Facile fabrication of a su-
perhydrophobic fabric with mechanical stability and easy-repairability, J.
Colloid Interface Sci. 380 (2012) 182–186, http://dx.doi.org/10.1016/j.
jcis.2012.04.063.
Acknowledgments
[18] X. Zhu, Z. Zhang, X. Men, J. Yang, K. Wang, X. Xu, et al., Robust super-
hydrophobic surfaces with mechanical durability and easy repairability, J.
Financial support from the Natural Sciences and Engineering Mater. Chem. 21 (2011) 15793, http://dx.doi.org/10.1039/c1jm12513c.
[19] H. Wang, H. Zhou, A. Gestos, J. Fang, T. Lin, Robust, superamphiphobic fabric
Research Council of Canada (NSERC) and the Ontario Research with multiple self-healing ability against both physical and chemical damages,
Fund – Research Excellence (ORF-RE) and access to the electron ACS Appl. Mater. Interfaces 5 (2013) 10221–10226, http://dx.doi.org/10.1021/
microscopy facility in the Ontario Centre for the Characterization am4029679.
[20] Y. Xiu, Y. Liu, D.W. Hess, C.P. Wong, Mechanically robust superhydrophobicity
of Advanced Materials (OCCAM) are acknowledged.
on hierarchically structured Si surfaces, Nanotechnology 21 (2010) 155705,
http://dx.doi.org/10.1088/0957-4484/21/15/155705.
[21] X. Deng, L. Mammen, Y. Zhao, P. Lellig, K. Müllen, C. Li, et al., Transparent,
thermally stable and mechanically robust superhydrophobic surfaces made
References
from porous silica capsules, Adv. Mater. 23 (2011) 2962–2965, http://dx.doi.
org/10.1002/adma.201100410.
[1] B. Bhushan, Y.C. Jung, Micro- and nanoscale characterization of hydrophobic [22] A. Davis, Y.H. Yeong, A. Steele, E. Loth, I.S. Bayer, Nanocomposite coating su-
and hydrophilic leaf surfaces, Nanotechnology 17 (2006) 2758–2772, http://dx. perhydrophobicity recovery after prolonged high-impact simulated rain, RSC
doi.org/10.1088/0957-4484/17/11/008. Adv. 4 (2014) 47222–47226, http://dx.doi.org/10.1039/C4RA08622H.
[2] Y.C. Jung, B. Bhushan, Mechanically durable carbon nanotube - Composite [23] ASTM, ASTM D3359-09e2 Standard Test Methods for Measuring Adhesion by
hierarchical structures with superhydrophobicity, self-cleaning, and low-drag, Tape Test, 2009. http://dx.doi.org/10.1520/D3359-09E02.2.
ACS Nano 3 (2009) 4155–4163, http://dx.doi.org/10.1021/nn901509r. [24] R. Rust-Oleum, Multi Component Product Information Sheet 274232, Vemon
[3] A.J. Scardino, H. Zhang, D.J. Cookson, R.N. Lamb, R. de Nys, The role of nano- Hills, 2015.

Вам также может понравиться