You are on page 1of 29

Accepted Manuscript

Title: A Study of the Mechanical and Chemical Durability of


Ultra-Ever Dry Superhydrophobic Coating on Low Carbon
Steel Surface

Author: Li Wang Jieyi Yang Yan Zhu Zhenhua Li Tao Sheng


Y.M. Hu De-Quan Yang

PII: S0927-7757(16)30094-2
DOI: http://dx.doi.org/doi:10.1016/j.colsurfa.2016.02.022
Reference: COLSUA 20461

To appear in: Colloids and Surfaces A: Physicochem. Eng. Aspects

Received date: 14-11-2015


Revised date: 15-2-2016
Accepted date: 17-2-2016

Please cite this article as: Li Wang, Jieyi Yang, Yan Zhu, Zhenhua Li, Tao
Sheng, Y.M.Hu, De-Quan Yang, A Study of the Mechanical and Chemical
Durability of Ultra-Ever Dry Superhydrophobic Coating on Low Carbon Steel
Surface, Colloids and Surfaces A: Physicochemical and Engineering Aspects
http://dx.doi.org/10.1016/j.colsurfa.2016.02.022

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.
A Study of the Mechanical and Chemical Durability of Ultra-Ever Dry
Superhydrophobic Coating on Low Carbon Steel Surface

Li Wang1,2, Jieyi Yang2, , Yan Zhu1,* Zhenhua Li1,* Tao Sheng1, Y.M. Hu3, De-Quan Yang1,2 *

1
Faculty of Materials Science and Engineering, Kunming University of Science and Technology, 253 Xuefu Rd,
Kunming, Yunnan, 650093, China

2MaterialsResearch Lab., Wuxi Shunye Technology Co. Ltd., 29 Lianze Rd, Shanshui Cheng Tech Park, Suite
15, Binhu District, Wuxi, Jiangsu 214125, China

3School of Engineering, Dali University, 2 Hongsheng Rd. Dali, Yunnan 671003, China

Current address: Department of Biochemistry, McGill University, Montreal, Quebec, Canada; Rosalind and
Morris Goodman Cancer Research Centre, 1160 Pine Ave. West, Montreal, Quebec, Canada H3A 1A3

*Corresponding authors: dequan.yang@gmail.com (De-Quan Yang), Tel:.+86-510-66685886,

Tel:.Fax. +86-510-66685882.

zhuyankmust@foxmail.com (Yan Zhu); 88094709@qq.com (Zhenhua Li)


Graphical abstract
Highlights

1. Mechanical and chemical durability of Ultra-Ever Dry Superhydrophobic coating were


systematically evaluated.
2. XPS and SEM have been conducted to analyze the cause of the change in wettability of the
SH surface under mechanical and chemical test environment.
3. The corrosion process of salt spray induces pit-etching, which can result in the loss of
superhydrophobicity duo to the appearance of Fe-O nanoplates with high surface energy.
4. A new simple method has been proposed to repair the superhydrophobicity which has been
degenerated due to rusting caused by salt spray.

Abstract
Both mechanical and chemical durability of superhydrophobic (SH) surfaces are very important properties for

industrial applications. In this paper, the durability of the Ultra-Ever Dry, as a commercial SH product, sprayed

onto low carbon steel has been systematically studied by sand paper abrasion, waterfall/jet test and immersion

in solution of different pH values as well as salt spray test. The results show that the degeneration of

superhydrophobicity of the coating during the abrasion was mainly due to the loss of both the top layer of micro-

scale bumps and their nanoparticles (NPs). Waterfall/jet impact could also cause the loss of NPs on the micro-

scale bumps of the coating, which corresponds with the change of wettability. Different chemical processes elicit

complex effects on the SH Surface. The coating is able to maintain its superhydrophobicity in solutions of pH 1

to 12, whereas a solution of pH 14 causes both chemical change and loss of NPs, and as a result, the loss of the

water-repellent property. The salt spray test shows that the pitting corrosion on the surface and the degeneration

of superhydrophobicity are mainly induced by morphological and surface chemical changes such as the

formation of new Fe-O nanostructures that presented as pit-etching on the coating surface. The results also
confirm that surperhydrophobictity degeneration of the SH coating surface by salt spray can be easily recovered

through treatment with fluoroalkylsilane (FAS-17).

Keywords: Superhydrophobic surface, Ultra-Ever Dry coating, durability, salt spray

1. Introduction
Superhydrophobic surface, with a water contact angle greater than 150 degree and a sliding angle less than 10

degree, has attracted tremendous attention due to its potential applications in self-cleaning[1-3], anti-icing[4, 5],

anti-fouling[6, 7] and anti-corrosion[8-10]. Presently, countless methods of SH surface have been reported in the

literatures. However, facing with the challenge of practical application, an overwhelming majority of SH surfaces

are still stuck in laboratory due to the complex preparation process and their poor durability (i.e., their

superhydrophobicity cannot persist for a long time and/or in particular environmental conditions). These issues

may be attributed to the fact that the micro/nanostructures, an essential part of SH surface, are inherently fragile.

Despite the difficulties mentioned above, a few companies recently managed to launch their SH products into

the market. These products include Ultra-Ever Dry (Ultratech International. USA)[11], Supercoat (Wuxi Shunye

Technology, China)[12] and Never-Wet (NeverWet LLC, USA)[13]. These SH coatings may (or should) have a

better durability as they are commercial products for practical applications. Hence, a systematic study of their

mechanical and chemical durability should provide valuable information that may be of significance for people

to perform and compare research on SH surfaces in the future.

Recently, many researchers have paid much attention to enhancing the durability of SH surfaces [14-28];

however, the results cannot be easily compared due to the different evaluation methods or the same evaluation

method conducted with various procedures. In addition, there is still no standard sample available for a proper

comparison. A valid and recognized standard for evaluating the durability of SH surface has yet to be established

at present. These issues were comprehensively reviewed and discussed by Milionis Athanasios et al.,[14] and

Malavasi et al.,[15] in their recent papers.

In general, mechanical durability of a SH surface is widely measured by means of abrasion wear [16-27] and
water jet/fall impact[28-34]. The chemical durability of a SH surface is extensively evaluated by immersion of

varying pH solutions[35, 36] and electrochemical corrosion test[8, 37, 38]. Results demonstrate that most of the

SH surfaces show better corrosion resistance based on immersion in different pH solutions and electrochemical

test; however, it may not be enough for practical applications. A SH surface repels liquid by air underneath the

liquid, but not necessarily moisture. Thus for the surface of SH coatings, we need to consider the effects of

moisture. The salt spray test which provides a controllable moisture environment has been intensively used as

corrosion evolution of coatings in industry[39]. Through the salt spray test, one can determine the relative

corrosion-resistance information of various materials including coatings. However, there has been very few

reports on SH coatings that underwent the salt spray test despite the fact that superhydrophobic coatings were

described as an excellent corrosion protection.

In this work, we have systematically studied the durability of Ultra-Ever Dry as a commercially available SH

coating. With our results, we aim to provide an important and independent evaluation of the performance of this

commercial SH coating protocoled as a standard sample for comparison and reference data for future SH surface

evaluations. The mechanical durability of the Ultra-Ever Dry coating was first evaluated by abrasion wear test

with sandpaper and water fall/jet test, their degeneration mechanisms have been studied based on the change of

the surface chemical composition and morphology. The chemical durability of the coating was characterized by

immersion in solutions with different pH and salt spray test. The change of surface morphology was detected

and analyzed from SEM. Surface chemistry has been comprehensively studied using XPS. Lastly, we found that

the loss of superhydrophobicity of the SH surface induced by salt spray can be easily repaired due to the new

nano/micro structures formed during the salt spray.

2. Materials and methods


2.1 Materials

Ultra-Ever Dry was purchased from Ultra Tech International, Inc. with 1LB bottom packages both 4000(bottom

coating) and 4001(top coating). The substrate, a low carbon steel (120 mm length x 50 mm width x 0.28 cm

thickness), was purchased from Wuxi Guangyuan Auspicious Metal Materials Co., Ltd., and its nominal

compositions are listed in Table 1. The surface was successively polished using sandpaper of different grades
(e.g., 200#, 600# and 1200#), cleaned by sonication in ethanol for 5 minutes and quickly dried by an air blower.

For fear of potential corrosion, the edges and the other side of the substrate were protected by tape. According

to the instructions, the coatings were sprayed onto the substrate as shown in Fig. S1 of the supporting information.

Finally, samples were kept at room temperature for 24 hours prior to testing. The thickness of coating was

measured by coating thickness gauge GTS810F (Guangzhou Guo ou electronic Co., Ltd.), the average thickness

of the prayed coating is 45 um.

2.2 Characterizations and tests

2.2.1 Surface morphology and chemical composition

Surface morphology of samples was characterized by a Hitachi S-4800 scanning electron microscopy with field

emission electron gun. The X-ray photoelectron spectroscopy (XPS, PHI5000 Versa scanning probe-II XPS

Microprobe system, with Al k monochromatic X-ray source) was used to study the surface chemical

composition. The survey spectra have been carried out with 100 eV pass energy and 1 eV step. High spectra was

acquired at 20eV pass energy and 0.05eV step. All peaks were calibrated using C1s=284.6 eV. High resolution

XPS spectra were deconvoluted by CASA XPS.

2.2.2 Wettability of surface

Contact angles were measured using a SL200B Static and Dynamic Optical Contact Angle Goniometer (USA

Kino Industry Co., Ltd) with an accuracy of 1. The sessile drop method was used to determine the static contact

angles (CA) by employing 5 L water drops. Droplet shape photo analysis soft-ware was used to measure the

contact angles by the circle fitting method. The sliding angle (SA) was acquired by tilting the sample and

employing 8 L water drops. Average CA and SA values were determined by measuring three or more locations.

2.2.3 Mechanical durability test

Generally, the wear resistance is a very important property and often conducted by a linear shear abrasion on the

tested surface. The sandpaper abrasion test which functions in such a way has been extensively used to evaluate

mechanical robustness or durability of a SH surface [29, 30]. To study the wear resistance of the coating, the
obtained superhydrophobic surface loaded with pressure of 2.8 kPa was dragged forward on a 1500# grad SiC

sandpaper in one direction at a rate of 6cm/s for an abrasion length of 30 cm per cycle. After each cycle, the

sample was dried with a blower and its wettability was measured. The waterfall/jet test[40, 41] is designed to

investigate the mechanical durability of the SH surfaces in long-term exposure to water with different kinetic

energy by a setup which could provide a waterfall/jet flow as shown in Fig. S2 of the supporting information.

The sample was fixed on a substrate tilted at 45 and placed 35 cm below a water pipe (inner diameter 4 mm),

and then high-flux drops with impact velocity of 14 m/s were jetting within 120 min at pressure of 100 kPa.

During the test, the samples were continuously jetting for 30 min per cycle, rinsed with pure water and dried by

an air blower followed by CA and SA measurements.

2.2.4 Electrochemical test

As one of the electrochemical test techniques, potentiodynamic polarization offers information about corrosion

current and corrosion potential, which could estimate the corrosion resistance of the tested materials. Here,

electrochemical test was performed on an electrochemistry workstation (CS350, Wuhan Corrosion test

Instrument Co., Ltd.) by potentiodynamic polarization in a three-electrode system: a working electrode (WE), a

platinum stick counter electrode and a saturated calomel reference electrode (SCE). NaCl solution (3.5 wt %)

was used as the electrolyte. Samples were exposed to NaCl solution for 1h, then dynamic measurement of

polarization curves in a Tafel model was obtained at a scan rate of 1mV/s at room temperature. The potential

range for tests was fixed from -500 mV as the open circuit potential (OCP) value in the cathodic regime to 500

mV as OCP value in the anodic regime.

2.2.5 Immersion of neutral salt solution

The immersion of neutral salt solution (simulating seawater 3.5wt% NaCl, pH=7) at room temperature was

designed to assess the resistance of superhydrophobic surface in ionic environment. During the test, samples

immersed in solution for up to 30 days were randomly chosen in every 24 hours. They were then rinsed with

pure water and dried by an air blower. Finally, the wettability of these samples surface was measured.
2.2.6 Immersion of solution with different pH

To study the effect of solution with different pH on the wettability of SH coating, hydrochloric acid diluted with

deionized water was utilized to create an acidic environment (pH 1 and 4) while sodium hydroxide was used to

create an alkaline environment (pH 11 and 14). Samples were immersed in solutions with different pH for a

week. During the test, they were randomly taken out at different period of 1h, 2h, 4h, 6h, 8h, 12h, 24h, 48h and

72h, etc., rinsed with pure water and dried by an air blower prior to the measurement of their wettability and the

analysis of their durability.

2.2.7 Salt spray test

The neutral salt spray test was based on the ISO-9227:1990 standard. The experiment was carried out in a salt

spray chamber (Wuxi Suoyate Instrument Ltd.) at 35C under 100Kpa of pressure, and the salt spray solution

was 5% NaCl. Samples were checked in every 48 hours and images were recorded to compare the performance

of samples throughout the test period. At every time point, before the measurements of wettability were taken,

samples were successively rinsed with pure water and dried by a blower.

3. Results and discussion


3.1 Surface morphology

The surface morphology of the coating is shown in Fig. 1. The SEM image shows that the coating surface is a

typical asperity surface with micro/nano-structures. The diameter of most micrometer bumps is about 20-30m

with few larger micrometer bumps of about 100m (Fig. 1a). Both surfaces on the bumps and pits are composited

by 20nm nanoparticles (NPs) with nanopores (Fig. 1b, c). This is a typical structure for superhydrophobic

surfaces.

3.2 Mechanical properties and durability

The basic mechanical properties listed in Table 2 are measured by related industrial paints and coatings standards,

and the result indicates that this SH coating exhibits good mechanical property with the exception of hardness.

The wettability change of the surface as a function of abrasion wear test cycle is shown in Fig. 2. We noticed
that the surface can maintain its superhydrophobicity within 8 cycles despite the gradual increasing in SA;

however, the continuous abrasion brings about a rapid deterioration of the average value of SA with larger errors

while the average value of CA undergoes a small change. The variation of surface morphology after 12 cycles

of abrasion test was investigated (shown in Fig. 1d,e,f). It seems that the continuous abrasion has severely

scratched the top section of the micro bumps (Fig. 1d) on which the original nanoparticle layer seems to be

thoroughly abrased away (Fig. 1e,f). Therefore, these transformations of micro/nano-structure of surface induced

by abrasion could be the main reason to the loss of superhydrophobicity. Fig. 3 shows the waterfall/jet test result

of the coating. The change of CA, decreasing from 155 degree to 142 degree along the water flow impact, is not

as significant as the SA whose value increases from 2 degree to 25 degree within 90 min and rapidly degenerates

to a non-roll off over 90 min under 100kPa pressure water jet impacting. It should be noted that the top layer has

been washed away during water flow impact as shown in the optical photograph (Fig. S3 of the supporting

information), and this is also confirmed by SEM[42].

3.3 Corrosion resistance by neutral salt solution of SH coating

Electrochemical analysis is a common method to assess the corrosion resistance of surfaces. The change of

surface corrosion resistance with and without the SH coating can be observed from the polarization curves as

shown in Fig. 4. The superhydrophobic surface produces positive results on the low carbon steel surface due to

the anodic reaction, whose currents are reduced by more than two orders of magnitude; the corrosion potential

shifts positively for about 0.7 V compared with the uncoated surface. This suggests that the SH coating retards

the dissolution of the low-carbon steel between the interface of steel surface and the salt solution. This has been

further demonstrated by the study of salt immersion, and the effect of immersion (3cm in depth) on the wettability

of SH coatings under neutral salt solution (simulating seawater 3.5wt % NaCl, pH=7) is shown in Fig. S4 of

supporting information. The results indicate that neutral salt has little effect on CA and SA of the SH coating

within 30 days. Both electrochemical test and immersion of neutral salt solution indicate that this commercial

SH coating exhibits high chemical durability and anti-corrosion in seawater. The mechanism of anti-corrosion

is that the abundant air trapped in hierarchical rough structures could effectively prevent the infiltration of

corrosive ions[43].
3.4 The effects of immersion on the wettability of the SH coating in different pH value solutions

Although neutral NaCl solution that is similar to seawater shows little effect on the wettability of the coating as

described above, one needs to consider the effects of solution with different pH on the wettability. Results (see

in Fig. 5) show the different influences of pH value on CA and SA of surface within a week of immersion time.

There seems to be little effect on the CA and SA of the immersed SH coating from pH=1 to 11. However, a

significant effect on the SH coating under the solution of pH=14 was observed in Fig. 5 (the inset plot). In order

to study the cause of the change on wettability induced by pH=14 solution immersion, we performed XPS and

SEM analysis of tested surfaces. SEM microphotographs indicated that the nanoparticle layers both on the bumps

and the pits were eliminated by the alkaline solution as shown Fig. 1(g,h,i). Both XPS Survey (Fig. 6) and high

resolution spectra (Fig. 7 and Fig. 8) demonstrate the change of surface chemistry after immersion in pH=14

solution. As estimated through XPS survey spectra, a quantitative chemical composition is listed in Table 3.

Results in Table 4 also confirmed the chemical change of the surface after the immersion in pH=14 solution,

which indicates that the harsh alkali solution (pH=14) has not only decomposed the surface chemistry (Fig. 6-8

and Table 3-4), but also destroyed the surface nanostructured (SEM in Fig. 1h,i). As shown in Fig. 8 and Table

4, chemical etching of harsh solution of pH=14 induces bond breakage of both C-F2 chain and C-Si chain, loss

of Si-O bonds and hydroxylation, which will increase the surface tension compared to the original surface.

Therefore, based on Fig. 1, 6-8 and Table 3 and 4, one can conclude that the increase of CA and the decrease of

SA are caused by surface chemistry changes and the loss of surface nanostructures duo to the erosion caused by

the strong alkaline solution.

3.5 Salt spray effects on the wettability of the SH coating surface

A representative photograph of the coating surface as a function of the salt spray time is shown in Fig. S5 of the

supporting information. After two days of salt spray time, pit-etchings has appeared as shown in Fig. 9. The

results indicate that the size of the etched pits grows with the increase of salt spray time. At the same time, the

rust areal ratio (Fig. 9) of the pit-etchings increases linearly with the salt spray time for the first 6 days, and then

rust area density become large. This trend suggests that the early period (first 6-day) is the nuclei stage of pit-
etching which transits to the growth stage of the pit-etching later. The average size of the pit-etching increased

with the salt spray time, which suggests the growth of the pit-etching is almost linear in relation with time.

Interestingly, the total area of the etched pits correlates with sliding angle as shown in Fig. 10. This correlation

suggests that the SA is directly related with the contact area between water droplet and high surface energy

(rust)[44]. Please note that twenty rusts were randomly chosen and their diameters were measured using an

optical microscope. The corresponding average diameter value (D) was then measured and calculated, and

therefore, the rust areal ratio can be calculated by an equation: (N *1/4 D2)/S. Here, N is total number of rusts;

D is average diameter value; S is the effective surface area value of a tested sample.

According to the observation of the surface morphology during the salt spray test, the surface could be divided

into two typical parts: non-pit-etching region and pit-etching region, whose wettability has distinct performance

(Fig. 11). It has been noted that there is little effect of salt spray time on the wettability of the surface within 30

days in the non-pit etching region (Fig. 11a). However, a significant effect of salt spray time on the wettability

in pit-etching region (Fig. 11b) could be observed. This evidence reveals that the SA increases with the salt spray

time prior to the first 8 day, and then the water droplet is pinned on the surface in the rust region. Surface

morphology changes were observed by SEM are shown in Fig. 1(j,k,l), which illustrates that there are a lot of

nanoplates appearing as the etched pits of the surface (Fig. 1l), while the only NPs layer in the non-pit-etching

region has little visible changes in its topography in the SEM images(Fig. 1k). XPS analysis was used to evaluate

surface chemistry of the surface 12 days after salt spray for these different regions, non-pit-etching region (A),

mild pit-etching region (B) and heavy pit-etching region (C) namely as shown in Fig. 12. It should be noted that

the XPS analysis area size is about 200um in diameter; the XPS signal from B and C region may be larger than

the size of the etched pits, which explains the appearance of significant Fe2p signal for region C. At the same

time, there is a tiny Fe XPS signal in region B (Fig. 13). A comparison of chemical composition with different

salt spray treatments in different regions is listed in Table 5. The high resolution spectra comparison can be found

in Fig. S6 and Fig. 14. The major changes of surface chemical composition are Fe2p and O1s after 12-day salt

spray test (Fig. 14) in the C region, indicating corrosion of the surface in this region. Fe can be attributed to the

mixture of Fe2O3 and FeOOH based on the appearances of satellites, and the position Fe2p 3/2 [45, 46].

Deconvolution of O1s in this region confirms the presence of both Fe2O3 (529.6eV) and FeOOH (530.5eV)[47]as
shown in Fig. 14b. This indicates that the corrosion of substrate (low carbon steel) surface generates the product

in the SH surface through electrochemical process between the SH surface and low carbon steel surface. This

corrosion can be attributed to the inherent porosity between nanoparticles. During the salt spray, some large

pores of the SH surface may not be able to block H2O molecules during the salt spray test, and once water

molecules penetrate the coating, lots of micro-chemical batteries will be formed on the steel side. Consequently,

low carbon steel as a substrate will be rapidly corroded, the process of oxidation is probably due to the following

corrosion reactions[48]:

H2OH++OH- (1)

Fe2++OH-Fe(OH)2 (2)

Fe(OH)2+O2FeOOH (3)

FeOOHFe2O 3+H2O (4)

Fe(OH)2FeO+H2O (5)

FeO+O2Fe2O3 (6)

Fe(OH)2+Cl-FeOH-+Cl-+OH- (7)

Cl- has been established as an active chemical ion that has a strong penetration ability and accelerates the rate of

corrosion[49]. When Cl- is absorbed by Fe(OH)2, reaction (7) occurs [50]. Once stable Fe(OH)2 is destroyed,

more salt spray oxygen and water get into the coating, which will further enhance the rate of corrosion. This then

results in surface microstructure changing from nanoparticles to nanoplates in the corroded region.

4. Reparation of degenerated SH surface after salt spray test


As previously shown, the wettability can be changed with serious rust on the SH coating, which may cause the

surface to lose its water-repellent property. The apparent rust was Fe-O nanoplates (rust or pit-etching) with high

surface tension comparing to the original C-Fx molecules. An important question arises from this: can we recover

the SH surface after rusting? And the answer is that it is possible. We found that the surface is once again

superhydrophobic after the immersion of the corroded samples in ethanol solution with 2 wt% fluoroalkylsilane

(1H, 1H, 2H, 2H- Perfluorodecyltrimethoxysilane) for 10 minutes followed by the quick desiccation with a

blower. Fig. 15 shows the water droplet interaction photograph before and after reparation. The
superhydrophobicity of the repaired surface has been confirmed by CA and SA measurement data for different

corrosion pattern (with different pit-etching size) as shown in Fig. 16. This evidence suggests that we can easily

repair the loss of superhydrophobicity induced by rust from salt spray treatment regardless of the rust size and

rust area density. The reparation of the superhydrophobic surface can be attributed to the formation of the

nanoplates (Fe-O) as a new nanostructure despite the missing of original nanoparticles and surface chemical

modification of rusted region with low surface energy chemistry. It should be mentioned that the mechanical

durability of the repaired SH surface after salt spray test is dependent on the rust area percentage based on our

initial experimental results. More detailed information will be reported in our next paper.

5. Conclusion
We have extensively studied the durability of Ultra-Ever Dry coating including mechanical and chemical

durability. The study demonstrated that immersion of the extremely alkaline solution (pH=14), salt spray test,

sand-paper wear abrasion and waterfall/jet could lead to the loss of superhydrophobicity of the coating.

Specifically, the loss of the SH surface performance, under the harsh solution, was caused by the decomposition

of low surface composition, sand-paper wear could destroy both micro- and nano-structures, and the waterfall/jet

induced a loss of nanoparticles on the SH surface. Salt spray treatment resulted in pit-etching with the formation

of pit-etching rust whose microstructures were composed of Fe-O nanoplates. We proposed a simple method to

repair the loss of SH surface by using low surface energy to modify the rusted surface. This study also

demonstrated that the SH coating had an excellent water-repellent property and anti-corrosion in both neutral

salt solution and solution of pH<12; however, the corrosion resistance to the salt spray or moisture environment

is still limited for the SH surface with porous structures, which allows etching molecules to pass through the

porous and induces the corrosion of the metal substrate.

Acknowledgements
All experiments are conducted in Materials Research Lab of Wuxi Shunye Technologies. L.W expresses his

appreciation to Shunye Technology for financial supporting. The authors acknowledge financial support from

the Wuxi City Innovation Fund (Project No. CGE01G1209), Wuxi City New Technology for

Industrialization (Project No. 12CGFX0042). The authors also appreciate the support from the Natural Science
Foundation of Yunnan Province of China (Grant No. 2012FB127) and the young talents support program of

Faculty of Materials Science and Engineering, Kunming University of Science and Technology.

References
[1] B. Bhushan, Y.C. Jung, K. Koch, Self-cleaning efficiency of artificial superhydrophobic surfaces, Langmuir, 25 (2009)
3240-3248.
[2] C. Neinhuis, W. Barthlott, Characterization and distribution of water-repellent, self-cleaning plant surfaces, Annals
of Botany, 79 (1997) 667-677.
[3] L. Jiang, Y. Zhao, J. Zhai, A lotus leaf like superhydrophobic surface: a porous microsphere/nanofiber
composite film prepared by electrohydrodynamics, Angewandte Chemie, 116 (2004) 4438-4441.
[4] L. Cao, A.K. Jones, V.K. Sikka, J. Wu, D. Gao, Anti-icing superhydrophobic coatings, Langmuir, 25 (2009) 12444-
12448.
[5] P. Kim, T.-S. Wong, J. Alvarenga, M.J. Kreder, W.E. Adorno-Martinez, J. Aizenberg, Liquid-infused nanostructured
surfaces with extreme anti-ice and anti-frost performance, ACS nano, 6 (2012) 6569-6577.
[6] X.-M. Li, D. Reinhoudt, M. Crego-Calama, What do we need for a superhydrophobic surface? A review on the
recent progress in the preparation of superhydrophobic surfaces, Chemical Society Reviews, 36 (2007) 1350-1368.
[7] L. Zhao, Q. Liu, R. Gao, J. Wang, W. Yang, L. Liu, One-step method for the fabrication of superhydrophobic surface
on magnesium alloy and its corrosion protection, antifouling performance, Corrosion Science, 80 (2014) 177-183.
[8] T. Ishizaki, Y. Masuda, M. Sakamoto, Corrosion resistance and durability of superhydrophobic surface formed on
magnesium alloy coated with nanostructured cerium oxide film and fluoroalkylsilane molecules in corrosive NaCl
aqueous solution, Langmuir, 27 (2011) 4780-4788.
[9] A.V. Rao, S.S. Latthe, S.A. Mahadik, C. Kappenstein, Mechanically stable and corrosion resistant superhydrophobic
solgel coatings on copper substrate, Applied Surface Science, 257 (2011) 5772-5776.
[10] F. Zhang, L. Zhao, H. Chen, S. Xu, D.G. Evans, X. Duan, Corrosion resistance of superhydrophobic layered double
hydroxide films on aluminum, Angewandte Chemie International Edition, 47 (2008) 2466-2469.
[11] Please visit: http://www.ultraeverdrystore.com/.
[12] Z.-J. Yu, J. Yang, F. Wan, Q. Ge, L.-L. Yang, Z.-L. Ding, D.-Q. Yang, E. Sacher, T.T. Isimjan, How to repel hot water
from a superhydrophobic surface?, Journal of Materials Chemistry A, 2 (2014) 10639-10646.
[13] Please visit: http://www.neverwet.com/.
[14] A. Milionis, E. Loth, I.S. Bayer, Recent advances in the mechanical durability of superhydrophobic materials,
Advances in Colloid and Interface Science, (2015).
[15] I. Malavasi, I. Bernagozzi, C. Antonini, M. Marengo, Assessing durability of superhydrophobic surfaces, Surface
Innovations, 3 (2014) 49-60.
[16] L. Boinovich, A. Emelyanenko, The behaviour of fluoro-and hydrocarbon surfactants used for fabrication of
superhydrophobic coatings at solid/water interface, Colloids and Surfaces A: Physicochemical and Engineering
Aspects, 481 (2015) 167-175.
[17] L.B. Boinovich, A.M. Emelyanenko, A.D. Modestov, A.G. Domantovsky, K.A. Emelyanenko, Synergistic effect of
superhydrophobicity and oxidized layers on corrosion resistance of aluminum alloy surface textured by nanosecond
laser treatment, ACS applied materials & interfaces, 7 (2015) 19500-19508.
[18] A.M. Emelyanenko, F.M. Shagieva, A.G. Domantovsky, L.B. Boinovich, Nanosecond laser micro-and
nanotexturing for the design of a superhydrophobic coating robust against long-term contact with water, cavitation,
and abrasion, Applied Surface Science, 332 (2015) 513-517.
[19] I.S. Bayer, A. Brown, A. Steele, E. Loth, Transforming anaerobic adhesives into highly durable and abrasion
resistant superhydrophobic organoclay nanocomposite films: a new hybrid spray adhesive for tough
superhydrophobicity, Applied physics express, 2 (2009) 125003.
[20] J. Cheek, A. Steele, I.S. Bayer, E. Loth, Underwater saturation resistance and electrolytic functionality for
superhydrophobic nanocomposites, Colloid and Polymer Science, 291 (2013).
[21] H. Cho, D. Kim, C. Lee, W. Hwang, A simple fabrication method for mechanically robust superhydrophobic surface
by hierarchical aluminum hydroxide structures, Current Applied Physics, 13 (2013) 762-767.
[22] H. Jin, X. Tian, O. Ikkala, R.H. Ras, Preservation of superhydrophobic and superoleophobic properties upon wear
damage, ACS applied materials & interfaces, 5 (2013) 485-488.
[23] I.A. Larmour, G.C. Saunders, S.E. Bell, Compressed metal powders that remain superhydrophobic after abrasion,
ACS Applied Materials & Interfaces, 2 (2010) 2703-2706.
[24] B. Li, J. Zhang, L. Wu, A. Wang, Durable superhydrophobic surfaces prepared by spray coating of polymerized
organosilane/attapulgite nanocomposites, ChemPlusChem, 78 (2013) 1503-1509.
[25] T.M. Schutzius, M.K. Tiwari, I.S. Bayer, C.M. Megaridis, High strain sustaining, nitrile rubber based, large-area,
superhydrophobic, nanostructured composite coatings, Composites Part A: Applied Science and Manufacturing, 42
(2011) 979-985.
[26] A. Steele, I. Bayer, E. Loth, Adhesion strength and superhydrophobicity of polyurethane/organoclay
nanocomposite coatings, Journal of Applied Polymer Science, 125 (2012) E445-E452.
[27] A. Steele, A. Davis, J. Kim, E. Loth, I.S. Bayer, Wear Independent Similarity, ACS Applied Materials & Interfaces,
(2015).
[28] I.S. Bayer, A.J. Davis, E. Loth, A. Steele, Water jet resistant superhydrophobic carbonaceous films by flame
synthesis and tribocharging, Materials Today Communications, 3 (2015) 57-68.
[29] X. Zhu, Z. Zhang, X. Men, J. Yang, K. Wang, X. Xu, X. Zhou, Q. Xue, Robust superhydrophobic surfaces with
mechanical durability and easy repairability, Journal of Materials Chemistry, 21 (2011) 15793-15797.
[30] T. Verho, C. Bower, P. Andrew, S. Franssila, O. Ikkala, R.H. Ras, Mechanically durable superhydrophobic surfaces,
Advanced Materials, 23 (2011) 673-678.
[31] S. Hoshian, V. Jokinen, V. Somerkivi, A.R. Lokanathan, S. Franssila, Robust Superhydrophobic Silicon without a
Low Surface-Energy Hydrophobic Coating, ACS applied materials & interfaces, 7 (2014) 941-949.
[32] E. Huovinen, L. Takkunen, T. Korpela, M. Suvanto, T.T. Pakkanen, T.A. Pakkanen, Mechanically robust
superhydrophobic polymer surfaces based on protective micropillars, Langmuir, 30 (2014) 1435-1443.
[33] E. Huovinen, J. Hirvi, M. Suvanto, T.A. Pakkanen, Micromicro hierarchy replacing micronano hierarchy: a
precisely controlled way to produce wear-resistant superhydrophobic polymer surfaces, Langmuir, 28 (2012) 14747-
14755.
[34] J. Groten, J.r. Rhe, Surfaces with combined microscale and nanoscale structures: a route to mechanically stable
superhydrophobic surfaces?, Langmuir, 29 (2013) 3765-3772.
[35] S. Wang, L. Feng, L. Jiang, One Step Solution Immersion Process for the Fabrication of Stable Bionic
Superhydrophobic Surfaces, Advanced Materials, 18 (2006) 767-770.
[36] T. Liu, Y. Yin, S. Chen, X. Chang, S. Cheng, Super-hydrophobic surfaces improve corrosion resistance of copper in
seawater, Electrochimica Acta, 52 (2007) 3709-3713.
[37] T. Ishizaki, J. Hieda, N. Saito, N. Saito, O. Takai, Corrosion resistance and chemical stability of super-hydrophobic
film deposited on magnesium alloy AZ31 by microwave plasma-enhanced chemical vapor deposition, Electrochimica
Acta, 55 (2010) 7094-7101.
[38] T. He, Y. Wang, Y. Zhang, T. Xu, T. Liu, Super-hydrophobic surface treatment as corrosion protection for aluminum
in seawater, Corrosion science, 51 (2009) 1757-1761.
[39] J. Liang, Y. Hu, Y. Wu, H. Chen, Facile formation of superhydrophobic silica-based surface on aluminum substrate
with tetraethylorthosilicate and vinyltriethoxysilane as co-precursor and its corrosion resistant performance in
corrosive NaCl aqueous solution, Surface and Coatings Technology, 240 (2014) 145-153.
[40] Y.C. Jung, B. Bhushan, Mechanically durable carbon nanotube composite hierarchical structures with
superhydrophobicity, self-cleaning, and low-drag, ACS nano, 3 (2009) 4155-4163.
[41] A. Davis, Y.H. Yeong, A. Steele, E. Loth, I.S. Bayer, Spray impact resistance of a superhydrophobic nanocomposite
coating, AIChE Journal, 60 (2014) 3025-3032.
[42] W.-H Hu et al., A simple method to improve mechanical robust of the superhydrophobic coating. manuscript
in preparation.
[43] T. Liu, S. Chen, S. Cheng, J. Tian, X. Chang, Y. Yin, Corrosion behavior of super-hydrophobic surface on copper in
seawater, Electrochimica Acta, 52 (2007) 8003-8007.
[44] P. Hao, C. Lv, Z. Yao, F. He, Sliding behavior of water droplet on superhydrophobic surface, EPL (Europhysics
Letters), 90 (2010) 66003.
[45] I. Welsh, P. Sherwood, Photoemission and electronic structure of FeOOH: Distinguishing between oxide and
oxyhydroxide, Physical Review B, 40 (1989) 6386.
[46] T. Yamashita, P. Hayes, Analysis of XPS spectra of Fe 2+ and Fe 3+ ions in oxide materials, Applied Surface Science,
254 (2008) 2441-2449.
[47] S. Poulin, R. Franca, L. Moreau-Blanger, E. Sacher, Confirmation of X-ray photoelectron spectroscopy peak
attributions of nanoparticulate iron oxides, using symmetric peak component line shapes, The Journal of Physical
Chemistry C, 114 (2010) 10711-10718.
[48] Z. Zhi-hong, X. Xiang-xin, R.L.-y. He, D. Pei-ning, Effect of Different Boron Content on Corrosion Resistance of Low
Carbon Steel Iron and Steel, 48 (2013) 56-59.
[49] Q. Zhang, J. Wang, J. Wu, W. Zheng, J. Chen, A. Li, Effect of ion selective property on protective ability of rust
layer formed on weathering steel exposed in the marine atmosphere, Acta metallurgica sinica, 37 (2001) 193-196.
[50] J. Dawson, M. Ferreira, Electrochemical studies of the pitting of austenitic stainless steel, Corrosion Science, 26
(1986) 1009-1026.
Fig. 1 SEM photographs of the SH coating: (a) low magnification, (b) and (c) high magnification
of both bump and concave surfaces. For the surface after 12 cycle abrasion test: (d) low
magnification, (e) and (f) high magnification in the bump surface. For the surface immersed in
solution of pH 14 after 6 hrs: (g), (h) (bumps in g) and (i) (non-bumps region in g). And for the
surface after 12 days salt spray test: (j) low magnification, (k) high magnification in non-pit-
etching region and (l) high magnification in pit-etching region.
Fig. 2 The CA and SA as a function of sandpaper abrasion cycle number under 2.8 kPa pressure.

Fig. 3 The CA and SA as a function of water impact time under given water flow pressure
of 100kPa.
Fig. 4 Polarization curves of both uncoated and SH coated samples.
Fig. 5 CA and SA as a function of pH value in aqueous solution after a week. Insert is CA and SA evolution with
immersion time at pH= 14.

Fig. 6 XPS survey spectra of the coating before and after immersion in pH=14 for 6hrs.
(a) (b) (c)

(d) (e) (f)

Fig. 7 A comparison of high resolution XPS spectra of the surfaces before and after immersion
in pH=14 for 6 hrs: (a) C1s, (b) O1s, (c) F1s, (d) Si2p, (e) Cl2p and (f) Fe2p.
(a) (b)

(c) (d)

XXX

Fig. 8 Deconvolution of high resolution XPS spectra of samples (a) C1s, (c) O1s and immersion
in pH=14 for 6hrs: (b) C1s, (d) O1s.
Fig. 9 Rust areal ratio and average size of pit- Fig. 10 SA as a function of the rust areal
etching evolution with salt spray time. ration in percentage (N*1/4 D2/S).

(a) (b)

Fig.13 CA and SA as function of salt spray time in (a) non-pit-etching region, (b) pit-etching region.

Fig. 11 CA and SA as a function of salt spray time in (a) non-pit-etching region and (b) pit-etching region.
(a) (b)
A

B C

Fig. 12 The representative photographs of rusted coating surface (a) pit-etching distribution and (b)
enlarged region marked letter for XPS analysis.

Fig. 13 XPS survey spectra comparison in different regions for SH sample and sample in non-
pit-etching region (A), mild pit-etching region (B) and heavy pit-etching region (C) after 12-
day salt spray test.
(a) (b)

Fig. 14 XPS high resolution spectra for heavy pit-etching region C: (a) Fe2p and (b) O1s.

Fig. 15 Photographs demonstrated the different wettability of the same surface: the first row
photos are of pre-repair, the second row photos are of post-repair.

Fig. 16 CA and SA evolution of pre- and post-repair of different rust pitting size.
Table 1 Chemical composition of the low carbon steel in this work.

Elements C Mn S P Si Fe
Content(%) 0.13 0.60 0.05 0.015 0.01 residue

Table 2 List of mechanical properties of the SH coating.

Physical Impact resistance


Hardnessa Adhesionb Flexibility (mm)c
parameters (kg.cm)d

value HB 1-2 1 >50


a Evaluated by Paints and varnishes- determination of film hardness by pencil test ( ISO 15184:1998(E))
b
Evaluated by Paints and varnishes - Cross-cut test (ISO 2049:1992(E))
c
Evaluated by Determination of flexibility of films (BG/T 1731-93)
d
Evaluated by Paints and varnishes Rapid-deformation (impact resistance) tests: Part 1: Falling-weight

test, large-area indenter (EN ISO 62721:2004)


Table 3 Surface chemical composition change of before and after immersion in pH=14 for 6 hrs estimated by
XPS (at.%).
Sample C O F Si Cl

Before 21.74 32.09 31.99 14.79 0

After 37.32 25.23 27.03 9.39 1.04

Table 4 Surface chemistry of before and after immersion in pH=14 for 6 hrs estimated by XPS (at.%).
C1s O1s
Sample
C-C(H) C-O COO C-F2 C-F3 C-Si C-F C-O Si-O OH
Before 20.87 14.63 4.10 42.29 7.95 10.15 0 63.17 36.83 0
After 49.35 26.31 5.37 9.38 2.11 2.29 5.20 75.75 24.25 7.78

Table 5 Surface chemical composition of before and after 12-day salt spray test estimated by XPS (at.%).
Sample C O F Si Fe Cl Na
SHS 20.19 28.05 24.74 27.02 0 0 0
A 24.06 27.21 22.81 25.92 0 0 0
B 24.55 28.37 20.52 24.26 0.85 1.43 0
C 27.34 38.82 11.36 9.51 9.74 3.22 0

You might also like