Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Full Paper

Enhancing the Mechanical Properties of


Superhydrophobic Atmospheric Pressure
Plasma Deposited Siloxane Coatings

Charles E. Nwankire, Gregory Favaro, Quynh-Huong Duong,


Denis P. Dowling*

Surfaces with water contact angles above 1508 are regarded as superhydrophobic. In this study
the use of atmospheric pressure plasma jet system called PlasmaStreamTM to deposit super-
hydrophobic coatings is investigated. The coatings were deposited from the following liquid
precursors: hexamethyldisiloxane (HMDSO), tetramethyl cyclotetrasiloxane (Tomcats) and a
mixture of Tomcats and fluorosiloxane. The objective of the study is to investigate how
precursor type and deposition conditions, influences the morphology and mechanical per-
formance of the deposited superhydrophobic coatings. Optical profilometry, AFM, SEM,
Ellipsometry, XPS, Water contact angle and FTIR techniques were used to evaluate the surface
roughness, morphology, thickness and chemical functionality of the deposited coatings. The
mechanical properties were evaluated using
the Nano Tribometer, Nano Scratch, Ultra
Nanoindentation and ultrasonic abrasion
tests. Superhydrophobic coatings deposited
from a precursor mixture of Tomcats and fluor-
osiloxane yielded a substantial enhancement
in coating adhesion and mechanical durability
compared to the superhydrophobic coatings
obtained with either Tomcats or HMDSO pre-
cursors alone.

Introduction properties have been reported to depend on both chemical


composition and surface morphology.[1–3] In a review,
Surfaces with water contact angles greater than 1508 are Genzer et al.[2] showed that superhydrophobic surfaces can
generally regarded as superhydrophobic.[1] There is con- be achieved by tailoring the chemical composition and
siderable interest in these surfaces due to their anti-wetting physical appearance of surfaces. Lee et al.[1] achieved SH
and self-cleaning properties. The superhydrophobic (SH) surface by grafting polyacrylic acid chains onto nylon
surfaces with high surface roughness. Oner et al.[3] prepared
ultrahydrophobic surfaces by photolithography and hydro-
C. E. Nwankire, D. P. Dowling
School of Electrical Electronic and Mechanical Engineering,
phobised using silane reagents. The surfaces were termed
University College Dublin, Ireland ultrahydrophobic because they exhibited advancing and
E-mail: denis.dowling@ucd.ie receding water contact angles above 1708. Some of the other
G. Favaro, Q.-H. Duong techniques that have been used to fabricate superhydro-
CSM Instruments SA, Peseux, Switzerland phobic surfaces include wax solidification,[4] lithography,[3]

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim wileyonlinelibrary.com DOI: 10.1002/ppap.201000069 305
C. E. Nwankire, G. Favaro, Q.-H. Duong, D. P. Dowling

carbon nano tube alignment by plasma enhanced chemical fluorosiloxane and Tomcats (TCFS). The flow rates investigated
vapour deposition,[5] polymer reconformation,[6] sol gel,[7] for these liquid precursors were in the range 3 to 8 mlmin1. Helium
and electrochemical processing.[8] Atmospheric pressure (He) gas flow rate was kept constant at 5 slm. N2 gas (flow rate of
100 mlmin1) was added to the helium plasma to help reduce the
plasmas have also been used and their advantage is firstly
formation of particulates.[17] In this study, for all the coatings
the relative speed of superhydrophobic coating deposition,
investigated the plasma power was maintained at 7.6 W and the
and generally only a single step deposition process is
coatings were deposited by passing the jet 3 times across the silicon
required.[9–12] Amongst the precursors used to deposit wafer substrates at a CNC speed 10 mms1.
these atmospheric plasma deposited coatings have been Coating contact angle measurements were carried out six days
fluorocarbon gases,[12] hexamethyldisilazane[10] and hexa- after deposition in order to avoid any effects due to hydrophobic
methyldisiloxane (HMDSO).[9,13] In a previous study recovery.[16] The coatings were stored in Petri dishes under normal
involving the PlasmaStreamTM atmospheric plasma jet room temperature conditions. Coating contact angles were
deposition system at University College Dublin (UCD), a SH measured at room temperature using a video capture apparatus
coating was deposited from HMDSO.[9] The superhydro- OCA 20 from Dataphysics Instruments. Deionised water, diiodo-
phobicity of the HMDSO coating depended on a relatively methane and ethylene glycol were used as test liquids. Contact
angles were measured at three different locations on the coating
fragile needle-like morphology. It has previously been
surface and averaged. The OWRK (Owens, Wendt, Rabel and Kaelbe)
reported that the mechanical durability of SH coatings
method was used to determine the surface energy of the deposited
deposited by plasma fluorination of polybutadiene films coatings from the contact angle measurements.[18]
was enhanced 8-fold after thermal treatment at 155 8C for Fourier transform infrared spectroscopy (FTIR) measurements
1 h.[14] The objective of this study is to evaluate if the were carried out using a Bruker Vertex – 70 system, and the
mechanical durability of the superhydrophobic coatings transmission spectra were collected in the range of 400–4 000 cm1
can be enhanced by altering the siloxane precursor using a spectral resolution of 4 cm1 after 64 scans.
chemistry used in the deposition of the SH coatings. X-ray photoelectron spectroscopy (XPS) was performed using a
In addition to HMDSO, the liquid siloxane precursors Kratos Analytical Axis Ultra photoelectron spectrometer. The
investigated were perfluorooctyl triethoxysilane (fluorosi- instrument has a spherical mirror analyzer with an integrated
loxane), tetramethyl cyclotetrasiloxane (Tomcats) and a automatic charge neutralizer and a magnetic lens. A monochro-
mated (Al Ka) X-ray source was used to record spectra at normal
mixture of Tomcats and fluorosiloxane. The ability of any of
emission. The area analysed was 700  300 mm2 with a take-off-
these precursors to form superhydrophobic coatings has
angle of 908. The elemental composition is presented as the average
not been reported previously. of three analyses. Each position was analysed in the survey mode
(Pass Energy 160 eV) to determine the elements present at the
surface and their relative concentrations. CasaXPS (Casa Software
Experimental Part Ltd) data processing software was used to calculate the area under
Equipment and Materials peaks representative of elements detected, which were then
normalized to take into account relative sensitivity to provide
The siloxane coatings were deposited onto one-side polished P-type relative concentrations.
silicon wafers (450 mm thick). These substrates were ultrasonically Surface roughness and morphology analyses of the coating was
cleaned in acetone and methanol for 10 min, respectively, and were examined using a Wyko NT1100 optical profilometer scanning at
then dried in air. Coatings were deposited using the Dow Corning least 3 different areas of 45  59 mm2 in vertical scanning
atmospheric pressure plasma jet system called PlasmaStreamTM, interferometer (VSI) mode, unless stated otherwise in the text.
which has been reported in details elsewhere.[15,16] The system was The reported average (Ra) and root mean square (Rq) surface
configured with a dielectric head housing two pin electrodes either roughness is the mean of three measurements on different areas of
side of a pneumatic nebuliser (Burgener Ari Mist nebuliser), each sample taken in order to verify surface homogeneity. Surface
through which an aerosol of the precursor was introduced. The morphology examination was also carried out using a CP-II (Veeco)
resultant helium – aerosol mix exited the system through a 75 mm Atomic force microscope (AFM) scanner in non contact mode and a
long  16 mm diameter dielectric tube. Low frequency electrical TM-1000 Hitachi TableTOP scanning electron microscope (SEM).
power was delivered to both electrodes from a modified PTI 100W rf The coating thickness and optical properties were measured with
power supply at a frequency of approximately 5  25 kHz. The the J.A. Woollam M2000 variable angle ellipsometer. Coating
entire plasma device was moved over the surface of the substrate in thickness was obtained by using multiple angle measurements
a raster pattern (xy directional scan) using a CNC device with a line (658, 708 and 758) over a wavelength of 250 to 1 690 nm. The
speed of 10 mms1 and a step interval of 2.5 mm over the area of refractive index was measured at the wavelength of 632.8 nm and
100  100 mm2. The substrate to source distance (gap distance) was fitted using Cauchy model.[19]
varied from 2 to 8 mm. The mechanical properties of the coatings were evaluated with
Coatings were deposited using the following precursors the CSM Nano Scratch Tester, Nano Tribometer and Ultra
hexamethyldisiloxane (HMDSO) ((CH3)3SiOSi(CH3)3, Aldrich Nanoindentation Testers (CSM Instruments, Switzerland).[20]
98.5%), tetramethyl cyclotetrasiloxane (Tomcats) ((HSiCH3O)4, The Nano Scratch Test was carried out using a 908 sphero-conical
Aldrich 99%), perfluorooctyl triethoxysilane (fluorosiloxane or FS) stylus with an indenter radius of 10 mm. Three scratches were
(C14H19F13O3Si, Aldrich 98%) and a 50/50 wt.-% mixture of obtained from each test surface, with an initial scanning load of

Plasma Process. Polym. 2011, 8, 305–315


306 ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/ppap.201000069
Enhancing the Mechanical Properties of Superhydrophobic . . .

0.3 mN and final loads of 40 mN (for Tomcats and HMDSO coatings) given time period and tested, it was re-inserted in the test
and 80 mN (for TCFS coating). The critical loads were determined apparatus for the longer term assessment of abrasion resistance.
from the recorded normal force vs. penetration depth curves along The results of this ‘in-house’ abrasion test were correlated with
the scratch; the respective images have also been taken. other mechanical test results obtained using Nano scratch,
The Nano Tribometer system was used to determine the friction Nano Tribometer and Ultra Nanoindentation equipment (CSM
coefficient of the coatings. Loads in the range of 0.05 mN to 1 N were Instruments).
used. A 4 mm diameter 100 Cr6 steel ball (mounted on a stiff lever)
was loaded onto the coated samples with a normal force of 0.5 mN.
This was moved against the coated samples at a linear speed of
1 cms1, and was stopped after 500 laps. The friction coefficient Results and Discussion
was determined by measuring the deflection force on the
cantilevers. HMDSO Superhydrophobic Coatings
Mechanical properties of the coatings such as Hardness and
The HMDSO coatings were deposited at a precursor flow
Elastic Modulus were obtained using an Ultra Nanoindentation
[21] rate of 3 mlmin1 and a gap distance of 8 mm. These
Tester equipped with a Berkovich diamond indenter tip. The
maximum loading/unloading rate was 40 and 200 mNmin1, parameters were previously established as producing
while the maximum load was 20 and 100 mN for the coated and superhydrophobic coatings using the PlasmaStreamTM
uncoated samples respectively. The system was used to measure system.[9] The optical profilometry image of a typical
hardness, H (resistance to local deformation) and elastic modulus, E HMDSO coating deposited using these deposition condi-
of the coatings. The latter was estimated by analysing the load- tions is given in Figure 1. This image demonstrates that the
displacement curves based on the Oliver – Pharr method.[22] The coating exhibits a nano-textured ‘needle-like’ morphology.
resistance of the coatings to plastic deformation (H3/E2) and E to H The water droplet with contact angle measured as 1568 is
ratio were obtained.[23] also shown in this figure. Coating surface energy is 5
Abrasion resistance is an important parameter in the evaluation
mNm1, while the Ra and Rq are 20 and 26 nm,
of the wear resistance of the coatings. Xiu et al.[24] evaluated the
respectively. The coating thickness obtained with ellipso-
abrasion resistance of superhydrophobic surfaces formed on silicon
metry was approximately 350 nm. The needle-like coating
wafer by potassium hydroxide and metal assisted etching, using a
‘Technicloth’ wipe as the abrasion surface. It was reported that the morphology is most likely due to precursor fragmentation
superhydrophobic surfaces lost their self-cleaning ability after the and gaseous phase reactions in the plasma.[17]
abrasion tests. To our knowledge, an abrasion resistance test has
not been reported previously for use with nm thick coatings. A
comparative laboratory test is therefore proposed as part of this Tomcats Coatings
study. The test involves immersion of the silicon wafer coated
substrates in a 300 ml beaker containing a mixture of 50 wt.-% Superhydrophobic Tomcats coatings were obtained using
deionised water with 20 mm diameter precision micro-abrasive the conditions detailed in Table 1.
silicon carbide (SiC) particles (Comco Inc, USA). The volume of the The morphology of a superhydrophobic Tomcats coating
aqueous mixture was fixed at 200 cm3 and into this mixture a obtained by AFM is shown in Figure 2. The coating exhibits
UP200H ultrasonic processor probe (Hielscher Ultrasonics GmbH, both micro and nano rough morphology, which has been
Germany) was lowered to a distance of
approximately 5 mm above the base of the
beaker. The height of the ultrasonic probe was
maintained at this distance in order to stop the
abrasive particles from settling at the bottom
of the beaker, thus maintaining uniform
sample abrasion. This 200 W, 24 kHz frequency
ultrasonic probe operated at a 100% pulse cycle
rate and 50% amplitude was housed in a
SB3-16 sound protection box. After abrasion
with the SiC particles for periods of between 1
and 150 min, the test samples were removed,
washed with deionised water, followed by air
drying. Any remaining abrasive particles on
the tested surface were removed using com-
pressed air. The adhesion and abrasion resis-
tance of the tested coatings were evaluated
using FTIR and water contact angle measure-
ments. Each test was carried out in triplicate Figure 1. Optical profilometry image of a superhydrophobic HMDSO coating showing the
and once the coating had been removed after a needle-like morphology and water contact angle on the coated surface.

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.plasma-polymers.org 307
C. E. Nwankire, G. Favaro, Q.-H. Duong, D. P. Dowling

Table 1. Plasma deposition parameters for superhydrophobic Table 2. Effect of gap distance on Tomcats coating properties.
Tomcats coatings. WCA is the water contact angle.

Plasma No. of Gap CNC Gap Ra Thickness WCA Surface Si  CH3 /


power Passes distance speed distance energy Si  O  Si

W mm mms1 mm nm nm - mNm1

7.6 3 2 10 2 19 353  25 155 5 0.095


8 152 410  41 90 24 0.134

table, it is shown that as gap distance is increased, the water


contact angle for the deposited Tomcats coatings decreased
from 1558 to 908. This is likely to be due to an enhanced level
of coating oxidation as the distance from the substrate to
the jet orifice is increased. This conclusion is supported by
the work previously carried out by O’Neill et al.[25] They
reported that enhanced coating oxidation at the higher gap
distance could be attributed to the fluid dynamics of the
atmospheric plasma jet system; as the flow of gas exiting
the PlasmaStreamTM can entrain air which would subse-
quently oxidize the precursor as it travels towards the
substrate.
Figure 2. AFM image of Tomcats coating Ra 12 nm demonstrating Surface roughness measurements demonstrated that, as
the micro and nano rough morphology. the gap distance increased from 2 to 8 mm, the coatings
showed an increase in surface roughness from 19 to 152 nm.
The corresponding optical profilometry images are given in
reported previously to be important for superhydrophobi- Figure 3. It was also observed that the coating thickness
city.[1–3] The Ra value of 12 nm is the mean of three different increased from approximately 353 to 410 nm with this
scans. increase in gap distance. The latter thickness value is
The optimised gap distance between the jet orifice and however only an approximation due to the very high
the substrate as detailed in Table 1 was found to be 2 mm. In surface roughness.
order to investigate the influence of this gap distance on The FTIR spectra of the Tomcats precursor monomer and
coating properties, coatings were also deposited at 8 mm plasma polymerised coatings deposited at the gap dis-
gap distances, while keeping the other deposition para- tances of 2 and 8 mm, under conditions of 7.6 W plasma
meters constant. The influence on coating properties of this power, 10 mms1 CNC speed and 3 passes of the jet, are
change in the gap distance is given in Table 2. From this given in Figure 4. These spectra confirm retention of the

Figure 3. Optical profilometry images showing the morphology of Tomcats coatings at the gap distance of 2 mm (left) Ra 19 nm and 8 mm
(right) Ra 152 nm, respectively. Note the decrease in water contact angle for the latter coating.

Plasma Process. Polym. 2011, 8, 305–315


308 ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/ppap.201000069
Enhancing the Mechanical Properties of Superhydrophobic . . .

respectively. The corresponding absorption peaks for the


Tomcats coating were observed at approximately 2 182 and
2 242 cm1. Further FTIR analysis was carried out by
integrating[16] the peak ratio of the SiCH3 to SiOSi
functional groups at 1 255 and 850 cm1 respectively. The
results which are given in Table 2 demonstrate that the
SiCH3 to SiOSi ratio increased from 0.095 to 0.134 for
the coatings deposited at 2 and 8 mm gap distance
respectively. This 29% increase in the SiCH3 to SiOSi
ratio of the deposited coatings with increased gap distance,
is an indication of higher organic presence in the coating
deposited at the lower gap distance, thus supporting the
conclusion that the coatings deposited at the 2 mm gap
distance are less oxidised.
The refractive index of the Tomcats coatings was
obtained using ellipsometry and it was observed that the
index of the coatings increased from 1.14 to 1.27 as the gap
distance increased from 2 to 8 mm. An increase in siloxane
coating refractive index has previously been associated
with a decrease in the organic fraction of the coating, again
suggesting a higher level of coating oxidation.[19] The lower
refractive index of the coating deposited at the 2 mm gap
distance may thus be due to a higher organic composition,
which would contribute to its superhydrophobicity.

Figure 4. FTIR spectra of the Tomcats monomer and coatings


deposited at 2 and 8 mm gap distances. Note the retention of the Superoleophilicity
functional chemistry in the plasma polymerised coatings.
The superhydrophobic Tomcats coating with a water contact
angle of 1558, was tested for oil repellency by measuring the
monomer chemistry in the plasma polymerised coatings, as contact angle of decane[29] (surface tension g 23.8 mNm1).
previously reported for other plasma deposited coatings.[16] Decane completely wetted the surface with contact
The peak at 1 255 cm1 is assigned to the SiCH3 functional angles < 58. The surface energy of the Tomcats coated
group (CH3 symmetric deformation). The SiC stretching is sample is 5 mNm1. A possible explanation for the wetting
observed at 760 cm1 and the peaks at 2 170 and 890 cm1 by decane of the superhydrophobic Tomcats coatings may be
are attributed to the absorption and bending frequencies of the oleophilic character associated with the methyl groups
SiH group. The asymmetric stretching of the SiOSi within this polymer (see Figure 4). One potential application
group is absorbed at 1 090 cm1. The absorption band at of surfaces that possess both superhydrophobic (water
2 970 cm1 is due to CH stretching in CH3 functional contact angle >1508) and superoleophilic (oil contact
group.[26] The broad peak around 3 300 cm1 which is angle < 58) properties is their use as oil sorbents. These have
assigned to OH group is almost absent in the case of the previously been reported for the separation of diesel from
2 mm gap distance but is relatively intense in the spectrum water.[30] The HMDSO coating also exhibited superoleo-
of the coating deposited at the 8 mm gap distance.[27] This philicity as decane completely wetted the surface.
confirms the higher level of oxidation of the latter coating.
A comparison of the SiOSi peaks in the FTIR spectra of
Superhydrophobic Tomcats/Fluorosiloxane Coatings
the precursor monomer and that of the coatings deposited
at the 2 and 8 mm gap distances, demonstrated a broad- Deposition studies were carried out with fluorosiloxane
ening of the SiOSi peak for the coating deposited at the precursor at flow rate of 5 mlmin1 and 2 mm gap distance.
larger gap distance (see Figure 4). This shift has previously The resulting coatings however did not exhibit super-
been reported with increased exposure of plasma poly- hydrophobic properties. In this study, the highest water
merised siloxanes to air.[28] This paper also reported that contact angle achieved using a fluorosiloxane only
due to the back-bonding of oxygen with silicon, the position precursor was 1208, although the coatings exhibited low
of the SiH stretching vibration observed at 2 170 cm1 surface energy of 6 mNm1, similar to that of HMDSO
moves to a higher energy positions of 2 180 and 2 240 cm1. superhydrophobic coating. The addition of the fluorosil-
These two peaks are assigned to SiH2O2 and SiHO3, oxane to Tomcats was investigated as a means of enhancing

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.plasma-polymers.org 309
C. E. Nwankire, G. Favaro, Q.-H. Duong, D. P. Dowling

Table 3. Comparison between the Tomcats and TCFS coatings.

Coating Thickness Ra Rq WCA


type
nm nm nm -

Tomcats 370 18 29 155


TCFS 352 13 17 158

water droplet volume. It was attempted to also measure the


contact angle with water volumes of 2 to 4 ml, which have
been reported by other authors previously.[29] At these
volumes the contact angle appeared to be considerably
Figure 5. FTIR spectra of the Tomcats, fluorosiloxane (FS) and the higher but the droplet simply slid off the surface, making it
TCFS coatings.
impossible to obtain a consistent contact angle measure-
ment. A comparison was made between the properties of
the Tomcats and TCFS coatings both of which were
the mechanical performance of the Tomcats SH coating. It is deposited on silicon wafer substrates and also had similar
anticipated that the alkoxy functional group in the thickness as detailed in Table 3.
fluorosiloxane may increase cross-linking reactions within While both coatings exhibit similar SH water contact
the mixture, and hence may yield a more mechanically angles, the roughness of the TCFS coating is considerably
robust polymer coating.[16] A mixture (50/50, v./v.) of lower. This is demonstrated by the SEM and profilometry
Tomcats and fluorosiloxane was nebulised into the plasma images given in Figure 6. A much higher level of particulates
and deposited onto silicon wafer substrates at a 2 mm gap formation is observed in the case of the Tomcats coating.
distance. The resulting Tomcats / fluorosiloxane coating These particulates are formed due to gas phase plasma
mixture is given the abbreviation TCFS. Based on a series of reactions.[17] The reduced particulate formation in the case
TCFS deposition studies the following were the conditions of the TCFS coating indicates that more polymerisation is
which consistently yielded SH coatings:
7 mlmin1 precursor flow rate, 5 slm
helium flow rate, 100 mlmin1 nitrogen
flow rate, 7.6 W plasma power, 10
mms1 CNC speed and a gap distance
of 2 mm.
FTIR analysis of the fluorosiloxane,
Tomcats and TCFS coatings is given in
Figure 5. The main peaks at 1 270 cm1 is
attributed to the CF2 functional group,
while that at 785 cm1 is ascribed to the
CF3 group. The absorption at 1 100 cm1 is
assigned to the asymmetric SiOSi
stretching.[31] These three functional
groups are consistent with the linear
chemical formula (C14H19F13O3Si) of the
fluorosiloxane monomer, and demon-
strate that the precursor chemistry is
retained in the plasma deposited coating
as reported previously.[16] It should be
noted that the TCFS has peaks observed in
the Tomcats and Fluorosiloxane precur-
sors. The average water contact angle of
the TCFS coated wafers was 1588 and Figure 6. SEM (top) and Optical profilometry (down) images of the Tomcats (left) Ra
surface energy value of 3 mNm1. This 18 nm and TCFS (right) Ra 12 nm coated substrates. The images confirm the presence of
contact angle was obtained with a 10 ml numbers of particulates and lower surface roughness for the TCFS coating.

Plasma Process. Polym. 2011, 8, 305–315


310 ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/ppap.201000069
Enhancing the Mechanical Properties of Superhydrophobic . . .

Table 4. XPS comparison of the elemental composition of the increase in the oxygen content was observed. The
precursor monomer and plasma deposited coating. These were
stoichiometric elemental composition for carbon in the
deposited at a 2 mm gap distance and flow rates of 5 and
7 mlmin1 for the Tomcats and TCFS precursors respectively. TCFS coating reduced by 10% compared to the 1:1 TC:FS
monomer mixture, while the oxygen and fluorine content
C O Si F increased by 3 and 6%, respectively. No significant change
was observed for the elemental composition of silicon. The
at.-% at.-% at.-% at.-% observed decrease in the elemental composition of carbon
and increase in oxygen in both coatings is due to coating
Tomcats Monomer 33 33 33 –
oxidation which is to be expected in an atmospheric
Coating 27 43 27 – pressure helium plasma process.
TCFS Monomer 41 19 15 26 XPS deconvolution of the C 1s curve for both the TCFS and
Coating 31 22 14 32 Tomcats coatings are shown in Figure 7. The curve fitting
confirms that more cross-linking and polymerization
reactions increased in the TCFS coating as anticipated,
taking place on the substrate surface than in the gas phase. with the formation of new bonds (CC¼O) and (CO), that
When comparing the TCFS and Tomcats precursors with were not present in the chemical formula of the mixture.
that of HMDSO, a factor influencing the increase in the level
of the latter precursors reaction in the plasma rather than at
Mechanical Properties of the SH Coatings
the substrate, may be due to its lower flash point. The flash
point of a liquid is that temperature at which the liquid For the comparison trials on the effect of precursor
emits sufficient vapour to form a combustible mixture with chemistry on the mechanical durability of superhydropho-
air.[32] HMDSO precursor has a flash point of 0.6 8C, while bic coatings, in the case of the Tomcats coatings these were
that of the Tomcats and fluorosiloxane precursors are 24 therefore deposited at a gap distance 2 mm and a precursor
and 97 8C, respectively. flow rate of 5 mlmin1. Nano scratch, nano tribometer and
XPS was used to calculate the stoichiometric elemental ultra indentation tests were carried on the HMDSO,
composition based on the chemical formula of Tomcats and Tomcats and TCFS coatings in order to assess their
TCFS (50/50 vol/vol mixture) monomer. The elemental mechanical properties. The Nano Scratch test was used to
composition of the plasma deposited Tomcats and TCFS compare the scratch resistance of the three SH coatings and
coatings compared to these values are given in Table 4. The their adhesion to the silicon wafer substrates. This Nano
results demonstrate that a 6% reduction in the stoichio- Scratch technique uses different sensors (such as force
metric elemental composition (i.e. when compared with the sensor and depth sensor) combined with an optical
Tomcats monomer) for carbon and silicon, while a 10 at.-% microscope to determine the critical loads. The minimum

Figure 7. XPS C 1s curve fitting of the a) TCFS and b) Tomcats coatings.

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.plasma-polymers.org 311
C. E. Nwankire, G. Favaro, Q.-H. Duong, D. P. Dowling

that the coating was not broken. In


order to determine the Nanoindentation
hardness and elastic modulus, a matrix of
6 to 10 indentations were performed on
the coated samples using the Ultra
Nanoindentation tester. The Indentation
Hardness (HIT) and Elastic modulus (EIT) of
the coated and uncoated silicon wafer
samples are given in Table 5. The results
demonstrate that the hardness of the
Figure 8. Image of first critical load (Lc1) (left) and second critical load (Lc2) (right) on the TCFS coating is 19 and 46 times higher
Tomcats SH coating. Lc1 is the normal load at which partial delamination appears on the than the Tomcats and HMDSO super-
film, while Lc2 is the normal load at which the full delamination of the film occurs. hydrophobic coatings respectively. The
elastic modulus of the TCFS coating is 3
and 8 fold higher than that of the Tomcats
load at which adhesive failure occurs is called critical load and HMDSO coatings. Some researchers have suggested
(Lc) and is also known as adhesion strength, which is the that comparing the Hardness and Elastic modulus of
stress required to remove a coating from the substrate.[33] polymers alone could be misleading, as these properties
The first critical load (LC1) is normally ascribed to the onset of could be largely influenced by variations in their micro-
failure, second critical load (LC2) (see Figure 8) indicates film structure, morphology, anisotropy, molecular weight and
delamination.[34] It has been shown that LC can be used to crosslink density.[21,23] The resistance to plastic deforma-
evaluate the performance of a Diamond-like Carbon (DLC) tion (H3/E2) is reportedly preferred for evaluating the wear
[35]
coating and its adhesion to the substrate. A high LC value resistance of any material.[23,34] The calculations suggest
would indicate good scratch resistance and adhesion to the that the TCFS had the highest resistance to plastic
substrate. The Lc1 and Lc2 values for the Tomcats, HMDSO deformation value of 8 in comparison to Tomcats (0.01)
and TCFS coatings are given in Table 5. It should be noted and HMDSO (0.01) coatings. It has been reported that for a
that the Lc1 event did not occur for the TCFS coating. These material with lower E/H ratio any contact with the material
results suggest that the adhesion of the TCFS coating to the will more likely be elastic than plastic.[34] The E/H ratio 7 for
substrate, and its scratch resistance is superior to that of the the TCFS coating is considerably lower compared with the
HMDSO and Tomcats coatings. Another important obser- 41 and 39 for the Tomcats and HMDSO coatings,
vation is that the scratch length at which Lc2 occurred is respectively. This suggests that an elastic deformation
1.4 mm for the TCFS coating; however it is 0.5 and 0.4 mm mechanism is more likely to occur as the TCFS coating
for the HMDSO and Tomcats coatings respectively. transmits an external force to the substrate, than the
The coefficient of friction which is the ratio of the Tomcats and HMDSO coatings. This quality would be
frictional force to the normal load was determined using preferred in industrial applications as the deformation in
Nano Tribometry.[36] A lower shear strength between the the material would be reversible. These results indicate that
sliding ball and the coated surface would result in a lower the mechanical durability of the TCFS superhydrophobic
friction force, and thus lower friction coefficient. It was coating has been enhanced probably due to increased
observed that the TCFS coating has the lowest friction plasma polymerisation and cross-linking reactions with the
coefficient which at 0.34 is lower than that of the Tomcats TCFS coating, which resulted in better coating adhesion to
and HMDSO coatings. The friction coefficient was almost the substrate. It is interesting to note however that the nano
constant in all cases during the test, which is an indication hardness of TCFS coating at 370 MPa (See Table 5) is

Table 5. Critical load, friction coefficient, Nano hardness (H) and Elastic modulus (E) of the coatings and the uncoated silicon wafer.

Coatings Thickness LC1 LC2 Friction H E


coefficient
nm mN mN MPa MPa

Tomcats 370 3.2 15 0.38 19 785


HMDSO 350 8.1 21 0.52 8 313
TCFS 352 – 59 0.34 370 2500
Uncoated Si wafer – – – – 14398 167112

Plasma Process. Polym. 2011, 8, 305–315


312 ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/ppap.201000069
Enhancing the Mechanical Properties of Superhydrophobic . . .

approximately 6 times higher than the nano hardness


reported for a plasma fluorinated SH coating at 64 MPa,
which was thermally treated for enhanced mechanical
durability.[14] In order to verify and compare the results,
similar tests were conducted on an uncoated silicon wafer
substrate and the H3/ E2 calculation equates to 107, while E/
H ratio is 12. This is to be expected as the uncoated silicon
wafer substrate exhibits much higher hardness with better
resistance to plastic deformation than the plasma poly-
merised coatings.

Abrasion Tests
These tests were carried out for the SH coated silicon wafer
substrates using the ultrasonic probe immersed in the Figure 9. Abrasion tests for the Tomcats and TCFS coatings with
water / SiC powder mixture. The test results given in 370 nm and 352 nm thickness, respectively. These coatings were
Figure 9 and Table 6 are the average obtained for 3 samples deposited at a 2 mm gap distance. The 350 nm thick HMDSO
per test. All coatings exhibited a loss in superhydrophobic coating was deposited at an 8 mm gap distance.
properties, but no coating was found to delaminate after
the 150 min abrasion treatment. The water contact angle of
the TCFS coating decreased from 156 to 958 during the XPS analysis was used to compare the siloxy chemistry of
abrasion treatment (39%) reduction after the test period, in the as deposited Tomcats, HMDSO and TCFS coatings. The
comparison to Tomcats 73% (154 to 418) and HMDSO 56% three superhydrophobic coatings were deposited at the gap
(156 to 688) coatings. These abrasion test results correlate distances detailed in Figure 9. The silicon Si 2p core level
with the Nano scratch test results detailed earlier. curve was fitted according to a methodology previously
FTIR peak integration[16] of the SiO and SiCH3 reported by O’Hare et al.[37] The simplified notation M
functional groups for the three coatings before and after [(CH3)3SiO1/2], D [(CH3)2SiO2/2], T [(CH3)SiO3/2] and Q [SiO4/2]
the 150 min abrasion treatment is presented in Table 6. The was used to represent the siloxy unit indicating the number
results confirm that the coatings survived the abrasion test of atoms attached to the silicon. A high Q-type siloxy
with some slight reduction in the peak integration chemistry would suggest increased oxidation, whereas
intensities, which is probably due to the wearing away high M-type siloxy chemistry would indicate high organic
and subsequent reduction in the coating thickness. An content.
analysis of FTIR peak intensity ratios of the SiO to SiCH3 As demonstrated in Table 7, the TCFS coating was more
peaks before and after the 150 min abrasion tests demon- oxidised with the highest Q-type siloxy chemistry than the
strate a small increase in the relative intensity of the SiO Tomcats and HMDSO coatings. The enhanced mechanical
peaks for all three coatings. This suggests a preferential loss performance of the TCFS as demonstrated earlier using
in the organic fraction of the coatings, which would Nano Scratch, Nanoindentation, Nano Tribology and Nano
contribute to the loss in coating superhydrophobicity. Abrasion tests may be associated with this higher level of Q-
Although this abrasion test should only be considered as type and T-type siloxy chemistry. In a previous study by
qualitative rather than quantitative it was used to conclude Twomey et al.[38] comparing the mechanical behaviour
that coating abrasion durability of the SH coatings is ranked of nm thick atmospheric plasma coatings, it was demon-
as follows: TCFS > HMDSO > TC. strated that coatings with high Q-type siloxy chemistry

Table 6. FTIR peak integration of the SiO and SiCH3 peaks of the coated samples after abrasion tests.

Coatings SiO SiCH3 SiO / SiCH3 SiO / SiCH3

a.u. a.u.

Before After Before After Before After

Tomcats 65  3 45  4 9  0.3 6.1  0.1 7.22 7.38


HMDSO 52  3 41  2 7  0.1 5.5  0.1 7.43 7.45
TCFS 58  2 52  2 7.8  0.2 6.9  0.1 7.44 7.54

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.plasma-polymers.org 313
C. E. Nwankire, G. Favaro, Q.-H. Duong, D. P. Dowling

Table 7. XPS comparison of the siloxy chemistry in the SH. the needle-like profile observed in the 3-D scan of the
HMDSO coating. The relative height of the ‘needle’ peaks is
Siloxy chemistry Tomcats HMDSO TCFS reduced in the case of the Tomcats and particularly that of
% % % the TCFS coatings. The enhanced durability observed in the
TCFS coating may be partially associated with this
M 0.6 3.1 0 reduction in needle peak height particularly when com-
D 35.9 37.7 0.9 bined with enhanced levels of plasma polymerisation and
cross-linking.
T 43.2 33.4 67.5
Q 20.3 25.8 31.6
Conclusion
after increased plasma exposure had better mechanical
properties than their counterpart with lower Q-type Superhydrophobic coatings have been deposited using the
chemistry. It should be noted that, although the TCFS PlasmaStreamTM atmospheric plasma jet system using
coating exhibited higher levels of oxidation with high Q- HMDSO, Tomcats and the mixture of Tomcats and
type and T-type siloxy chemistry, it also exhibits an fluorosiloxane (TCFS) precursors. The main focus of the
increased inorganic character. It is anticipated that this research was to enhance the durability of the super-
would tend to increase the coating hydrophilicity. The hydrophobic coatings. This was successfully achieved
superhydrophobic properties of TCFS are therefore asso- through the deposition of the TCFS coating. XPS analysis
ciated with the presence of fluorine from the fluorosiloxane demonstrated that this coating exhibited enhanced levels
precursor. of cross-linking and polymerisation reaction compared
A 2-D optical profile of the Tomcats, HMDSO and TCFS with the Tomcats and HMDSO coatings. Nano Scratch, Nano
superhydrophobic coatings was obtained by scanning Tribology and Ultra Nanoindentation tests also demon-
areas of 45  59 mm2 using the optical profilometry strated that the TCFS coating also exhibited enhanced
(Figure 10). The profiles of the coatings further demonstrate adhesion, hardness and wear resistance compared with
these coatings. The abrasion resistance of the coatings was
assessed using an ultrasonic jet abrasion technique and the
abrasion performance correlate closely with the other
tribological tests. In addition to enhance polymer cross-
linking, a factor influencing the enhanced wear and
abrasion performance of the TCFS coating, is that its
superhydrophobic properties are achieved with a reduction
in the height of the ‘needle-like’ morphology, compared to
the other coatings examined. It is concluded from this study
that the superhydrophobic TCFS coating possess both
superior mechanical properties and better abrasive wear
resistance compared with that achieved with Tomcats and
HMDSO coatings.

Acknowledgements: This work is supported by Science Founda-


tion Ireland Grant 08/SRC11411.

Received: May 13, 2010; Revised: October 23, 2010; Accepted:


October 26, 2010; DOI: 10.1002/ppap.201000069

Keywords: fluorosiloxane; mechanical durability; nanoindenta-


tion; nano scratch; oleophilic; superhydrophobic

[1] H. J. Lee, S. Michielsen, J. Polym. Sci. B: Polym. Phys. 2007, 45,


253.
Figure 10. 2-dimensional optical profile of the SH Tomcats, [2] J. Genzer, K. Efimenko, Biofouling 2006, 22, 339.
HMDSO and TCFS coatings. [3] D. Oner, T. J. McCarthy, Langmuir 2000, 16, 7777.

Plasma Process. Polym. 2011, 8, 305–315


314 ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim DOI: 10.1002/ppap.201000069
Enhancing the Mechanical Properties of Superhydrophobic . . .

[4] T. Onda, S. Shibuichi, N. Satoh, K. Tsujii, Langmuir 1996, 12, [22] M. Doerner, W. Nix, J. Mater. Res. 1986, 1, 601.
2125. [23] A. Leyland, A. Matthews, Wear 2000, 246, 1.
[5] A. Milella, R. D. Mundo, F. Palumbo, P. Favia, F. Fracassi, [24] Y. Xiu, Y. Liu, D. W. Hess, C. P. Wong, Nanotechnology 2010, 21,
R. d’Agostino, Plasma Processes Polym. 2009, 6, 460. 155705.
[6] H. Erbil, Science 2003, 299, 1377. [25] L. O’Neill, C. O’Sullivan, Chem. Vap. Deposition 2009, 15,
[7] K. Tadanaga, N. Katata, T. Minami, J. Am. Ceram. Soc. 1997, 80, 21.
3213. [26] J. Widodo, W. Lu, S. Mhaisalkar, J. Sudijono, L. Hsia, L. Shen,
[8] M. Li, J. Zhai, H. Liu, Y. Song, L. Jiang, D. Zhu, J. Phys. Chem. B Z. Zeng, Thin Solid Films 2005, 472, 195.
2003, 107, 9954. [27] A. Fidalgo, L. M. Ilharco, J. Non-Cryst. Solids 2001, 283, 144.
[9] L. O’Neill, A. P. F. Herbert, C. Stallard, D. P. Dowling, Plasma [28] A. Gupta, S. Kluge, C. Schulz, H. Wiggers, G. Duisburg, Mat. Res.
Processes Polym. 2010, 7, 43. Soc. Symp. Proc. 2010, 1207, 1207–N02.
[10] S. Yang, C. Liu, C. Su, H. Chen, Thin Solid Films 2009, 517, 5284. [29] A. Tuteja, W. Choi, M. Ma, J. M. Mabry, S. A. Mazzella,
[11] Y.-Y. Ji, S.-S. Kim, O. P. Kwon, S.-H. Lee, Appl. Surf. Sci. 2009, 255, G. C. Rutledge, G. H. McKinley, R. E. Cohen, Science 2007,
4575. 318, 1618.
[12] S. H. Kim, J.-H. Kim, B. K. Kang, H. S. Uhm, Langmuir 2005, 21, [30] L. Feng, Z. Zhang, Z. Mai, Y. Ma, B. Liu, L. Jiang, D. Zhu, Angew.
12213. Chem. 2004, 116, 2046.
[13] G. Chen, S. Chen, W. Feng, W. Chen, S.-Z. Yang, Appl. Surf. Sci. [31] N. B. Colthup, L. H. Daly, S. E. Wiberley, Introduction to
2008, 254, 3915. infrared and Raman spectroscopy, Academic Press, New York
[14] I. Woodward, W. C. E. Schofield, V. Roucoules, J. P. S. Badyal, 1975.
Langmuir 2003, 19, 3432. [32] H.-J. Liaw, Y.-Y. Chiu, J. Hazard. Mater. 2003, 101, 83.
[15] J. Albaugh, C. O’Sullivan, L. O’Neill, Surf. Coat. Technol. 2008, [33] H. Park, D. Kwon, Thin Solid Films 1997, 307, 156.
203, 844. [34] D. V. Shtansky, N. A. Gloushankova, I. A. Bashkova, M. I.
[16] C. E. Nwankire, M. Ardhaoui, D. P. Dowling, Polym. Int. 2009, Petrzhik, A. N. Sheveiko, F. V. Kiryukhantsev-Korneev, I. V.
58, 996. Reshetov, A. S. Grigoryan, E. A. Levashov, Surf. Coat. Technol.
[17] D. P. Dowling, A. Ramamoorthy, M. Rahman, D. A. Mooney, 2006, 201, 4111.
J. M. D. MacElroy, Plasma Processes Polym. 2009, 6, S483. [35] G. Ramı́rez, E. Tarrés, B. Casas, I. Valls, R. Martı́nez, L. Llanes,
[18] D. K. Owens, R. C. Wendt, J. Appl. Polym. Sci. 1969, 13, 1741. Plasma Processes Polym 2009, 6, S588.
[19] B. Twomey, M. Rahman, G. Byrne, A. Hynes, L.-A. O’Hare, [36] B. Bhushan, J. Israelachvili, U. Landman, Nature 1995, 374,
L. O’Neill, D. P. Dowling, Plasma Processes Polym. 2008, 5, 737. 607.
[20] P. Steinmann, Y. Tardy, H. Hintermann, Thin Solid Films 1987, [37] L.-A. O’Hare, B. Parbhoo, S. R. Leadley, Surf. Interface Anal.
154, 333. 2004, 36, 1427.
[21] M. VanLandingham, J. Villarrubia, W. Guthrie, G. Meyers, [38] B. Twomey, G. Byrne, A. Hynes, L. O’Neill, D. Dowling, Surf.
Macromolecular symposia 2001, 167, 15. Coat. Technol. 2009, 203, 2021.

Plasma Process. Polym. 2011, 8, 305–315


ß 2011 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim www.plasma-polymers.org 315

You might also like