You are on page 1of 8

Applied Surface Science 356 (2015) 408415

Contents lists available at ScienceDirect

Applied Surface Science


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

Reactive-ion etching of nylon fabric meshes using oxygen plasma for


creating surface nanostructures
Hernando S. Salapare III, Thierry Darmanin, Frdric Guittard
Universit Nice Sophia Antipolis, CNRS, LPMC, UMR 7336, Parc Valrose, 06100 Nice, France

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

Article history: A facile one-step oxygen plasma irradiation in reactive ion etching (RIE) conguration is employed to
Received 6 May 2015 nylon 6,6 fabrics with different mesh sizes to achieve surface nanostructures and improved wettability
Received in revised form 25 June 2015 for textile and ltration applications. To observe the effects of power and irradiation time on the samples,
Accepted 28 June 2015
the experiments were performed using constant irradiation time in varying power and using constant
Available online 10 August 2015
power in varying irradiation times. Results showed improved wettability after the plasma treatment. The
FTIR spectra of all the treated samples exhibited decreased transmittance of the amide and carboxylic
Keywords:
acid groups due to surface etching. The changes in the surface chemistry are supported by the SEM data
Nylon fabric
Reactive-ion etching
wherein etching and surface nanostructures were observed for the plasma-treated samples. The etching
Oxygen plasma of the surfaces is enhanced for higher power plasma treatments. The thermal analysis showed that the
Surface modication plasma treatment resulted in decreased crystallinity. Surface chemistry showed that the effects of the
Texturing plasma treatment on the samples have no signicant difference for all the mesh sizes. However, surface
DC self-bias voltage morphology showed that the sizes of the surface cracks are the same for all the mesh sizes but samples
with larger mesh sizes exhibited enhanced etching as compared to the samples with smaller mesh sizes.
Higher power induced higher negative DC self-bias voltage on the samples that favored anisotropic and
aggressive etching.
2015 Elsevier B.V. All rights reserved.

1. Introduction susceptibility of poly(ethylene terephthalate) (PET) and nylon


6,6 bers [10], and to achieve superhydrophilic or super-
Nylon 6,6 (polyhexamethylene adiamide) fabrics are semi- hydrophobic properties for lignocellulosic seagrass, PET, and
crystalline polyamides that possess outstanding versatile proper- poly(tetrauoroethylene) (PTFE) materials [1114]; gas discharges
ties that are used in textile, automobile, military, electronics, and in improving the wettability, superhydrophobicity and antibacte-
medical industries. Although the bers possess excellent mechan- rial properties of PTFE [1518]; laser to improve the dyeability of
ical, chemical, and thermal properties, there are still some limita- poly(amide) 6,6 knitted fabrics [19]; and microwave jet plasma for
tions on its uses and applications [110]. In general, to extend the the functionalization of electrospun poly(amide) 6 nanobers to
use and applications of polymers, several treatments are employed improve the adhesion properties during the production of com-
and some of them uses atmospheric pressure plasma to improve the posites in tissue engineering [7].
dye uptake of nylon 6 bers with different absorbed moisture [1] In this study, RF oxygen plasma in reactive-ion etching (RIE) con-
and to enhance the antibacterial activity and natural dyeing proper- guration was used to modify the chemistry and morphology of the
ties of nylon 6 fabrics [6]; direct low-pressure and low-temperature nylon 6,6 samples. This technique has some advantages over the
plasma to enhance the adhesion of carbon nanotubes on nylon other techniques since it is an extreme form of direct plasma treat-
6,6 fabrics [3], to improve the mechanical properties of nylon 6 ment and it has an increased DC bias level that favors anisotropic
plain woven fabrics [5], to correlate the crystallinity and plasma and aggressive etching on the samples [20,21]. The plasma cong-
uration under vacuum is favorable in the processing of materials
since the integrity of the plasma and the materials are maintained
throughout the irradiation process [1113,1518]. Plasma treat-
Corresponding author. Tel.: +334 92 07 61 59; fax: +334 92 07 61 56.
ments were found to be favorable over non-plasma techniques
E-mail addresses: salapare@unice.fr (H.S. Salapare III), darmanin@unice.fr
(T. Darmanin), Frederic.GUITTARD@unice.fr (F. Guittard). since the plasma may contain different species such as electrons,

http://dx.doi.org/10.1016/j.apsusc.2015.06.185
0169-4332/ 2015 Elsevier B.V. All rights reserved.
H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415 409

neutral species, charged species, and excited species that can inter- (Hm ) were extracted from the melting prole and the crystalliza-
act with the materials surface [22]. tion temperature (Tc ), and enthalpy of crystallization (Hc ) were
extracted from the crystallization prole. The crystallinity of the
2. Experimental procedures fabric samples were then calculated using the following equation:

2.1. Materials and plasma treatment Hm Hc


% Crystallinity = o 100%
Hm
Sefar Nitex nylon 6,6 fabrics (Sefar AG, Switzerland) of differ-
where Hm o is a reference value and represents the heat of melting
ent mesh sizes ranging from 25 m to 1550 m were used in the
experiments. Different mesh sizes were used because the pressure if nylon 6,6 were 100% crystalline [23].
that a fabric can support is highly dependent on the mesh size.
Before plasma treatment, the nylon 6,6 fabrics were soaked in ace- 2.2.4. Scanning electron microscopy
tone for 10 min followed by air-drying for 2 h at room temperature A JSM 6700F NT (JEOL Ltd., Japan) scanning electron microscope
in order to remove any impurities on the surface. The cleaned and was used to determine the morphology of the surface of the mate-
dried samples were then irradiated using oxygen plasma produced rial. The samples were scanned at 250, 10,000, and 25,000
in reactive ion etching conguration of a BSET EQ NT-1 plasma magnications.
device (Digit Concept Microelectronics and High Tech Equipment,
France) operating at an excitation frequency of 13.56 MHz. The 3. Results and discussion
plasma device operates on the principle of a capacitive coupled
plasma (CCP) reactor where the RF current through the oxygen Control nylon 6,6 fabrics of mesh sizes 25 m, 100 m, 200 m,
gas maintains a steady state discharge by heating the electrons. and 400 m exhibited high wettability since the water contact
The samples were placed on the power electrode, which is paral- angles are less than 10 and the water droplet spread on the surface
lel to the ground electrode at 50 mm distance from each other. A within 1 s of contact. This was expected since the nylon 6,6 fabrics
20 m3 /h rotary pump evacuated the system, and the base pressure used in this study are usually used for ltration applications. For
was kept at 100 mTorr. Oxygen gas was fed into the chamber at the control nylon 6,6 fabrics of 700 m and 1550 m mesh sizes,
a ow rate of 20 sccm. The plasma sheath produced between the the contact angle is 84 and 43 , respectively, since the volume of
power and ground electrodes induced negative DC self-bias volt- the water droplet used in the measurement is 2 L. This means
age on the samples that was measured using a voltmeter with a that the ber spacing is an important factor when studying wett-
low-pass lter. ability, for example, when the material is used for ltration, the
In order to determine the effects of power and irradiation time selection of an appropriate lter is necessary for different types of
on the treatment of the fabrics, two sets of plasma parameters were ltrates. After the plasma treatment, all the nylon 6,6 fabric sam-
used. The rst experiment was done with constant irradiation time ples exhibited improved wettability, especially for samples of mesh
of 600 s but with varying powers of 100 W, 200 W, and 300 W, and sizes 700 m and 1550 m. This improvement in the wettability
the second experiment was done with constant power of 600 W but is favorable when dealing with ltration using different types of
with varying irradiation times of 100 s, 200 s, and 300 s. The exper- ltrates.
imental parameters used in the rst set have the same incident For the FTIR, DSC, and SEM data, only the nylon 6,6 fabric with
energy as the second set ranging from 60 kJ to 180 kJ. mesh size of 100 m will be shown, since it was found in the study
that the effects of the plasma treatment on the samples have no
2.2. Characterizations signicant difference for all the mesh sizes (i.e., the same trend and
behavior on the surface chemistry, surface morphologies, thermal
2.2.1. Contact angle measurements properties, and crystallinity were observed for samples of mesh
A DSA-30 goniometer (Krss GmbH) equipped with a drop shape sizes 25 m, 100 m, 200 m, 400 m, 700 m and 1550 m).
analysis (DSA4) software was used to study the changes in the Fig. 1 shows the superimposed view of FTIR spectra of the con-
wettability of the samples. Sessile drop method was used to deter- trol and plasma-treated nylon 6,6 fabric samples with 100 m mesh
mine the apparent contact angles. 2 L of Milli-Q deionized water size. For spectra b, c, and d, the plasma treatment was performed
was dropped vertically onto the samples using a motorized syringe with constant irradiation time (600 s) of varying power (100 W,
mechanism. For each samples, contact angle measurements were 200 W, and 300 W). The spectrum of the control sample veries
performed at three different sites. that the material is nylon 6,6 since it contains the following chem-
ical functionalities: N H stretching (3264 cm1 ), C H asymmetric
2.2.2. Fourier transform infrared spectroscopy stretching (3085 cm1 ), asymmetric and symmetric stretching
The changes in the samples chemical functionalities were (2962 cm1 and 2851 cm1 , respectively), C O stretching from car-
obtained using a Spectrum 100 FT-IR spectrometer (Perkin Elmer, boxylic acid groups (1734 cm1 ), amide I band and amide II and
USA) with a diamond attenuated total reectance (ATR) top plate CH2 asymmetric deformation groups (1650 cm1 and 1539 cm1 ,
accessory. The samples were scanned 5 times at 4 cm1 spectral respectively), N H deformation and CH2 scissoring (1452 cm1 ),
resolutions over the range of 450 cm1 to 4000 cm1 . The spectra amide III band and CH2 wagging (1347 cm1 ), CCH symmet-
reported in this paper were obtained by subtracting the background ric bending and CH2 twisting (1145 cm1 and 1130 cm1 ), C C
spectra from the measured spectra. stretching (962 cm1 ), N H wagging and CH2 rocking (742 cm1 ),
C C bending (600 cm1 ), and O C N bending (548 cm1 ).
2.2.3. Differential scanning calorimetry For all the peaks except the C C stretching, it was found that the
Melting and crystallization proles were determined for each transmittance intensities decreased which are proportional to the
sample using a Jade Differential Scanning Calorimeter (DSC) (Perkin values of the incident energy. The existence of the amide I, amide
Elmer, USA). Samples were lled in a measurement cell of the DSC II, and carboxylic acid groups which are polar and hydrophilic sup-
and the temperature was rst heated from 25 C to 300 C at the rate ports the over-all hydrophilic properties of the nylon 6,6 fabrics.
of 20 C/min and held for 3 min. The samples were then quenched to The decrease in the transmittance peaks can be associated with the
25 C from 300 C at the rate 100 C/min. The glass transition tem- breakage of some amide linkage in the nylon 6,6 fabrics which cre-
perature (Tg ), melting temperature (Tm ), and enthalpy of melting ates free amino groups, this change in the chemistry of the material
410 H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415

Fig. 1. Superimposed view of FTIR spectra of the control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size. The plasma treatment was performed with
constant irradiation time (600 s) of varying power (100 W to 300 W).

Fig. 2. Superimposed view of FTIR spectra of the control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size. The plasma treatment was performed with
constant power (600 W) of varying irradiation times (100 s to 300 s).
H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415 411

Fig. 3. Superimposed view of the FTIR spectra of the plasma-treated nylon 6,6 fabric samples with 100 m mesh size. The two spectra represent the same incident energy
equal to 180 kJ but of different power (300 W and 600 W) and different irradiation time (600 s and 300 s).

is favorable for textile applications since it improves its dyeability with the materials surface. The higher the power also means that
as seen from the work of Bahtiyari [19] on polyamide fabric surface the plasma temperature is high which would possibly increase the
modication using laser. Another possible effect of the decrease in materials dyeability as shown by Kamide et al. [25] on their study
the transmittance peaks of amide I, amide II, and carboxylic acid of the structural factors governing the dyeability of poly(ethylene)
groups (which are sensitive to light and solvents) is the improve- terephthalate (PET) bers.
ment of the materials stability to light and solvents like alkali and Fig. 4 shows the dependence of the negative DC self-bias volt-
acids as shown by El-Garf et al. [24] on their work on treatment of age on power. The negative DC self-bias voltage increases linearly
nylon 6 materials using dichlorobenzenesulfonyl chloride. (R2 = 0.96553) with the square root of the power when the power
Fig. 2 shows the superimposed view of FTIR spectra of the control varies from 100 W to 600 W at constant pressure; this plasma
and plasma-treated nylon 6,6 fabric samples with 100 m mesh behavior is consistent with the capacitive sheath model proposed
size. For spectra b, c, and d, the plasma treatment was performed by Catherine et al. [26]. By correlating the FTIR data with Fig. 4, we
with constant power (600 W) of varying irradiation times (100 s, can assume that the decreases in the spectral transmittance peaks
200 s, and 300 s). The same observations as in Fig. 1 regarding the are proportional with the negative DC self-bias voltage induced on
transmittance peaks of all the chemical functionalities were seen the samples.
in the FTIR spectra in Fig. 2. Fig. 5 shows the stacked view of the melting diagrams for the
The main difference between Figs. 1 and 2 can be seen in Fig. 3 control and plasma-treated nylon 6,6 fabric samples with 100 m
which shows the superimposed view of the FTIR spectra of the mesh size obtained in the heating measurement and Fig. 6 shows
plasma-treated nylon 6,6 fabric samples with 100 m mesh size the stacked view of the crystallization diagrams for the control and
wherein the two spectra represent the same incident energy equal plasma-treated nylon 6,6 fabric samples with 100 m mesh size
to 180 kJ but of different power (300 W and 600 W) and different obtained in the cooling measurement. Based on the melting and
irradiation time (600 s and 300 s). It can be seen from Fig. 3 that crystallization diagrams, the thermal properties and crystallinity
the transmittance values of the 600 W, 300 s plasma-treated sam- of the nylon 6,6 fabrics were extracted and summarized in Table 1.
ples were higher than the 300 W, 600 s plasma-treated samples. The glass transition temperature of the plasma-treated samples
This observation is the same with the other samples of the same were found to be signicantly higher than the control sample, this
incident energy but of different power and irradiation time. From is an indication that the material after undergoing plasma treat-
this data, it can be assumed that the power plays the main part in ment has changed its molecular arrangement as also supported by
the surface modication process. Higher power treatments create the signicant decrease in the crystallinity of the samples [2,8].
more plasma particles than the lower power treatments [11,22] The decrease in the crystalline phase gives rise to the increase in
and the effect of the irradiation time was found to be of lower sig- the amorphous phase of the material. This change in the materials
nicance when dealing with the interaction of the plasma particles phase is favorable for textile applications as shown by Okuno et al.
412 H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415

Fig. 4. Dependence of negative DC self-bias voltage on power.

Fig. 5. Stacked view of the melting diagrams for the control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size obtained in the heating measurement.

Fig. 6. Stacked view of the crystallization diagrams for the control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size obtained in the cooling measurement.
H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415 413

Table 1
Summary of the thermal properties and the crystallinity for the control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size.

Experimental parameters Tg ( C) Tm ( C) Tc ( C) Hm (J/g) Hc (J/g) Crystallinity (%)

Control 123.2 220.2 173.8 137.9 1.9 53.1


100 W, 600 s 132.7 219.5 173.7 133.2 4.1 50.4
200 W, 600 s 133.2 219.8 173.2 115.5 4.8 47.0
300 W, 600 s 135.4 219.9 175.0 119.5 5.2 44.6
600 W, 100 s 132.0 219.5 176.5 116.8 8.1 42.5
600 W, 200 s 133.7 220.7 181.2 107.5 1.3 42.5
600 W, 300 s 137.3 220.8 172.5 103.5 2.5 39.5

[10] on their work on the effect of the crystallinity of PET and nylon Fig. 7 shows the representative SEM micrographs of the (a, d, g)
6,6 bers on the plasma etching and dyeability characteristics by control and plasma-treated nylon 6,6 fabric samples with 100 m
low temperature plasma treatment, wherein they found out that mesh size under (ac) 250, (df) 10,000, and (gi) 25,000
the increase in the materials crystallinity decreases its dyeability magnications. The SEM images of the treated samples represent
since the dyeable area in a material lies on the amorphous phase. the same incident energy equal to 180 kJ but of different power
The decrease in the crystallinity is supported by the FTIR data and different irradiation time: (b, e, h) 300 W, 600 s treated (c, f, i)
where there is an increase in the transmittance peaks in the 600 W, 300 s treated.
C C stretching bands of the plasma-treated samples, which cor- It can be observed that the plasma treatment induced etching
responds to the amorphous phase of the sample [19]. There is no as seen from the ripple-like structures on the surface of nylon 6,6
signicant change with the melting and crystallization tempera- fabrics. A closer look on the surfaces structures at 25,000 magni-
ture of the control and plasma-treated samples since based from the cations revealed that the plasma treatment, which induced etching,
FTIR data, the chemical composition did not changed due to plasma created surface cracks with openings of sizes ranging from 0.50 m
treatment. However, there is a general decrease in the enthalpy to 1.00 m. By comparing the 300 W, 600 s plasma-treated and
of melting and increase in the enthalpy of crystallization after the 600 W, 300 s plasma-treated samples reveals that the etching of the
plasma treatment, which also effected the decrease in the crys- samples surfaces also conrms the decrease in the transmittance
tallinity of the plasma-treated materials. This also means that the peaks of all the chemical groups in the FTIR spectra of the plasma
energy per mass unit being absorbed or released by the sample treated samples, this result is consistent with the BeerLamberts
changed due to the decrease in the transmittance peaks of all the law since the concentration of the functional groups decreased
chemical groups as seen from the FTIR data. because of the etching.

Fig. 7. Representative SEM micrographs of the (a, d, g) control and plasma-treated nylon 6,6 fabric samples with 100 m mesh size under (ac) 250, (df) 10,000, and
(gi) 25,000 magnications. The SEM images of the treated samples represent the same incident energy equal to 180 kJ but of different power and different irradiation
time: (b, e, h) 300 W, 600 s treated (c, f, i) 600 W, 300 s treated.
414 H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415

Fig. 8. SEM micrographs of the plasma-treated nylon 6,6 fabric samples with mesh sizes of (a and g) 25 m, (b and h) 100 m, (c and i) 200 m, (d and j) 400 m, (e and k)
700 m, and (f and l) 1550 m under 25,000 magnication. The SEM images from (af) represents the same incident energy as of (gl) but of different power and different
irradiation time: (af) 300 W, 600 s treated (gl) 600 W, 300 s treated.

In order to compare the etching effects on the different mesh treated samples exhibited decreased transmittance of the amide
sizes, Fig. 8 shows the SEM micrographs of the plasma-treated and carboxylic acid bands due to surface etching. The changes
nylon 6,6 fabric samples with mesh sizes of (a, g) 25 m, (b, h) in the surface chemistry are supported by the SEM data wherein
100 m, (c, i) 200 m, (d, j) 400 m, (e, k) 700 m, and (f, l) 1550 m etching and texturization of the surfaces were observed for the
under 25,000 magnication. The SEM images from (af) repre- plasma-treated samples. The etching of the surfaces is enhanced
sents the same incident energy as of (gl) but of different power and for higher power plasma treatments. Increasing the irradiation time
different irradiation time: (af) 300 W, 600 s treated (gl) 600 W, also enhances the etching on the samples surfaces, however; it is
300 s treated. The surface crack openings for all the mesh sizes the power and the negative DC self-bias voltage that played the
ranges from 0.05 m to 0.25 m for 300 W, 600 s treatments and major role in the etching process. The thermal analysis showed
from 0.50 m to 1.00 m for 600 W, 300 s treatments. The power that the plasma treatment resulted to decreased crystallinity and in
has a greater morphological effect on the surfaces as compared with consequence increased the amorphous phase of the samples. The
the irradiation time, which is also consistent with the observations data of the surface chemistry showed that the effects of the plasma
seen from the FTIR and DSC data. Although the increase in the sur- treatment on the samples have no signicant difference for all the
face crack openings are relatively the same for all the mesh sizes, a mesh sizes. However, surface morphology showed that the sizes of
closer look on Fig. 8 reveals that there is a greater etching on the fab- the surface cracks are the same for all the mesh sizes but samples
rics surfaces for larger mesh sizes (400 m, 700 m, and 1550 m) with larger mesh sizes exhibited enhanced etching as compared
as compared with the smaller mesh sizes (25 m, 100 m, 200 m) to the samples with smaller mesh sizes. Higher power induced
as exhibited by the enhanced ripple-like surface structures. Fabrics higher negative DC self-bias voltage on the samples that favored
with larger mesh sizes are more susceptible to etching since larger anisotropic and aggressive etching. The results of this study are
surface areas are exposed to the oxygen plasma. favorable for applications that need properties such as improved
The negative DC self-bias voltage also played a main role in dyeability, increased stability to light, and increased resistance to
the etching process; high power induces higher negative DC self- alkali and acid solvents.
bias voltage that increases the electronic density in the plasma and
the number of plasma species that interacts with the surface of
Acknowledgements
the materials [27]. Samples treated at longer irradiation times also
showed more enhanced and aggressive etching because etching
We would like to thank the Erasmus Mundus Mobility with Asia
rate is higher for samples exposed to high negative DC self-bias
(EMMA-West)2013 (Lot 12) for the nancial grant (Grant No.
voltage [28].
2012-2643/001-001-EMA 2) and the Centre Commun de Micro-
By visual inspection of Fig. 7, we can say that there are increases
scopie Applique of the Universit Nice Sophia Antipolis and the
in the surface roughness of the nylon 6,6 fabric surfaces after plasma
Microscopy and Imaging platform Cte dAzur (MICA) for the SEM
treatment. It can be assumed that the result of the plasma treatment
characterization.
on the materials wettability is consistent with the Wenzel wetting
state where hydrophilic surfaces tend to be even more hydrophilic
(when the apparent water contact angle of a smooth surface,  Y is References
less than 90 , the apparent contact angle of a rough surface,  W is
always less than ) when the surface roughness increases [2931]. [1] L. Zhu, C. Wang, Y. Qiu, Surf. Coat. Technol. 201 (2007) 74537461.
[2] T. Matsuda, T. Shimomura, M. Hirami, Polym. J. 31 (1999) 795800.
[3] W. Zhang, L. Johnson, S.R.P. Silva, M.K. Lei, Appl. Surf. Sci. 258 (2012)
82098213.
4. Conclusions [4] S. Dasgupta, W.B. Hammond, W.A. Goddard III, J. Am. Chem. Soc. 118 (1996)
1229112301.
[5] J. Yip, K. Chan, K.M. Sin, K.S. Lau, J. Mater. Process. Technol. 123 (2002) 512.
A facile one-step oxygen plasma irradiation in reactive-ion [6] A. Haji, A.M. Shoushtari, M. Mirafshar, Color. Technol. 130 (2013) 3742.
etching (RIE) conguration was employed to nylon 6,6 fabrics [7] D. Pavlink, J. Hnilica, A. Quade, J. Schfer, M. Alberti, V. Kudrle, Polym.
with different mesh sizes in achieving nanostructured surface and Degrad. Stab. 108 (2014) 4855.
[8] S.P. Rwei, Y.C. Tseng, K.C. Chiu, S.M. Chang, Y.M. Chen, Thermochim. Acta 555
improved wettability for textile and ltration applications. The
(2013) 3745.
effects of power and irradiation time on the samples were deter- [9] Z. Gao, J. Sun, S. Peng, L. Yao, Y. Qiu, J. Appl. Polym. Sci. 120 (2011) 22012206.
mined by using different parameters: (1) constant irradiation time [10] T. Okuno, T. Yasuda, H. Yasuda, Text. Res. J. 62 (1992) 474480.
in varying power and (2) constant power in varying irradiation [11] H.S. Salapare III, F. Guittard, X. Noblin, E. Tafn de Givenchy, F. Celestini, H.J.
Ramos, J. Colloid Interface Sci. 396 (2013) 287292.
times. In general, improved wettability was observed for all the [12] H.S. Salapare III, M.G.J.P. Tiquio, H.J. Ramos, Appl. Surf. Sci. 273 (2013)
samples after the plasma treatment. The FTIR spectra of all the 444447.
H.S. Salapare III et al. / Applied Surface Science 356 (2015) 408415 415

[13] J. Tarrade, T. Darmanin, E. Tafn de Givenchy, F. Guittard, D. Debarnot, F. [21] E.H. Kim, T.Y. Lee, B.C. Min, C.W. Chung, Thin Solid Films 521 (2012)
Poncin-Epaillard, Appl. Surf. Sci. 292 (2014) 782789. 216221.
[14] S. Zanini, R. Barni, R. Della Pergola, C. Riccardi, J. Phys. D: Appl. Phys. 47 (2014) [22] R. Hrach, V. Hrachov, M. Vicher, Proc. ICPP-25, Prague, Europhys. Conf. Abs.
325202. 22C (1998) 26472650.
[15] H.S. Salapare III, G.Q. Blantocas, W.L. Rivera, V.A. Ong, R.S. Hipolito, H.J. Ramos, [23] W.J. Sichina, Thermal Analysis, Perkin Elmer Inc., Connecticut, 2000.
Plasma Fusion Res. 6 (2011) 2406043. [24] S.A. El-Garf, S.M. El-Kemry, Text. Res. J. 67 (1997) 1317.
[16] H.S. Salapare III, G.Q. Blantocas, V.R. Noguera, H.J. Ramos, Appl. Surf. Sci. 255 [25] K. Kamide, T. Kuriki, Polym. J. 19 (1987) 315322.
(2008) 29512957. [26] Y. Catherine, P. Couderc, Thin Solid Films 144 (1985) 265280.
[17] H.S. Salapare III, B.A.T. Suarez, H.S.O. Cosinero, M.Y. Bacaoco, H.J. Ramos, [27] S. Djerourou, K. Henda, M. Djebli, J. Mod. Phys. 2 (2011) 954957.
Mater. Sci. Eng., C 46 (2015) 270275. [28] K. Takechi, M.A. Lieberman, J. Appl. Phys. 89 (2001) 53185321.
[18] H.S. Salapare III, G.Q. Blantocas, V.R. Noguera, H.J. Ramos, in: K.L. Mittal (Ed.), [29] Y.Y. Yan, N. Gao, W. Barthlott, Adv. Colloid Interface Sci. 169 (2011)
Contact Angle, Wettability, and Adhesion, vol. 6, VSP/Brill, Leiden, 2009, pp. 80105.
207216. [30] R.N. Wenzel, Ind. Eng. Chem. 28 (1936) 988994.
[19] M.I. Bahtiyari, Opt. Laser Technol. 43 (2011) 114118. [31] R.N. Wenzel, J. Phys. Colloid Chem. 53 (1949) 14661467.
[20] A. Vital, M. Vayer, C. Sinturel, T. Tillocher, P. Lefaucheux, R. Dussart, Appl. Surf.
Sci. 332 (2015) 237246.

You might also like