You are on page 1of 10

Applied Surface Science 475 (2019) 896–905

Contents lists available at ScienceDirect

Applied Surface Science


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

Full Length Article

Laser texturing of stainless steel under different processing atmospheres: T


From superhydrophilic to superhydrophobic surfaces
P. Poua, , J. del Vala,c, A. Riveiroa, R. Comesañab, F. Arias-Gonzáleza, F. Lusquiñosa,

M. Bountinguizaa, F. Quinteroa, J. Poua,c


a
Applied Physics Dept., University of Vigo, E.E.I., Lagoas-Marcosende, Vigo 36310, Spain
b
Materials Engineering, Applied Mechanics and Construction Dept., University of Vigo, E.E.I., Lagoas-Marcosende, Vigo 36310, Spain
c
Mechanical Engineering Dept., Columbia University, New York, NY 10027, USA

ARTICLE INFO ABSTRACT

Keywords: Wettability plays a major role in a variety of surface related phenomena, as corrosion, heat transfer or tissue adhesion
Laser texturing on implants. Consequently, great research effort is being devoted to control the wetting degree of functional surfaces.
Wettability Pulsed laser texturing at micro/nanometric level has been widely used for that purpose as a precision/time efficient
Stainless steel technique. This work studies the role of the processing atmosphere in controlling the wettability of commercial AISI
Processing atmosphere
304 by laser texturing. A pulsed laser source (λ = 532 nm) working at the nanosecond regime was employed and five
Superhydrophilic
Superhydrophobic
different atmospheres were tested (i.e. O2, Air, CO2, N2, Ar). The results show clear differences in the wetting behaviour
depending solely on the processing environment, ranging from hydrophilicity (31°) to hydrophobicity (125°). Those
differences in wettability were found to be a consequence of changes in surface chemistry between samples processed
under the various gases. The laser processing parameters showed a capability to tune the final wetting behaviour by
controlling the topography and modulating the chemical composition given by the processing environment. It is de-
monstrated how the effects of the atmosphere can be exploited to tailor the wettability of the untreated surfaces
(θ = 88°) up to the desired value, ranging from superhydrophilicity (θ = 0°) to superhydrophobicity (θ = 152°).

1. Introduction laser pulses, has proved to be a useful tool to modify the surface to-
pography at micro/nano scale over a wide area of material within a
Wettability is a crucial surface property governing diverse phenomena relatively short time, which makes it a precision/time efficient tech-
like corrosion [1], cell adhesion [2,3] or heat transfer [4], among others. nique. The lack of special requirements like vacuum environment or
Then, the control of the wettability of functional surfaces has a tremendous chemical coatings, together with its adaptability to different materials
potential benefit in the performance of engineering devices. For that reason, have already foster its use to tune the wettability of metals [13],
the fabrication of surfaces with extremely low wettability, designated as polymers [14], ceramics [15] or natural stones [16].
superhydrophobic, or with extremely high wettability, also known as su- Regarding metals, wettability modification by pulsed laser texturing
perhydrophilic, has been actively researched [5,6]. Such surfaces make entails topography changes based on melting/evaporation processes, which
possible the development of water harvesting systems, controlled drug re- can be controlled by the processing parameters to tune the size and period
lease devices or low drag surfaces for maritime transports [7,8]. of the generated structures [17,18]. Interestingly, a wettability transition
The wetting behaviour of a surface is governed by the interactions be- from highly hydrophilic freshly processed surfaces towards hydrophobic
tween the solid phases and the liquid at molecular level. Changes in the values as time is elapsed from processing has been reported and studied in
chemical composition of the surface and/or in the surface roughness modify many publications to this day, irrespective of the irradiated metal
its intrinsic wettability degree, as described by the classic models of Young [13,19–24]. Despite the mechanism is not totally clear, generally it is ac-
[9], Wenzel [10] and Cassie-Baxter [11,12]. On that basis, methods for cepted that high energy metal oxides are generated during laser processing,
modifying the wettability require the control of both surface chemistry and giving its initially hydrophilic behaviour to the surface. Those oxides are
topography, being plasma etching, sol-gel modification, or hydrothermal highly active and tend to react with surrounding molecules over time, de-
reaction some examples of the most used ones [7]. creasing the surface energy and the wettability in consequence [20,22].
Laser texturing based on short/ultrashort pulses, i.e. with ns/ps-fs Most of the current works have only explored tuning the wettability


Corresponding author at: University of Vigo, Applied Physics Dept., E.E.I., Lagoas-Marcosende, Vigo 36310, Spain.
E-mail address: ppou@uvigo.es (P. Pou).

https://doi.org/10.1016/j.apsusc.2018.12.248
Received 20 September 2018; Received in revised form 23 November 2018; Accepted 25 December 2018
Available online 26 December 2018
0169-4332/ © 2018 Elsevier B.V. All rights reserved.
P. Pou et al. Applied Surface Science 475 (2019) 896–905

by laser texturing under ambient conditions, i.e. with air as processing 532 nm in pulsed regime with a pulse width of 14 ns was used in this
atmosphere. Guan and co-workers showed how the processing atmo- work. The laser beam (TEM00 mode and quality factor M2 < 1.5) was
sphere can accelerate or slow down the wettability transition as a way guided with an XY scanning head equipped with two galvanometric
to extend the durability of superhydrophilic surfaces [25]. However, an mirrors and focused with a F-Theta lens (365 mm of focal length) onto
exploration of the hydrophobicity enhancement by laser texturing the surface of the samples, giving a spot size around 130 μm. Fig. 1
based on different processing atmospheres has not been addressed yet includes a schematic of the experimental setup.
to the best of the authors’ knowledge. The selected strategy for laser processing consisted in parallel
In this work, the role of the processing atmosphere in the wettability scanning lines (see Fig. 1). This strategy enables a fast processing of
of laser-textured commercially available AISI 304 stainless steel sur- large areas of material as well as it allows the study of the
faces was investigated. It is demonstrated how the effects of different influence of the different laser parameters on the processed surfaces in a
environments can be exploited to create either superhydrophilic and relatively easy way. Areas of 8 mm × 8 mm were processed for each
superhydrophobic surfaces using a nanosecond laser as single tool in a sample.
one-step method, without further coatings. In the present work, the role of the processing atmosphere in the
wettability of the steel samples was investigated. Experiments were
2. Materials and methods performed under five atmospheres with different degree of reactivity
with the material: oxygen, air, carbon dioxide, nitrogen and argon, all
2.1. Materials preparation of them of 99.5% minimum purity. A total volumetric flow rate of 90 l/
min was supplied to the processing area through two lateral gas nozzles,
AISI 304 stainless steel was used as base material in this study. Plates of each one with cross section of 2 × 40 mm2 (see Fig. 1). Besides, the
50 × 12 × 1 mm3 were prepared by mechanical procedures and rinsed capability of the laser processing parameters to modulate the effects
with ethanol prior to the laser processing. After the texturing process, the from the atmosphere was also evaluated, studying the role of the
samples were rinsed again with ethanol and blown by a stream of dry air to average laser irradiance I (W/cm2), the scanning speed v (mm/s), the
remove any debris that could affect the subsequent measurements. Samples pulse frequency f (kHz) and the scanning distance between parallel
were kept in sterile individual plastic bags filled with air during 7 days after lines sD (mm).
laser processing to allow a reasonable level of chemical stabilization [19]. The experimental strategy of the study comprised two main steps.
At that time, chemistry and contact angle characterizations were performed. Initially, the effects of the processing atmosphere in wettability, to-
pography and surface chemistry were investigated by texturing
samples under the five environments listed above for identical laser
2.2. Laser processing processing parameters. This step revealed the ability of the proces-
sing atmosphere of toggling the resulting wetting degree between
A Nd:YVO4 diode end-pumped laser emitting at a wavelength of hydrophilic and hydrophobic values. A second set of tests was then
performed with the aim of studying the capability of the laser pro-
cessing parameters to modulate the final wettability in both a re-
active and an inert atmosphere, i.e. in air and argon respectively. The
role of each laser processing parameter was screened with a 24 fac-
torial design in each atmosphere. The effects of the individual
parameters and the interactions between them were evaluated. The
significance of those effects was studied by an ANOVA procedure
with a confidence level of 95%. After that, each parameter was stu-
died independently within a redefined study range to understand the
relations between them and the resulting wetting behaviour. With
that knowledge, the laser processing parameters could be adjusted to
exploit the effect of the atmosphere and tailor the wettability of the
samples up to extreme levels: superhydrophobic and super-
hydrophilic. The atmospheres and processing parameters levels for
both steps are collected in Table 1.

2.3. Sample characterization

Surface topography was evaluated quantitatively via the ar-


ithmetic mean deviation of the roughness profile Ra (μm) according

Fig. 1. Schematic of experimental setup and processing strategy.

Table 1
Levels of the studied processing parameters for each experimental step.
1ST Step: Processing atmospheres 2ND Step: Laser parameters
Factorial 24 Individual study

Laser processing parameter Level Low level High level Low level High level

4 2
Average irradiance I (×10 W/cm ) 2.4 0.9 2.6 1.7 4.0
Scanning speed v (mm/s) 25 10 25 25 150
Pulse frequency f (kHz) 20 20 25 20 50
Scanning distance sD (mm) 0.1 0.05 0.1 0.05 0.2
Processing atmospheres Air, Ar, N2, CO2, O2 Air, Ar

897
P. Pou et al. Applied Surface Science 475 (2019) 896–905

to ISO 4287:1997. A roughness gauge TESA Rugosurf 10 G


from TESA technology was employed for the measurements,
performed perpendicularly to the laser scanning lines. Selected
samples were analysed with higher accuracy focus variation optical
profilometry using an Alicona IF device. The average of three in-
dependent measurements as well as the standard deviation were re-
gistered for each sample. Besides, a qualitative topography analysis
was performed via scanning electron microscopy (SEM) using the
secondary electrons detector (SE) from an XL-30 microscope from
Phillips.
Surface chemistry was evaluated semi-quantitatively via
energy-dispersive X-ray spectroscopy (EDS) technique using the SEM
X-ray detector. X-ray photoelectron spectroscopy (XPS) analyses
were performed to confirm quantitatively the elemental composition
results from EDS and to assess the chemical state of the
elements of interest. A Thermo Scientific K-Alpha ESCA instrument
was employed with Al Kα X-ray source at 1486.6 eV. Collection of Fig. 2. Wettability comparison for different processing atmospheres. (a) Base
photoelectrons was performed at a take-off angle of 90° from the Material with θ = 88 ± 2°. (b) Oxygen with θ = 31 ± 3°. (c) Air with
sample in constant analyser energy mode (CAE). Survey spectra were θ = 46 ± 7°. (d) Carbon dioxide with θ = 50 ± 3°. (e) Nitrogen with
collected with a 100 eV pass energy to assess quantitatively the ele- θ = 83 ± 9°. (f) Argon with θ = 125 ± 2°. Constant processing parameters:
I = 2.4 × 104 W/cm2, v = 25 mm/s, f = 20 kHz, sD = 0.1 mm.
mental composition of the samples. The chemical state of the ele-
ments of interest was studied by peak deconvolution from high re-
solution spectra at 20 eV of pass energy. The samples were analysed
in two different conditions. Initially, the samples were analysed in As it can be seen in Fig. 2a, the base material shows a contact
“as-received” condition, i.e. after processing and stabilization in air. angle with water close to 90°. This means that the non-treated surface
In order to improve the signal to noise ratio for the metal species and does not have a strong hydrophobic or hydrophilic behaviour. How-
evaluate atomic ratios for elements of interest, the samples were also ever, it is clear from Fig. 2b–d that samples processed under oxygen,
analysed after 45 s of sputter cleaning process with argon ions, air and carbon dioxide atmospheres become clearly hydrophilic.
partially removing the adventitious carbon contamination. The When the sample is processed in nitrogen this effect is weaker, as
survey spectra from this last condition were finally selected. shown by Fig. 2e. On the contrary, using argon as processing atmo-
The high resolution spectra were selected from the “as-received” sphere results in a hydrophobic surface, as displayed by Fig. 2f. Then,
condition, in order to avoid species reduction by argon ions. The C 1s there are clear evidences of an influence of the processing atmosphere
hydrocarbon photo peak at 285.0 eV was set as lower binding energy on the wettability of the textured surfaces. The following sections
reference. discuss the results of the characterization of the main surface prop-
Surface wettability was evaluated with an electronic goniometer erties that may explain those wetting tendencies: topography and
from Ossila Ltd. Contact angle θ (°) measurements were performed surface chemistry.
based on the sessile drop method. 2 μl drops of Mili-Q distilled water
were deposited on the samples with a micropipette, following the 3.1.2. Topography and roughness
recommendations given by the European standard EN 828:2013. Fig. 3 includes the SEM pictures of the base material and a groove
Contact angle hysteresis CAH (°) was characterized by the difference obtained by processing under identical laser parameters on each pro-
between the advancing and receding angles of an expanding and cessing atmosphere. The grooves are obtained by evacuation of molten
shrinking drop up to 6 μl. Stability of θ values for superhydrophobic material along the scanning line under the so-called recoil pressure.
surfaces was confirmed by measurements after 150 days elapsed from This mechanism is discussed further in Section 3.2.2. As it can be seen,
processing. Averages and standard deviation of three independent the resulting topography is very similar between the different atmo-
measurements were registered for each sample and characterized spheres. Roughness measurements with the roughness gauge revealed
angle. only small differences among the atmospheres. That result was con-
firmed by optical profilometry values, included in the figure, suggesting
3. Results and discussion that the previously observed wettability changes are not consequence of
roughness variations.
3.1. Laser texturing under different atmospheres Higher magnification pictures of the centre of the grooves pro-
cessed in oxygen and argon are shown in Fig. 4. There is a clear
3.1.1. Wettability contrast between a surface with cracks for oxygen processing and a
Stainless steel samples were processed under the five different smooth groove for argon processing. These cracks may result from the
above mentioned atmospheres (see Table 1) under identical laser pro- formation of an oxide layer due to the high temperature processes
cessing parameters. Fig. 2 shows the corresponding wettability mea- that take place during texturing. Besides, the groove processed in
surements of the textured surfaces together with the base material, oxygen shows also periodic sub-micro ripples that do not appear in
shown as reference. the case of argon atmosphere. It is known that the presence of oxygen

898
P. Pou et al. Applied Surface Science 475 (2019) 896–905

Fig. 3. Topography and roughness comparison for


different processing atmospheres. (a) Base Material
with Ra = 147 ± 5 nm. (b) Oxygen with
Ra = 800 ± 20 nm. (c) Air with Ra = 730
± 20 nm. (d) Carbon dioxide with Ra = 920
± 20 nm. (e) Nitrogen with Ra = 930 ± 40 nm.
(f) Argon with Ra = 1000 ± 20 nm. Constant pro-
cessing parameters: I = 2.4 × 104 W/cm2,
v = 25 mm/s, f = 20 kHz, sD = 0.1 mm. Scale bar
in all images is 50 μm.

decreases the surface tension of the molten pool [26]. As the oxygen
incorporation is more likely to happen at the higher temperatures of
the central area of the spot, a surface tension gradient can be estab-
lished. This gradient, in combination with the recoil pressure and/or
the surface tension gradient created by the temperature difference
itself, can eventually create a ripple by accumulation of material on
the edges of the spot [27]. Then, the micro/sub-micro topography
reveals some evidences of possible chemical differences between the
different processed surfaces.

3.1.3. Surface chemistry


Fig. 5 includes the relative atomic composition of the main elements
found in EDS analyses of the grooves processed under different atmo-
spheres. As it can be seen, there are no significant differences regarding
the relative amounts of the main components of the alloy, i.e. nickel
and chromium. However, substantial differences appear in the relative
oxygen quantity when processing in carbon dioxide, air and especially
in oxygen atmosphere. Furthermore, a relative increase in nitrogen is
also detected when processing in nitrogen or air atmospheres.
Those composition differences were quantitatively studied by XPS
measurements (refer to supplementary material S1 for XPS spectra and
data). Table 2 includes the values of the relative atomic amounts of
oxygen and nitrogen for all atmospheres obtained from the survey
spectra after the sputter cleaning process. The values overall confirm
the tendencies of EDS results. In this case there is no detection of ni-
Fig. 4. Higher magnification groove topography in oxygen (a) and argon (b). trogen in the sample processed in air. That can be explained by a
Constant processing parameters: I = 2.4 × 104 W/cm2, v = 25 mm/s, smaller concentration of nitrogen in the outer surface, as XPS analysis
f = 20 kHz, sD = 0.1 mm. collects signal from the first 10 nm of surface, whereas in EDS the signal
comes from depths up to 1 μm [28].
The chemical state of those elements was further analysed by high

899
P. Pou et al. Applied Surface Science 475 (2019) 896–905

Fig. 5. Relative element atomic amounts comparison for different processing atmospheres from EDS analyses. Constant processing parameters: I = 2.4 × 104 W/cm2,
v = 25 mm/s, f = 20 kHz, sD = 0.1 mm.

Table 2
Relative oxygen and nitrogen atomic amounts for different processing at-
mospheres from XPS survey spectra in sputter cleaned condition. Constant
processing parameters: I = 2.4 × 104 W/cm2, v = 25 mm/s, f = 20 kHz,
sD = 0.1 mm.
Sample O/Fe N/Fe

Base Mat. 0.44 –


O2 2.05 –
Air 1.81 –
CO2 1.72 –
N2 1.00 0.40
Ar 1.38 –

resolution XPS. Table 3 collects the total amounts of oxygen, nitrogen


and carbon in atomic percentage as well as the partial contributions of
the different chemical states obtained from the deconvolution of the
main spectral peaks in “as-received” condition.
As it can be seen, metal lattice oxides and hydroxides (–OH) are
responsible of nearly all the oxygen detected. The formation of those Fig. 6. Relation between the contact angle and the sum of the relative atomic
amounts of oxygen and nitrogen from XPS analyses in sputter cleaned condi-
oxides and hydroxides was confirmed by deconvolution of iron and
tion, for samples processed in the different atmospheres. Constant processing
chromium peaks (refer to supplementary material S1). Metal oxides are
parameters: I = 2.4 × 104 W/cm2, v = 25 mm/s, f = 20 kHz, sD = 0.1 mm.
the result of the reaction of metal atoms with oxygen at high tem-
peratures during processing. However, the high reactivity of oxide ions
in these oxides, acting as strong Lewis bases, favours its reaction with decreased after processing. This reduction is mainly due to the removal
ambient water molecules and the formation of hydroxides [20,29,30]. of polar groups (CeO, C]O, and COO−), whereas non polar carbon
This is likely to have happened during the 7 days elapsed between from hydrocarbon chains (CeC(H)) was rather recovered between
processing and XPS measurements. Nitrogen is found to be part of metal processing and measuring.
nitrides in the sample processed in nitrogen atmosphere. The base The above results are clear evidence of the formation of metal
material shows also a peak of nitrogen, probably arising from surface oxides and nitrides during processing and its dependence on the pro-
contamination, which is effectively removed by the treatment as shown. cessing atmosphere. Transition metals like iron, chromium and nickel
The presence of carbon is also likely to be the result of the adsorption of of austenitic stainless steels form oxides and nitrides with polar beha-
organic contaminants from the atmosphere, reacting with active metal, viour, due to the tendency of metal atoms to release the bonding
oxide and hydroxide sites between processing and XPS analyses electrons. In other words, the presence of oxides and nitrides can in-
[20,22,31]. The total carbon amount present in the base material is crease the polar component of the surface energy. Then, the joint

Table 3
Total element amounts in atomic percentage and partial chemical state contributions for oxygen, nitrogen and carbon for base material and different processing
atmospheres from XPS analyses in “as-received” condition. Constant processing parameters: I = 2.4 × 104 W/cm2, v = 25 mm/s, f = 20 kHz, sD = 0.1 mm.
Element O N C


Chemical state Total O Lattice Oxides eOH H2O /eCOO Total N Lattice nitrides C-NH2 Total C CeC(H) CeO C]O COO−

Base Mat. 33.62 8.00 23.21 2.41 2.78 – 2.78 52.94 34.61 8.30 1.65 8.38
O2 43.56 24.09 17.79 1.68 – – – 36.68 31.47 2.21 – 3.00
Air 43.39 27.15 14.71 1.53 – – – 36.96 30.94 2.75 – 3.27
CO2 41.83 23.91 16.57 1.35 – – – 39.50 33.22 2.08 – 4.20
N2 32.51 16.57 15.25 0.69 8.04 8.04 – 43.34 37.04 2.81 – 3.49
Ar 40.48 29.28 10.29 0.91 – – – 37.78 32.14 2.55 – 3.09

900
P. Pou et al. Applied Surface Science 475 (2019) 896–905

amount of nitrogen and oxygen, as a measure of the total reactivity of


the surface with the environment while processing, gives a rough es-
timation of the increase in the polar component of the surface energy.
This polarity is likely to enhance the tendency of the surfaces to interact
with polar water molecules [32]. In Fig. 6, the water contact angles
from the samples processed under the different processing atmospheres
are plotted against the sum of the relative amounts of oxygen and ni-
trogen from XPS survey spectra after the sputter cleaning process. It can
be seen a clear correlation between the increase in reactivity or polarity
and the increase in the wettability of the processed surface.
At this point, with the objective of obtaining superhydrophobic and
superhydrophilic surfaces, a selection of suitable processing atmo-
spheres was performed. Based on the above findings, to induce a hy-
drophilic behaviour, either air, carbon dioxide or oxygen could be
studied. However, due to its low cost and thereby easier application, the
air atmosphere was selected. On the other hand, the argon atmosphere
was selected as an inert atmosphere to create hydrophobic surfaces. The
following sections show how the laser processing parameters can be
tuned to take further advantage of the effects of the processing atmo-
sphere and tailor the wettability of the samples.

3.2. Tailoring the wettability with laser processing parameters

3.2.1. Screening with factorial design


In each one of the selected atmospheres (i.e. air and argon) a set of
experiments based on a factorial design was carried out as the first
approach to the identification of the effects of the laser processing
parameters on the wettability of the samples. The processing para-
meters limits for this set of experiments were listed in Table 1. Table 4
collects the results from the ANOVA procedure. The first column
identifies the individual parameter factors as well as the ones arising
from the interaction of two parameters. At a confidence level of 95%,
only factors with p-values below 0.05 have a statistically significant
effect.
As it can be seen, for air atmosphere only the average irradiance I
shows statistical significance, whereas in the case of argon atmosphere
both the irradiance and the frequency f have statistical significance. The
effect of the irradiance is by far the greatest one regarding its mean
square value.
Fig. 7 shows the contour plot for the contact angle versus irradiance
and frequency for both atmospheres. In the case of air processing in Fig. 7. Contour plot for contact angle θ (°) against average irradiance I and
Fig. 7a, the high relative importance of the irradiance is revealed by pulse frequency f for air atmosphere (a) and argon atmosphere (b). Base ma-
contour lines almost perpendicular to the horizontal axis. The decrease terial wettability: θ = 88 ± 2°.
of the contact angle as the irradiance increases, meaning an increase of
wettability, is also evident. On the contrary, for argon atmosphere in seen how decreasing the frequency implies a rise of the contact angle
Fig. 7b, a rise in the irradiance results in an increase of the contact values.
angle, which implies a reduction of the wettability. Besides, it can be These results indicate that the laser processing parameters, espe-
cially the irradiance, can tune the wettability behaviour in the regime
Table 4
dictated by each atmosphere (i.e. hydrophilic in air and hydrophobic in
ANOVA table for screening of laser processing parameters effects on wettability.
argon). In the following sections it is discussed how the surface prop-
Significant effects at 95% confidence level are labelled by (*).
erties that determine the observed wetting response can be controlled
Air Argon by the treatment.
Factor Mean Sq p-value Mean Sq p-value

−7* 3.2.2. Controlling the surface properties


I 45,679 4.72 × 10 4076.5 6.59 × 10−4*
v 3 0.7991 21.5 0.6088 To understand how the laser processing parameters are able to tune
f 137 0.1299 723.7 0.0249* the wettability, it is first required to identify the texturing mechanism
sD 99 0.1840 22.3 0.6025 that generates the final topography. On this matter, Fig. 8 shows the
I:v 48 0.3340 219.1 0.1418
plan view and cross section pictures from SEM of both the base material
I:f 117 0.1542 378.4 0.0706
I:sD 95 0.1911 55.5 0.4206
and a representative textured surface.
v:f 114 0.1588 66.4 0.3814 It is clear that the topography of the base material in Fig. 8a,b is
v:sD 36 0.3975 280.3 0.1057 altered by the generation of a groove pattern after laser texturing, as
f:sD 32 0.4219 6.9 0.7703 shown in Fig. 8c. At the cross section shown in Fig. 8d, a valley can be
identified at the centre of the irradiated spot and some resolidified

901
P. Pou et al. Applied Surface Science 475 (2019) 896–905

Fig. 8. Comparison of plan view and cross section of base material surface (a,b) and processed surface (c,d). Processing parameters: I = 4.0 × 104 W/cm2,
v = 10 mm/s, f = 20 kHz, sD = 0.1 mm, Atm: Air. Scale bar in all images is 50 μm.

material is found at both sides. This topography seems to be the result this enhancement on melt ejection leads to an increase of the roughness
of the ejection of molten material under the so-called recoil pressure of the samples. It is also displayed that, for air processing, the rise in
generated by the expanding metal vapour [22,33], produced at peak roughness is followed by an increase of the wettability.
irradiances higher than 107 W/cm2 for stainless steel under nanosecond In a similar trend, the remaining laser processing parameters also
pulses [34]. Note that the minimum peak irradiance used in this work is showed a capability to tune the resulting topography and roughness
3.2 × 107 W/cm2. In Fig. 8c it can be seen how each groove is itself levels, and the wettability in consequence, when the study ranges were
constructed by a periodic micro-ripple pattern along the track. An ex- widened from the ones initially selected in the factorial analysis (refer
planation for this morphology could be the periodic evacuation of to supplementary material S2 for further details). As summary and in
molten material away from the laser-interaction region, in a cycle in- accordance to related studies [22,35,36], reducing the scanning speed v
volving heating, melting, vaporization and ejection. Once the material increases the total energy per area, and generates more pronounced
is moved apart from the centre of the spot, it is solidified and the valleys in consequence. Also, reducing the pulse frequency f leads to
process would start again (refer to supplementary material S2 for fur- higher individual pulse energy, enhancing the melt ejection and the
ther discussion in this regard). topography modification. As well, when the scanning distance sD is
When the average irradiance is increased, the higher amount of reduced, the grooves are closer to each other, rising the average
energy per unit area increases the amount of molten material removed roughness of the surface.
from the centre of the irradiated spot, creating deeper and wider valleys According to the discussion in Section 3.1.3, the wetting regime (i.e.
as well as more prominent ridges, as shown by Fig. 9a and b and ac- hydrophilic or hydrophobic) is dictated by the surface chemistry, highly
cording to findings in literature [22,35,36]. In Fig. 9c it is shown how dependent on the reactivity of the selected atmosphere with the surface

Fig. 9. Cross section pictures of samples pro-


cessed under I = 1.9 × 104 W/cm2 (a) and
I = 3.0 × 104 W/cm2 (b) and evolution with ir-
radiance of roughness and wettability in air at-
mosphere (c). Constant processing parameters:
v = 10 mm/s for (a,b), v = 25 mm/s for (c),
f = 20 kHz, sD = 0.1 mm, Atm: Air. Scale bar in
images (a,b) is 50 μm.

902
P. Pou et al. Applied Surface Science 475 (2019) 896–905

reaction of oxygen and nitrogen molecules with the sample leads to the
formation of polar compounds, as indicated by XPS results. Therefore,
the resulting surface energy is high, attracting the water drop, which
presumably fills the grooves generated by laser texturing. When more
average energy per area is delivered in the treatment, for example by a
rise in the average irradiance or by a decrease in the scanning speed,
both the surface roughness and the surface energy increase, enhancing
the hydrophilic behaviour. The increase of the wettability with the
roughness seems to agree, at least qualitatively, with the model of
Wenzel [10], as a hydrophilic surface becomes more hydrophilic when
roughness is added. That is a consequence of the increase in the width
and depth of the grooves seen by the water drop, favouring its expan-
sion over the surface and increasing the total contact area [38]. In
addition to that, the increase of surface polarity for higher levels of
delivered energy helps to that enhancement of the hydrophilic beha-
viour. On the other hand, with an inert atmosphere like argon, an
averagely hydrophobic chemistry may result from the stabilization of
Fig. 10. Sum of the relative atomic amounts of oxygen and nitrogen from EDS hydrocarbons after processing, as indicated by XPS measurements. The
analyses, for base material and samples processed in increasing irradiance in air Wenzel’s model can explain again the observed tendencies, as the hy-
and argon atmospheres. Constant processing parameters: v = 25 mm/s, drophobic behaviour would be again enhanced with roughness. Also,
f = 20 kHz, sD = 0.1 mm. the lack of polar molecules and/or the same presence of hydrocarbons
might prevent the water drop from filling the grooves [20], enabling
the air entrapment state modelled by Cassie and Baxter [11,12] and
characteristic of low wettability surfaces.
The findings reported up to this point evidence the capability of the
presented method to tune the final wettability of the samples between
highly hydrophilic and highly hydrophobic levels. The following sec-
tion shows the feasibility of exploiting this result up to extreme wett-
ability regimes.

3.3. Superhydrophobic and superhydrophilic surfaces

The laser processing parameters were tuned taking into account the
previous results in order to maximize the hydrophilic and the hydro-
phobic behaviour respectively. According to the found relations, rela-
tively high values of irradiance and low values of scanning speed, pulse
frequency and scanning distance were selected in both cases. The re-
sults are shown in Fig. 12.
As it can be seen, the average wetting behaviour of the base material
Fig. 11. Wettability evolution in air and argon atmospheres. Processing para- (Fig. 12a, θ = 88 ± 2°) was successfully modified by means of laser
meters: I = 2.5 × 104 W/cm2, v = 25–150 mm/s, f = 20 kHz, sD = 0.1 mm. texturing. In air atmosphere, a superhydrophilic surface (θ = 0 ± 2°)
was obtained (Fig. 12b). The water drop completely extends on the top
of this surface, filling the groves generated by the laser treatment. In
argon atmosphere, a superhydrophobic surface (θ = 152 ± 4°) was
while processing. At the same time, the processing parameters also have
obtained (Fig. 12c). The water drop remains nearly spherical due to the
an effect on surface chemistry. On this matter, Fig. 10 shows the sum of
low tendency to extend on the surface. The stability of this wetting
the relative amounts of oxygen and nitrogen from EDS analyses of the
behaviour was confirmed by measurements after 150 days elapsed from
base material and the grooves processed under air and argon for in-
processing, with average values of θ = 154 ± 4°. The super-
creasing irradiance.
hydrophobic property was also characterized by a small contact angle
It can be observed that, when processing in air, an increase in the
irradiance produces more oxygen and nitrogen incorporations on the
sample surface, in accordance to findings of similar studies in literature
[37]. Higher amounts of average energy per unit area lead to higher
temperatures, increasing the reactivity of the material with the en-
vironment and explaining this result. In the case of argon processing,
the sum of the relative amounts of oxygen and nitrogen does not show
significant variations with the irradiance, staying approximately at the
levels of the unmodified surface.
Therefore, from the previous results it is clear that the laser para-
meters have an effect on both the topography and the surface chemistry
of the samples. Fig. 11 summarizes the repercussion of those effects on
the resulting wettability of the samples. Fig. 12. Comparison of the wettability of (a) base material with θ = 88 ± 2°,
It can be seen how the roughness evolution, responding to changes (b) superhydrophilic surface with θ = 0 ± 2° and (c) superhydrophobic sur-
in the laser processing parameters, is successfully translated into more face with θ = 152 ± 4°. Processing parameters: (b) I = 2.6 × 104 W/cm2,
hydrophilic samples when processing in air and more hydrophobic v = 10 mm/s, f = 20 kHz, sD = 0.05 mm, Atm: Air; (c) I = 2.3 × 104 W/cm2,
samples for argon atmosphere. On the one hand, in air processing, the v = 25 mm/s, f = 20 kHz, sD = 0.025 mm, Atm: Argon.

903
P. Pou et al. Applied Surface Science 475 (2019) 896–905

Fig. 13. Comparison of the impact of a water drop on the base material (a), superhydrophilic surface (b) and superhydrophobic surface (c). Processing parameters:
(b) I = 2.6 × 104 W/cm2, v = 10 mm/s, f = 20 kHz, sD = 0.05 mm, Atm: Air; (c) I = 2.3 × 104 W/cm2, v = 25 mm/s, f = 20 kHz, sD = 0.025 mm, Atm: Argon.

hysteresis of CAH = 7 ± 4° (refer to supplementary material S3 for ranging from 31° to 125°. Whereas similar roughness values and micro
further details about θ stability and CAH measurements). topography were found for all those surfaces, sub-micro features in-
To show how these surfaces perform, the impact of a water drop on dicated possible chemical differences among them. EDS analyses ef-
each one was recorded with a high speed camera. Captures of those fectively showed variations in the relative amounts of oxygen and ni-
recordings are shown in Fig. 13 (refer to supplementary material S4 for trogen. XPS measurements confirmed the appearance of different
video version). Compared to the small expansion of the water drop over amounts of oxides and nitrides for each atmosphere. The measured
the base material surface shown in Fig. 13a, the extension is clearly contact angles for the various atmospheres seem to be related with the
higher in the case of the superhydrophilic surface, in Fig. 13b. On the sum of the relative amounts of oxygen and nitrogen detected. The po-
contrary, the drop can bounce several times on the top of the super- larity of oxides and nitrides of the transition metals of the alloy was
hydrophobic surface due to the low adhesion among them, as displayed proposed as an explanation to the increase of wettability as the joint
by Fig. 13c. oxygen and nitrogen amount increases with the reactivity of the pro-
Overall, the results presented in this work demonstrate the great cessing atmosphere.
influence of the processing atmosphere, in particular of its reactivity In order to take advantage of the capability of the processing at-
with the processed surface, on the wettability of commercial stainless mosphere of toggling the final wettability of the textured surfaces, the
steel sheets after laser texturing. In addition, it is demonstrated how role of different laser processing parameters on the process was first
those effects can be exploited to produce either superhydrophobic or screened by a full factorial design in selected air and argon atmo-
superhydrophilic surfaces. Moreover, the use of a nanosecond laser spheres. The average irradiance demonstrated to be the most significant
source, that, in general terms, is cheaper than a ultrashort pulsed laser factor on the wettability. An increase of the irradiance led to an in-
such as a femtosecond one, together with the absence of costly vacuum crease of the hydrophilic/hydrophobic behaviour in air/argon en-
chambers or additional coatings, reinforce the potential of this tech- vironments respectively.
nique. The capability to produce either superhydrophobic or super- The analysis of the topography showed a groove pattern as result of
hydrophilic surfaces with a single tool can find an application in the the texturing process. The evacuation of molten material by means of
generation of wettability patterns or gradients to enhance pool boiling the evaporation recoil pressure was found to likely be the main topo-
heat transfer [4] or to selectively promote/prevent the adhesion of graphy modification mechanism. Periodic features along the grooves
cells/bacteria on implants and other biomedical devices [2,3,39]. indicated the feasibility of the existence of heating, melting, evapora-
tion and ejection cycles.
The effect of the average irradiance I on wettability was found to be
4. Conclusions related to changes in the amount of ejected material, which controls the
size of the generated grooves and, consequently, the average roughness
The wettability control of AISI 304 stainless steel plates was ex- of the samples. Similar relations could be found for the remaining laser
plored by means of changing the processing atmosphere during laser processing parameters. Besides, an increase of the reactivity of the
texturing. The results clearly demonstrate a major role of the atmo- material with the environment when more energy is delivered was
sphere on the final wetting behaviour of the laser processed surfaces. detected in air atmosphere. Consequently, changes in surface chemistry
The study of the effects of the processing environment was devel- with processing parameters seem to contribute also to the overall
oped by texturing samples into oxygen, air, carbon dioxide, nitrogen wettability tendencies.
and argon atmospheres. Wettability differences were found between Taking advantage of the observed effects of the processing
samples processed under the various atmospheres, with contact angles

904
P. Pou et al. Applied Surface Science 475 (2019) 896–905

atmosphere, both superhydrophilic (θ = 0°) and superhydrophobic of laser surface texturing: an application on Zimbabwe black granites, Appl. Surf.
(θ = 152°) surfaces could be successfully obtained by means of a one- Sci. 418 (2017) 463–471, https://doi.org/10.1016/j.apsusc.2016.12.196.
[17] R. Jagdheesh, J.J. García-Ballesteros, J.L. Ocaña, One-step fabrication of near su-
step texturing process without further coatings, using a nanosecond perhydrophobic aluminum surface by nanosecond laser ablation, Appl. Surf. Sci.
laser as single-tool. 374 (2016) 2–11, https://doi.org/10.1016/j.apsusc.2015.06.104.
[18] S. Moradi, S. Kamal, P. Englezos, S.G. Hatzikiriakos, Femtosecond laser irradiation
of metallic surfaces: effects of laser parameters on superhydrophobicity,
Acknowledgements Nanotechnology. 24 (2013) 415302, , https://doi.org/10.1088/0957-4484/24/41/
415302.
The authors wish to thank the assistance with XPS analyses from Dr. [19] A.M. Kietzig, S.G. Hatzikiriakos, P. Englezos, Patterned superhydrophobic metallic
surfaces, Langmuir 25 (2009) 4821–4827, https://doi.org/10.1021/la8037582.
C. Serra (University of Vigo), and the technical staff of CACTI [20] J. Long, M. Zhong, H. Zhang, P. Fan, Superhydrophilicity to superhydrophobicity
(University of Vigo) and the New Materials Group (University of Vigo) transition of picosecond laser microstructured aluminum in ambient air, J. Colloid
for their help with sample characterization. This work was partially Interface Sci. 441 (2015) 1–9, https://doi.org/10.1016/j.jcis.2014.11.015.
[21] D.V. Ta, A. Dunn, T.J. Wasley, R.W. Kay, J. Stringer, P.J. Smith, C. Connaughton,
supported by the European Union (EAPA_151/2016 Interreg Atlantic
J.D. Shephard, Nanosecond laser textured superhydrophobic metallic surfaces and
Area), Government of Spain (MAT2015-71459-C2-1-P (MINECO/ their chemical sensing applications, Appl. Surf. Sci. 357 (2015) 248–254, https://
FEDER), FPU16/05492, and PRX17/00157), and Xunta de Galicia doi.org/10.1016/j.apsusc.2015.09.027.
(ED431B 2016/042, ED481D 2017/010, ED481B 2016/047-0). [22] J.T. Cardoso, A. Garcia-Girón, J.M. Romano, D. Huerta-Murillo, R. Jagdheesh,
M. Walker, S.S. Dimov, J.L. Ocaña, Influence of ambient conditions on the evolution
of wettability properties of an IR-, ns-laser textured aluminium alloy, RSC Adv. 7
Appendix A. Supplementary data (2017) 39617–39627, https://doi.org/10.1039/C7RA07421B.
[23] C.V. Ngo, D.M. Chun, Fast wettability transition from hydrophilic to super-
hydrophobic laser-textured stainless steel surfaces under low-temperature an-
Supplementary data to this article can be found online at https:// nealing, Appl. Surf. Sci. 409 (2017) 232–240, https://doi.org/10.1016/j.apsusc.
doi.org/10.1016/j.apsusc.2018.12.248. 2017.03.038.
[24] C.V. Ngo, D.M. Chun, Control of laser-ablated aluminum surface wettability to
superhydrophobic or superhydrophilic through simple heat treatment or water
References boiling post-processing, Appl. Surf. Sci. 435 (2018) 974–982, https://doi.org/10.
1016/j.apsusc.2017.11.185.
[1] F. Zhang, S. Chen, L. Dong, Y. Lei, T. Liu, Y. Yin, Preparation of superhydrophobic [25] Y.C. Guan, F.F. Luo, G.C. Lim, M.H. Hong, H.Y. Zheng, B. Qi, Fabrication of metallic
films on titanium as effective corrosion barriers, Appl. Surf. Sci. 257 (2011) surfaces with long-term superhydrophilic property using one-stop laser method,
2587–2591, https://doi.org/10.1016/j.apsusc.2010.10.027. Mater. Des. 78 (2015) 19–24, https://doi.org/10.1016/j.matdes.2015.04.021.
[2] Y. Arima, H. Iwata, Effect of wettability and surface functional groups on protein [26] F.A. Halden, W.D. Kingery, Surface tension at elevated temperatures. II. Effect of C,
adsorption and cell adhesion using well-defined mixed self-assembled monolayers, N, O and S on liquid iron surface tension and interfacial energy with Al2O3, J. Phys.
Biomaterials 28 (2007) 3074–3082, https://doi.org/10.1016/j.biomaterials.2007. Chem. 59 (1955) 557–559, https://doi.org/10.1021/j150528a018.
03.013. [27] T.R. Anthony, H.E. Cline, Surface rippling induced by surface-tension gradients
[3] A. Riveiro, A.L.B. Maçon, J. del Val, R. Comesaña, J. Pou, Laser surface texturing of during laser surface melting and alloying, J. Appl. Phys. 48 (1977) 3888–3894,
polymers for biomedical applications, Front. Phys. 6 (2018) 16, https://doi.org/10. https://doi.org/10.1063/1.324260.
3389/fphy.2018.00016. [28] D. Brandon, W.D. Kaplan, Microstructural Characterization of Materials, John
[4] A.R. Betz, J. Xu, H. Qiu, D. Attinger, Do surfaces with mixed hydrophilic and hy- Wiley & Sons, Ltd, Chichester, UK, 2008. http://dx.doi.org/10.1002/
drophobic areas enhance pool boiling? Appl. Phys. Lett. 97 (2010) 141909, , 9780470727133.
https://doi.org/10.1063/1.3485057. [29] H. Tamura, A. Tanaka, K. ya Mita, R. Furuichi, Surface hydroxyl site densities on
[5] L. Feng, S. Li, Y. Li, H. Li, L. Zhang, J. Zhai, Y. Song, B. Liu, L. Jiang, D. Zhu, Super- metal oxides as a measure for the ion-exchange capacity, J. Colloid Interface Sci.
hydrophobic surfaces: from natural to artificial, Adv. Mater. 14 (2002) 1857–1860, 209 (1999) 225–231, https://doi.org/10.1006/jcis.1998.5877.
https://doi.org/10.1002/adma.200290020. [30] H. Tamura, K. Mita, A. Tanaka, M. Ito, Mechanism of hydroxylation of metal oxide
[6] J. Drelich, E. Chibowski, D.D. Meng, K. Terpilowski, Hydrophilic and super- surfaces, J. Colloid Interface Sci. 243 (2001) 202–207, https://doi.org/10.1006/
hydrophilic surfaces and materials, Soft Matter 7 (2011) 9804–9828, https://doi. jcis.2001.7864.
org/10.1039/c1sm05849e. [31] S. Takeda, M. Fukawa, Y. Hayashi, K. Matsumoto, Surface OH group governing
[7] S. Wang, K. Liu, X. Yao, L. Jiang, Bioinspired surfaces with superwettability: new adsorption properties of metal oxide films, Thin Solid Films 339 (1999) 220–224,
insight on theory, design, and applications, Chem. Rev. 115 (2015) 8230–8293, https://doi.org/10.1016/S0040-6090(98)01152-3.
https://doi.org/10.1021/cr400083y. [32] F.M. Fowkes, Attractive forces at interfaces, Ind. Eng. Chem. 56 (1964) 40–52,
[8] B. Su, Y. Tian, L. Jiang, Bioinspired interfaces with superwettability: from materials https://doi.org/10.1021/ie50660a008.
to chemistry, J. Am. Chem. Soc. 138 (2016) 1727–1748, https://doi.org/10.1021/ [33] M. Stafe, C. Negutu, I.M. Popescu, Theoretical determination of the ablation rate of
jacs.5b12728. metals in multiple-nanosecond laser pulses irradiation regime, Appl. Surf. Sci. 253
[9] T. Young, An essay on the cohesion of fluids, Philos. Trans. R. Soc. London. 95 (2007) 6353–6358, https://doi.org/10.1016/j.apsusc.2007.01.060.
(1805) 65–87, https://doi.org/10.1098/rstl.1805.0005. [34] J.J. Chang, B.E. Warner, E.P. Dragon, M.W. Martinez, Precision micromachining
[10] R.N. Wenzel, Resistance of solid surfaces to wetting by water, Ind. Eng. Chem. 28 with pulsed green lasers, J. Laser Appl. 10 (1998) 285–291, https://doi.org/10.
(1936) 988–994, https://doi.org/10.1021/ie50320a024. 2351/1.521863.
[11] A.B.D. Cassie, Contact angles, Discuss. Faraday Soc. 3 (1948) 11–16, https://doi. [35] V.D. Ta, A. Dunn, T.J. Wasley, J. Li, R.W. Kay, J. Stringer, P.J. Smith, E. Esenturk,
org/10.1039/df9480300011. C. Connaughton, J.D. Shephard, Laser textured superhydrophobic surfaces and their
[12] A.B.D. Cassie, S. Baxter, Wettability of porous surfaces, Trans. Faraday Soc. 40 applications for homogeneous spot deposition, Appl. Surf. Sci. 365 (2016) 153–159,
(1944) 546–551, https://doi.org/10.1039/tf9444000546. https://doi.org/10.1016/j.apsusc.2016.01.019.
[13] A.M. Kietzig, M.N. Mirvakili, S. Kamal, P. Englezos, S.G. Hatzikiriakos, Laser-pat- [36] Y. Cai, W. Chang, X. Luo, A.M.L. Sousa, K.H.A. Lau, Y. Qin, Superhydrophobic
terned super-hydrophobic pure metallic substrates: Cassie to Wenzel wetting tran- structures on 316L stainless steel surfaces machined by nanosecond pulsed laser,
sitions, J. Adhes. Sci. Technol. 25 (2011) 2789–2809, https://doi.org/10.1163/ Precis. Eng. (2018), https://doi.org/10.1016/j.precisioneng.2018.01.004.
016942410X549988. [37] S. Razi, K. Madanipour, M. Mollabashi, Laser surface texturing of 316L stainless
[14] A. Riveiro, R. Soto, R. Comesaña, M. Boutinguiza, J. del Val, F. Quintero, steel in air and water: a method for increasing hydrophilicity via direct creation of
F. Lusquiños, J. Pou, Laser surface modification of PEEK, Appl. Surf. Sci. 258 (2012) microstructures, Opt. Laser Technol. 80 (2016) 237–246, https://doi.org/10.1016/
9437–9442, https://doi.org/10.1016/j.apsusc.2012.01.154. j.optlastec.2015.12.022.
[15] J. Qiao, L. Zhu, W. Yue, Z. Fu, J. Kang, C. Wang, The effect of attributes of micro- [38] D. Quéré, Wetting and roughness, Annu. Rev. Mater. Res. 38 (2008) 71–99, https://
shapes of laser surface texture on the wettability of WC-CrCo metal ceramic coat- doi.org/10.1146/annurev.matsci.38.060407.132434.
ings, Surf. Coatings Technol. 334 (2018) 429–437, https://doi.org/10.1016/j. [39] W. Teughels, N. Assche, I. Sliepen, M. Quirynen, Effect of material characteristics
surfcoat.2017.12.001. and/or surface topography on biofilm development, Clin. Oral Implants Res. 17
[16] A. Chantada, J. Penide, A. Riveiro, J. del Val, F. Quintero, M. Meixus, R. Soto, (Suppl 2) (2006) 68–81, https://doi.org/10.1111/j.1600-0501.2006.01353.x.
F. Lusquiños, J. Pou, Increasing the hydrophobicity degree of stonework by means

905

You might also like