You are on page 1of 35

Accepted Manuscript

Full Length Article

On a simplified method to produce hydrophobic coatings for aeronautical ap-


plications

F. Piscitelli, F. Tescione, L. Mazzola, G. Bruno, M. Lavorgna

PII: S0169-4332(18)31016-X
DOI: https://doi.org/10.1016/j.apsusc.2018.04.062
Reference: APSUSC 39061

To appear in: Applied Surface Science

Received Date: 16 October 2017


Revised Date: 26 March 2018
Accepted Date: 6 April 2018

Please cite this article as: F. Piscitelli, F. Tescione, L. Mazzola, G. Bruno, M. Lavorgna, On a simplified method to
produce hydrophobic coatings for aeronautical applications, Applied Surface Science (2018), doi: https://doi.org/
10.1016/j.apsusc.2018.04.062

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
On a simplified method to produce hydrophobic coatings for

aeronautical applications
F. Piscitelli1*, F. Tescione2, L. Mazzola1, G. Bruno1, M. Lavorgna2

1
CIRA - Italian Aerospace Research Centre, Via Maiorise 1,81043 Capua, Italy

*Corresponding author e-mail: f.piscitelli@cira.it

2
CNR - Institute for Polymers, Composites and Biomaterials, National Research Council, Portici (NA), Italy

Abstract

The deposition of ice on wings and tails surfaces of aircrafts often alters the airflow reducing the

lift force and causing aerodynamic stall. In order to prevent these kind of serious problems, de-icing

and anti-icing systems, which act by removing the ice once it has formed, and by preventing its

formation, respectively, are usually employed. In particular, the anti-icing superhydrophobic and

icephobic coatings can be achieved by designing and realizing nanostructured surfaces by taking

inspiration from effective surface present in nature, e.g., the lotus leaf. However, this process is

often expensive and requires specialized equipment. The effort of the present work is to find out a

simplified and not-expensive method to prepare superhydrophobic coatings to be applied on large

aircraft surfaces. Among the investigated methods, the deposition of several layers of a

homogeneous hydrophobic silica nanoparticles by using a simple aerograph allowed to build-up a

coating which exhibits a contact angle of 154°. The coating withstands to aging tests simulating

flight conditions. Also during typical flight conditions, i.e., -12°C and 0.5 bar, it retains its

1
hierarchical superficial morphology with roughness features from nano- to micro-scale and shows

contact angle higher than the commercial reference.

Keywords: icephobic coating, silica nanoparticles, hierarchical structures, surface free

energy, wettability.

2
1. Introduction

Cold weather conditions are considered to be critical factors in the design, analysis, and

performance tests of the airplanes [1]. Only from 1990 to 2005, there have been 240 reports of

aviation accidents due to icing, and most of them included icing on aircraft wings [2]. Typically, the

presence of tiny pieces of ice or supercooled liquid water in the clouds, which remains liquid below

zero and suddenly turns to ice, are the main sources of ice deposition during a flight. The presence

of ice on surfaces alters the airflow over the wing and tail, and then reduces the lift force that keeps

the plane in the air. This potentially causes aerodynamic stall, a condition that can lead to a

temporary loss of control of the aircraft. Passive methods, such as superhydrophobic and icephobic

coatings, are potentially employed to prevent the ice adhesion surfaces of aircrafts [3]-[10]. The

design of new icephobic coatings is very complex and several thermodynamic conditions have to be

taken into account in order to realize an effective, cost-sustainable and durable material. Alongside

with pressure and temperature the nature of clouds and so of the supercooled water droplets, change

with the altitude. In addition, the airflow around the aeronautical components involves the dynamic

effect of impingement of the droplets on the surfaces. Hence, not only the static pressure is present

but also the dynamic pressure, which affects the behavior of droplets on the surface [11]-[15]. In

this contest, the main function of the icephobic coatings is to reduce the adhesive forces between

the supercooled water or ice, and the surface of the aeronautic component. The recent findings on

the superhydrophobic and icephobic surfaces are inspired to nature, and specifically to the surface

of lotus leaves, which as effect of a combination of hierarchical micro and nanostructures exhibit a

low surface energy. Thus, the fabrication of synthetic superhydrophobic surfaces with

morphological topography similar to lotus leaf typically involves the nanostructuring of surface by

using nanoparticles, photolithography, mesoporous polymers, and surface etching, sometimes in

conjunction with chemical modification to reduce surface energy [16].


3
However, these coatings may be realized only through complex synthetic conditions, e.g., etching

and high temperature, [17]-[18], complex fabrication techniques [19]-[20], and are often limited by

the substrate type and geometry that can be successfully coated [21]. In more details, a top-down

approach, which includes the wet chemical etching [22]-[24], the dry etching [25], and the laser

irradiation [26]-[27] directly generates surface roughness on a material, thus the final micro/nano

textures are an integral part of the substrate, with no boundary or interfacial adhesion issues [28], so

enhancing the durability of coating. The top-down methods are typically substrate-dependent,

which limits their wider applications. In contrast, other bottom-up methods, e.g., the spray coating

[29], the electrospinning [30], the spin-coating [31]-[32], and the dip-coating [33], are generally

substrate independent. However, they typically exhibit mechanical durability issues, such as the

possibility of delamination, and thus grafting and/or crosslinking methods are generally employed

to enhance their mechanical durability [34]-[35]. Therefore, these approaches are often expensive

and/or require specialized equipment. Among bottom-up procedures, interesting results have been

obtained by using silica nanoparticles deposited by layer-by-layer methods. Bravo et al. [20]

controlled the placement and level of aggregation of differently sized nanoparticles within the

multilayer film, tuning the surface roughness and achieving the superhydrophobic behavior.

Transparent superhydrophobic films were created by the sequential deposition of silica

nanoparticles and poly(allylamine hydrochloride), and a final treatment with silane. In details, the

process consists in the deposition of adhesion layers, body layers and top layer, each of them

consists of several bilayers of polyelectrolyte and silica, i.e., 5 bilayers for the adhesion layer, from

10 to 40 bilayers for the body layers and 3 bilayers for the top of layers. Some layers are calcinated

at 550°C. Finally, a fluorosilane was deposited on the top coat. A similar layer-by-layer approach

was employed by Yilgor et al. [31]. They prepared the coating by spinning on the substrate some

layers of a hydrophobic silica dispersion. The adhesion between the coating and the substrate was

4
assured by a polymer solution which was deposited before the coating itself. Finally, the coated

surfaces were treated at 50°C to realize compact and durable coatings.

Common factors, which are generally taken into account during the designing of the

superhydrophobic and icephobic coatings are the creation of low surface energy materials and/or

rough surface textures.

To this end and in this scenario, the goal of the present work is to find out a simplified method able

to prepare an effective superhydrophobic coating to be applied on large surfaces, such as the wings

or tails of the aircrafts, starting from commercial hydrophobic silica nanoparticles. Several methods

have been taken into account for both the preparation and the application of the coatings, i.e., the

sonication and the magnetic stirring for the dispersion of silica nanoparticles, and the paintbrush

and the aerograph for their deposition, respectively, using methods and tools usually employed in

the pre-industrial and industrial scale. Starting materials, such as the hydrophobic silica, are chosen

among low-cost and easily available ones. Developed coatings were characterized by static contact

angles at room temperature, and at -12°C and 0.5 bar. Aging tests simulating flight conditions were

performed in order to assess the durability of the coatings. Optical and electron microscopies were

carried out to correlate the hydrophobic behavior to the morphology of coatings.

2. Experimental

2.1. Materials

Hydrophobic silica Aerosil R812 which consists of fumed silica modified with

Hexamethyldisilazane (HDMS), having a specific surface area of 260 m2/g, was provided by

Evonik. The dimension of the primary particle size is about 7 nm, which aggregate in agglomerates

of 100-500 nm [36]. Industrial grade transparent Polyurethane (PU), RS 132633, with a viscosity of

5
2500-3000 cps, was purchased by RS and used as primer layer. Tetrahydrofuran (THF) anhydrous

(≥99.9%) and Isopropyl alcohol (IPA) were purchased by Sigma Aldrich.

2.2. Superhydrophobic coating preparation

Several procedures for the preparation and the application of the superhydrophobic coating were

adopted. The coating was prepared by starting from a dispersion of nanosized hydrophobic silica in

THF with a concentration of 2.5% by weight, obtained by sonication or by magnetic stirring, and

then applied by paintbrush or by aerograph. A sonicator ultrasonic processor manufactured by

Qsonica was used for the dispersion of silica nanoparticles in pulse mode by operating 10 s on and

10 s off, with an amplitude of 40%, for 180 min or for 30 min. Due to the high energy generated, a

cold-water bath was used to avoid overheating. Alternatively, the dispersion was prepared by

stirring the silica and solvent for 60 min at room temperature using a commercial stirrer and a

magnetic anchor. Two kind of supports, i.e., the opaque epoxy-carbon fiber composites painted with

Aerowave 5001 from Akzo Nobel, labelled OP-, and the transparent neat Poly(methyl methacrylate)

(PPMA), denoted as TR-, were used as substrates and reference samples. In particular, PMMA

surface was cleaned with IPA before the coating or primer applications. All samples were prepared

following a multi-step procedure able to build-up a layered structure consisting of the primer and

several nanoparticles layers. In this way, it was possible to realize layered structures as

schematically depicted in Figure 1. For most samples, a primer consisting of PU in THF (5 or

20wt% (see Table 1)) was applied on the substrate before the nanostructured coating, in order to

enhance the adhesion between the hydrophobic silica and the substrate.

6
Figure 1: schematic structure of the coating obtained by the multi-step procedure.

Table 1: key-designing variables for the preparation and the application of the superhydrophobic

coatings.

ID sample Dispersion of silica Primer Application Evaporation of

R812 in THF method THF

(2,5wt%)

OP-DR-n Sonication for 180 min 20wt% PU in THF Paint brush 75°C for 10 min

(0÷15 layers)

OP-S-n Magnetic stirring for 60 20wt% PU in THF Paint brush 75°C for 10 min

min (0÷50 layers)

OP-SiO2 Sonication for 30 min 5wt% PU in THF Aerograph (3 layers) 75°C for 5 min

TR-PR-SiO2 Sonication for 30 min 5wt% PU in THF Aerograph (3 layers) 75°C for 5 min

TR-SiO2 Sonication for 30 min NO PRIMER Aerograph (3 layers) 75°C for 5 min

The concentration of the primer was chosen in order to not affect the transmittance of the

transparent PMMA substrate. The same solvent, i.e., the THF, was employed for both the primer

and silica dispersion. This because, as Yilgor et al. suggested [31], the THF dissolves/swells the
7
surface layer of most polymers, being a good solvent for most of them. In this way, it helps the

silica nanoparticles to partially penetrate into the first layer of polymer substrate, getting firmly

embedded into them. Afterwards, several layers of the silica suspension were deposited. The

procedure and silica layers number change with the samples, as reported in Table 1, which

summarizes all details about the samples preparation. The layer-by-layer approach allowed to

achieving, through a mechanism induced by the solvent evaporation, an intrinsic surface roughness,

which, together with the low surface energy of the selected hydrophobic silica nanoparticles,

induces the superhydrophobic behavior of the surfaces.

2.2. Characterization methods

Contact angle measurements were performed at room temperature by using three different liquids

(i.e., water and formamide as polar liquids and diiodomethane as dispersive liquid) with the aim to

determine the surface free energy (SFE) and its components, adhesion work and wettability. During

tests, 10 drops of 3 µl of each liquid were deposited on the sample surface. The SFE was calculated

according to the well-accepted Owens-Wendt (OW) method [37]-[38]. Owens and Wendt

envisioned the surface energy of a solid as being comprised of two components: a dispersive

component and a polar component. The dispersive component, theoretically, accounts for the Van

der Waals and other non-site specific interactions that a surface establishes with applied liquids.

Whereas theoretically, the polar component accounts for the dipole-dipole, dipole-induced dipole,

hydrogen bonding, and other site-specific interactions [39]-[40]. In this contest, the diiodomethane

has a relatively high overall surface tension of 50.8 mN/m, and due to its molecular symmetry, it

has not any polar component. The formamide, on the contrary, has both component of the SFE,

namely 22.4 mN/m as dispersive and 34.6 mN/m as polar component. Being predominant, the last

one was used as polar liquid. Water, obviously, was used as polar liquid. According to the OW

8
method, the SFE of the solid can be calculated starting from the geometric mean of the dispersive

and polar components of the surface tension of solid and liquid, respectively, as reported in the

equation (1):

(1)

where, θ is the contact angle between the liquid droplet and surface, is the liquid total surface

tension, is the dispersive component of liquid surface tension and is the polar component of

liquid surface tension. The unknown terms in the equation (1) are which is the dispersive

component of the solid SFE and which is the polar component of the solid SFE. The required

value of total SFE of the solid is obtained as sum of the two components, according to the equation

(2):

(2)

Dividing by the equation (1), it can be rewritten as:

(3)

The left-hand side of the equation (3) contains quantities which are measured experimentally (i.e.,

θ) or available in the literature (i.e., , so that a plot of the left-hand side of equation (3) versus

quantity gives a straight line with slope and intercept .

The surface tension of employed liquids and its components are reported in literature [41].

However, these values are referred to standard conditions of ambient temperature and pressure at

9
sea level, which do not replicate the flight conditions at which the ice is formed. In order to simulate

the actual flight conditions, a new instrument, able to reproduce pressures and temperatures until a

flight altitude of about 16000 meters, and contemporary to measure the contact angles, was

designed and developed at CIRA [42]-[43]. According to the standard certification tests [44], in the

icing wind tunnels, the highest probability to have icing phenomena, is 5000 meters, where the

temperature is about -12÷-15 °C and the pressure is about 0.5 bar. It is evident that, in real

conditions, the behavior and dynamic of the interactions between the supercooled liquid droplets

and surfaces are more complex, because the supercooled water droplets impact the aircraft

components at high speeds. The dynamics of interaction, happening in a few milliseconds, are in

non-equilibrium condition and there are several phenomena that can occur, i.e., infiltration due to

the water hammer pressure, instantaneous freezing or rapid rolling of the droplets. Contrary, the

characterization technique developed at CIRA allowed to studying the icephobic surface in static

conditions. This represents the worst condition since the supercooled water droplet has long time to

reach the equilibrium state and then to freeze on the surface. In order to determine how the SFE

changes with temperature and pressure, the equation (3) can be further processed. By using the

Karbanda’s equation, it is possible to rescale the values of surface tension with temperature:

(4)

where represents the surface tension of liquid at 20 whereas represents the critical

temperature of the selected liquids, namely 373.94°C, 376.45°C and 474.42°C for water, formamide

and diiodomethane, respectively. Once rescaled in temperature, it is necessary to rescale these new

values for the new pressure, using the Laplace equation:

10
= (5)

where K is a constant value, function of employed liquids. For the liquids employed in these

experiments, a value of K equal to 2 was used.

Aging tests of coating were performed in an ANGELANTONI CH 2000CHSP according to the

MIL standard [45] aimed at simulating real flight conditions at 16000 meters in altitude of a Multi-

Role Combat Aircraft (MRCA). Figure 2 (A) reports the samples fixed to supports designed

according to the ASTM-D2247 and tilted at 15°. They were contemporarily undergone to thermal

loads and humidity variations according to altitudes ranging between the sea level and 16000 m in

three different stages: cold/dry; warm/humid; and warm/dry (see Figure 2 (B)). In details, the cold-

dry regime simulates a parking stage on the ground in a cold environment at the minimum

temperature of the overall cycle, i.e., -67 °C, and after a take-off phase at the cruise altitude of

16000 m with a rate of 10 m/s, according to the standard. In this phase, the humidity was not

controlled. The warm/humid cycle simulates a descent phase from cruise altitude to the ground

level at 10 m/s, then a parking on the ground at a relative humidity of 95%, performed at the

maximum facility rate and at the temperature of 32 °C. Finally, the warm/dry cycle simulates the

parking on the ground in a hot environment at the temperature of 43 °C and at the minimum

humidity less than 30%, and then a take-off phase at the cruise altitude reached at 10 m/s. In this

phase, the humidity was not controlled. The overall cycle had a duration of 963 minutes and was

repeated for 10 time.

11
(A)

(B)

Figure 2: (A) samples mounted on the test fixture, (B) setting of altitude, temperature and humidity

during aging tests.

After aging tests, measurements of contact angles at room temperature were carried out in order to

highlight possible degradations due to the typical flight conditions simulated during the aging tests.

Optical and electron microscopies were carried out in order to correlate the hydrophobic properties

to the morphology of the coatings. To this end, optical microscopy was performed by using a

12
LEICA MZ12 and LEICA DM RXE microscopies, whereas the Scanning Electron Microscopy

(SEM) analyses was carried out by using an Inspect F model FEI apparatus at an accelerating

voltage of 10 kV. In this case, the samples were metallized twice in order to improve the quality of

images. Measurements of surface roughness were obtained using a confocal and interferometry 3D

microscope (Leica DCM 3D), by using a 20x objective. The measurements were carried out along

two directions, parallel and perpendicular to the acquisition directions, respectively, on a scan size

of about 1 mm2. Two coupons for each kind of sample were tested (referred as sample 1 and sample

2).

Measurements of transmittance in the visible field (400-700 nm) were performed on the transparent

TR-series by UV−Vis analysis, using an Agilent Technologies Cary 60 UV−Vis spectrophotometer.

Finally, in order to highlight the effective deposition of the coating deposition on the selected

substrates, FT-IR analysis was performed by using a Nicolet IR operating in simple ATR mode.

The spectra were collected in the range 600-4000 cm-1 with a resolution of 2 cm-1.

2.3. Results and discussion

For both series of coated samples, identified as OP-DR-n and OP-S-n, the contact angle measured at

room temperature increases with the deposition layers number (see Figure S1 and Figure S2 of the

Supporting Information), reaching respectively a value of 148° with 15 layers, and 138° with 30

layers. Therefore, when the silica nanoparticles were dispersed by sonication, higher values of

contact angle were achieved with a lower number of deposited layers. Additionally, the

hydrophobic behavior of the OP-S-n samples disappears once the samples were submitted to the

aging test. Morphological investigations addressed to evidence the effect of aging on the coated

surface showed that the formation and the aggregation of silica in micro and nano-sized sheets-like

13
features are well evident soon after coating deposition, but they disappear as effect of the aging

treatment (see Figure S3 of the Supporting Information).

Hence, the OP-SiO2, TR-SiO2, and TR-PR-SiO2 series were prepared by ultra-sonicating the silica

nanoparticles into the THF for 30 minutes (see Table 1) in order to assure homogeneous dispersion

meanwhile reducing the preparation time with respect to the OP-DR-n series. The silica dispersion

layers were deposited by using an aerograph. With this method a contact angle value of 152°

(Figure 3 and Figure 4), which does not change significantly after aging, was obtained for the OP-

SiO2 sample. However, at -12°C and 0.5 bar this contact angle value decreases to 128° (Figure 3). It

is worth observing that the coating developed in this paper evidences a contact angle, which is

always higher than the reference coating, regardless the conditions of measurements, i.e., ambient

and harsh temperatures and pressures, and after aging (Figure 3). Contrary to what happens for the

OP-S-n series (see Figure S3 of the Additional Information), the sheets-like structures are stable to

the aging test (Figure 4), since the only effect of the aging was to reduce the dimensions of these

aggregates.

14
Figure 3: contact angles measurements performed at room temperature, before and after aging, and

measurements at -12°C and 0.5bar, for the OP-REF and OP-SiO2 samples. In the insert, a picture of

the OP-SiO2 sample during the measurements of contact angle at -12°C and 0.5bar.

Figure 4: optical images of OP-REF and OP-SiO2 samples before and after aging at 50.4x,

100.8x, and 550x, respectively. On the left side, contact angles at room temperature with

water.

The enhanced contact angle can be ascribed to the increased roughness of the developed coating

with respect to the reference; in fact, as reported in Table 2 and in Figure 5, the Ra value increases

15
from 0.218 m to 0.310 m with the application of the coating. How the surface roughness plays an

important role in the tunable wettability of surfaces is widely reported in literature [46]-[49].

Directly correlated to the surface roughness are both the size of the exposed particles as well as

particles aggregate [5] and the coverage degree of the surface with those particles/aggregates [31].

Indeed, Conti et al. [49] found that the roughness and the surface wettability change with the size of

the silica nanoparticles, reaching a maximum value with nanoparticles of 80 nm in size. The

coverage degree of substrate, for its part, has to provide the desired distance between the

nanoparticles [31], [49], [50] in order to guarantee the Cassie-Baxter condition [51]. In the present

study, it is evident that the particle aggregates created a sort of mountain-like structures, visible in

Figure 5, able to guarantee that favorable condition. Additionally, SEM images (Figure 6) highlight

the formation of a porous structure, onto the surface of these mountain-like structures, which

probably allows the trapping of small-scale air pockets, reducing the contact area between the liquid

drop and the surface [32]. Moreover, a hierarchical roughness characterized by features on the

micro-scale appears on the top of the coating, responsible of its superhydrophobic behavior. In

correspondence of the cross section of the coating, distinguishable particles together with the typical

porous structure of silica can be observed in Figure 5 (see bottom row of Figure 5).

Therefore, the combination of the multiscale roughness (micro- and nano-), the presence of the air

pockets and the relative distance among the peaks worked in a synergic way to enhance the

superhydrophobic behavior of the developed coating.

16
Table 2: surface roughness measurements of the OP-REF and OP-SiO2 samples.

Ra [m]

OP-REF OP-SiO2

Parallel Perpendicular Parallel Perpendicular

sample 1 sample 1

0.194 0.225 0.243 0.246

0.231 0.244 0.290 0.274

0.200 0.215 0.232 0.410

sample 2 sample 2

0.193 0.208 0.301 0.350

0.229 0.183 0.334 0.307

0.213 0.278 0.349 0.388

Average Ra [m]

0.218 0.310

17
Figure 5: 3D surface topography of OP-REF (a) and OP-SiO2 (b) samples obtained by

interferometric/confocal microscopy.

Figure 6: SEM images of the OP-SiO2 coating (top) and cross-section (bottom).

18
FT-IR measurements performed on both the OP-REF and the OP-SiO2 samples are shown in Figure

7. The spectrum of the hydrophobic silica R812 is also reported. It is worth noting that in the OP-

SiO2 spectrum, the symmetrical and asymmetrical stretching of Si-O in Si-O-Si groups at about

1060 cm-1 and the rocking of CH3 in Si(CH3)3 at 843 cm-1 and Si(CH3)2 at 808 cm-1 [52] appear.

Both signals are characteristic of the silica R812, whereas the hydrophilic -OH stretching at 3300

cm-1 [52], present in the OP-REF, are quite absent in the OP-SiO2 coating. Hence, the new coating

completely covers the hydrophilic possible sites characteristic of the commercial coating.

Figure 7: FT-IR measurements of OP-REF and OP-SiO2 samples. The spectrum of the hydrophobic

silica R812 is reported.

19
Figure 8 shows the images of the contact angles measurements carried out with water, formamide

and diiodomethane at room temperature for the OP-SiO2 and OP-REF samples, before and after the

aging test. The results show that the coating developed in this work is very stable to the aging

treatment, since the contact angles do not change as effect of the treatment. On the contrary, its

wettability to the diiodomethane slightly increases from 30.5° to 16°. An opposite behavior can be

observed for the OP-REF coating, since after aging the water contact angle decreases from 73° to

62°, and contemporary the diiodomethane contact angle slightly increases. It should also be noticed

that the wettability of OP-SiO2 surface to diiodomethane is higher than the reference OP-REF

(Figure 8). This means that the new coating allows to improving the apolar interactions with respect

to the polar ones, as the very low wettability to water and formamide demonstrated (Figure 8).

These measurements allowed to assessing the SFE and its components, whose values are shown in

Figure 9 for both the OP-SiO2 and OP-REF samples.

20
Figure 8: contact angles of OP-REF and OP-SiO2 samples before and after aging performed using

water, diiodomethane and formamide.

21
Figure 9: SFE and its components for the Op-REF and OP-SiO2 sample, measured at room

temperature before and after aging and at -12°C and 0.5 bar.

It was found that the SFE of the OP-SiO2 coating is always higher than the reference OP-REF, in

both ambient and harsh temperatures and pressures, and after aging test. Additionally, for both the

OP-SiO2 and OP-REF samples, the dispersive component of the SFE is always predominant with

respect to the polar component (Figure 9). The apparent SFE depends on both the superficial

chemistry and roughness of the coating. On the one hand, the chemistry of surface is largely

changed by the application of the superhydrophobic developed coating, as FT-IR spectra (Figure 7)

and wettability to diiodomethane (Figure 8) demonstrate. In details, the FT-IR spectra show a

reduction of the –OH groups confirmed by the improvement of the affinity to dispersive liquid

(such as diiodomethane) with respect to the reference sample, which causes an increase in the SFE

dispersive component. The latter is responsible of the increased total SFE especially in flight

conditions (see Figure 9). On the other hand, the SEM images demonstrate that the roughness

increases since a hierarchical structure appears on the coating (Figure 6). Finally, the increased

22
roughness and the decreased hydrophilic –OH groups number cause the reduction of both

wettability and work of adhesion between water and surface observed for the OP-SiO2 sample with

respect the OP-REF (Figure 10).

Figure 10: wettability and work of adhesion of the OP-REF and OP-SiO2 sample measured at room

temperature before and after aging, and at -12°C and 0.5 bar.

The described bottom-up preparation method is substrate independent. In fact similar results in

terms of contact angles, SFE, work of adhesion and wettability were obtained for specimens

realized by using PMMA substrates (see from Figure S5 to Figure S9 in the Additional

Information), in spite of the reduced micro-roughness, which is an intrinsic characteristic of the OP-

substrate.

23
Additionally, the effect of the coating deposition on the transparency of TR-SiO2 and TR-PR-SiO2

samples was assessed by transmittance measurements in the visible spectrum 400-700 nm. Figure

11 shows that the transmittance of the coated PMMA specimens ranges between 50 and 80% for the

TR-SiO2 and TR-PR-SiO2 samples.

Figure 11: measurements of transmittance in the visible spectrum for the TR-REF, TR-PR-SiO2, and

TR-SiO2 samples.

A picture of the TR-PR-SiO2 sample with a droplet of water on the top, highlighting the

hydrophobic behavior of the coating, is reported in Figure 12 a). The transparency of TR-coated

samples is compared to the reference TR-REF in Figure 12 b). These images confirm the goodness

of the transparency despite of the measured transmittance values (Figure 11).

24
Figure 12: (a) picture of the TR-PR-SiO2 sample with a droplet of water on the top, and (b)

comparison of transparency among TR-REF, TR-PR-SiO2 and TR-SiO2 samples.

Finally, a video reproducing the behavior of droplets of water, which fall down on the new coated

and reference surfaces, tilted at 45°, were recorded. In Figure 13 and in Figure 14 some frames are

shown. It must be highlighted that when the droplet of water hits the surface of the OP-SiO2 sample

(Figure 13), it flattens into a pancake shape, then regains a round shape and bounces up, showing

the typical behavior of superhydrophobic coatings.

Figure 13: frames recorded meanwhile a water droplet falls over the OP-SiO2 (top) and OP-REF

(bottom) surfaces.

25
Whereas, when the droplet of water hits the surface of the TR-PR-SiO2, it rolls over the surface, as

shown in Figure 14. This is a specific characteristic largely required for self-cleaning surfaces. On

the contrary, when the droplet of water hits both reference samples, i.e., the OP-REF and TR-REF,

it adheres to its surface (see bottom of Figure 13 and Figure 14). The behavior displayed by the

superhydrophobic coatings can be ascribed to the low work of adhesion between water and

surfaces, since the lower the work of adhesion, the lower the energy needed to detach the water

droplet, realized with a combination of both changed surface chemistry and hierarchical micro- and

nano-roughness.

Figure 14: frames recorded meanwhile a water droplet falls over the TR-PR-SiO2 (top) and TR-REF

(bottom) surfaces.

Conclusions

In this paper, a simplified method to achieve large superhydrophobic surfaces is described. The

coating consists of a dispersion in THF of hydrophobic silica nanoparticles. It was found that a

short sonication of 30 minutes is enough to obtain a stable and homogenous silica dispersion. Spray

coating was found to offer benefits of rapid patterning over large areas on both planar and curved
26
surfaces. Results showed that the superhydrophobicity was achieved by the chemical modification

of the surface, which reduces the surface free energy, and by the nanoscale roughness imparted by

the silica nanoparticles in combination with the micro-scale roughness. This roughness is inherent

on the one side to the intrinsic substrate roughness and on the other side to the aggregation of silica

nanoparticles, which occurs during the coating deposition and is promoted by the solvent

evaporation. Therefore, a hierarchical roughness features, with overlapped micro- and nano-sized

roughness, were achieved. In addition, it was found that the silica is organized in porous aggregates,

which allow the trapping of small-scale air pockets.

Aknowledgements

The authors thank to Mrs. Angela Ferrigno for the optical microscopy observations, Mr. Vincenzo

Fiorillo for the acquisition of videos with a camera at high velocity. The authors are grateful Prof.

Antonino Squillace and Dr. Carla Velotti from DICMaPI Department of Federico II UNiversity of

Naples, for the profilometer measurements.

References

[1] K. Venkataramani, L. McVey, R. Holm, K. Montgomery, Inclement weather considerations

for aircraft engines, AIAA-2007-0695 AIAA, Reston (2007).

[2] Mason, W. Strapp, P. Chow, The ice particle threat to engines in flight, Vol 4 (2006) 2445-

2465, AIAA-2006-0206.

[3] F. Wang, C. Li, Y. Lv, Y.Du, A facile Superhydrophobic Surface for Mitigating Ice

Accretion, Properties and Applications of Dieletric Materials, (2009) 150-153.

[4] S.Q. Yang, Q. Xia, L. Zhu, J. Xue, Q. Wang, Q. Chen, Research on the icephobic properties

of fluoropolymer-based materials, Applied Surface Science 257 (11); 2011: 4956–4962.

27
[5] L.L. Cao, A.K. Jones, V.K. Sikka, J. Wu, D. Gao, Anti-icing superhydrophobic coatings,

Langmuir 25 (21); (2009) 12444–12448.

[6] P Tourkine, M. Le Merrer, D. Quéré, Delayed freezing on water repellent materials,

Langmuir 25 (13); (2009) 7214–7216.

[7] M. Susoff, K. Siegmann, C. Pfaffenroth, M. Hirayama, Evaluation of icephobic coatings-

Screening of different coatings and influence of roughness, Applied Surface Science 282

(2013): 870-879.

[8] G. Momen, M. Farzaneh, Facile approach in the development of icephobic hierarchically

textured coatings as corrosion barrier, Applied Surface Science 299 (2014) 41-46.

[9] T. Strobl, S. Storm, M. Kolb, J. Haang, M. Hornung, Development of a hybrid ice protection

system based on nanostructured hydrophobic surfaces, 29th Congress of the International

Council of the Aeronautical Science, St. Petersburg, Russia, September 7-12, (2014).

[10] H. Sojoudi, M. Wang, N. D. Boscher, G. H. McKinley, K. K. Gleason, Durable and

scalable icephobic surfaces: similarities and distinctions from superhydrophobic surfaces,

Soft Matter 12 (2016) 1938-1963.

[11] A. Lazauskas, A. Guobiene, I. Prosycevas, V. Baltrusaitis, V. Grigaliunas, P. Narmontas, J.

Baltrusaitis, Water droplet behavior on superhydrophobic SiO 2 nanocomposite films

during icing/deicing cycles, Materials characterization 82 (2013) 9-16.

[12] A. Dotan, H. Dodiuk, C. Laforte, S. Kenig, The relationship between water wetting and

ice adhesion, Journal of Adhesion Science and Technology 23 (15); (2009) 1907–1915.

[13] H. Saito, K. Takai, G. Yamauchi, A study on ice adhesiveness to water-repellent coating,

Materials Science Research International 3 (3); (1997) 185–189.

[14] S.A. Kulinich, M. Farzaneh, Ice adhesion on super-hydrophobic surfaces, Applied Surface

Science 255 (18) (2009) 8153–8157.

28
[15] S.A. Kulinich, S. Farhadi, K. Nose, X.W. Dull, Superhydrophobic surfaces: are they really

ice-repellent?, Langmuir 27 (1); (2011) 25–29.

[16] X.M. Li, D. Reinhoudt, M. Crego-Calama, What do we need for a superhydrophobic

surface? A review on the recent progress in the preparation of superhydrophobic surfaces,

Chem. Soc. Rev., 36 (8) (2007) 1550-1568.

[17] J. Fresnais, J. P. Chapel, L. Benyahia, F. Poncin-Epaillard, Plasma-Treated

Superhydrophobic Polyethylene Surfaces: Fabrication, Wetting and Dewetting Properties,

J Adhes. Sci. Technol., 23 (2009) 447-467.

[18] M. Hikita, K. Tanaka, T. Nakamura, T. Kajiyama, A. Takahara, Super-Liquid-Repellent

Surfaces Prepared by Colloidal Silica Nanoparticles Covered with Fluoroalkyl Groups,

Langmuir, 21 (2005) 7299-7302.

[19] H. T. Yang, P. Jiang, Self-Cleaning Diffractive Macroporous Films by Doctor Blade

Coating, Langmuir, 26 (2010) 12598-12604.

[20] J. Bravo, L. Zhai, Z. Z Wu, R. E. Cohen, M. F. Rubner, Transparent Superhydrophobic

Films Based on Silica Nanoparticles, Langmuir, 23 (2007) 7293-7298.

[21] Hikita et al. 2005; Yang & Jiang, 2010; Furstner, R.; Barthlott, W.; Neinhuis, C.; Walzel,

P. Wetting and Self-Cleaning Properties of Artificial Superhydrophobic Surfaces,

Langmuir, 21 (2005) 956-961.

[22] N. Wang, D. Xiong, Y. Deng, Y. Shi, K. Wang, Mechanically Robust Superhydrophobic

Steel Surface with Anti-Icing, UV-Durability, and Corrosion Resistance Properties, ACS

Appl. Mater. Interfaces, 7 (2015) 6260–6272.

[23] L. Li, V. Breedveld, D. W. Hess, Creation of Superhydrophobic Stainless Steel Surfaces by

Acid Treatments and Hydrophobic Film Deposition, ACS Appl. Mater. Interfaces, 4

(2012) 4549–4556.

29
[24] S. Barthwal, Y. S. Kim and S.-H. Lim, Mechanically Robust Superamphiphobic

Aluminum Surface with Nanopore-Embedded Microtexture, Langmuir, 29 (2013) 11966–

11974.

[25] D. Infante, K. W. Koch, P. Mazumder, L. Tian, A. Carrilero, D. Tulli, D. Baker, V.

Pruneri, Durable, superhydrophobic, antireflection, and low haze glass surfaces using

scalable metal dewetting nanostructuring, Nano Res., 6 (2013) 429–440.

[26] A. Steele, B. K. Nayak, A. Davis, M. C. Gupta, E. Loth, Linear abrasion of a titanium

superhydrophobic surface prepared by ultrafast laser microtexturing, J. Micromech.

Microeng., 23 (2013) 115012-115020.

[27] A. Y. Vorobyev, C. L. Guo, Multifunctional surfaces produced by femtosecond laser

pulses, J. Appl. Phys., 117 (2015) 0331031-0331037.

[28] C.-H. Xue, J.-Z. Ma, Long-lived superhydrophobic surfaces, J. Mater. Chem. A, 1 (2013)

4146–4161.

[29] D. T. Ge, L. L. Yang, G. X. Wu, S. Yang, Spray coating of superhydrophobic and angle-

independent coloured films, Chem. Commun., 50 (2014) 2469–2472.

[30] Y. Liao, C.-H. Loh, R. Wang, A. G. Fane, Electrospun Superhydrophobic Membranes with

Unique Structures for Membrane Distillation, ACS Appl. Mater. Interfaces, 6 (2014)

16035–16048.

[31] I. Yilgor, S. Bilgin, M. Isik, E. Yilgor, Facile preparation of superhydrophobic polymer

surfaces, Polymer, 53 (2012) 1180–1188.

[32] J. Liu, Z.A. Janjiua, M. Roe, F. Xu, B. Turnbull, K.S. Choi, X. Hou, Super-

hydrophobic/icephobic coatings based on silica nanoparticles modified by self-assembled

monolayers, Nanomaterials 6 (2016) 232-242.

30
[33] S. Amigoni, E. T. de Givenchy, M. Dufay, F. Guittard, Covalent Layer-by-Layer

Assembled Superhydrophobic Organic−Inorganic Hybrid Films, Langmuir, 25 (2009)

11073–11077.

[34] H. Sojoudi, G. H. McKinley, K. K. Gleason, Linker-free grafting of fluorinated polymeric

cross-linked network bilayers for durable reduction of ice adhesion, Mater. Horiz., 2

(2015) 91–99.

[35] E. J. Lee, J. J. Kim, S. O. Cho, Fabrication of Porous Hierarchical Polymer/Ceramic

Composites by Electron Irradiation of Organic/Inorganic Polymers: Route to a Highly

Durable, Large-Area Superhydrophobic Coating, Langmuir, 26 (2010) 3024–3030.

[36] https://www.aerosil.com/sites/lists/RE/DocumentsSI/Technical-Overview-AEROSIL-

Fumed-Silica-EN.pdf

[37] D. K. Owens, R. C. Wendt, Estimation of the SFE of polymers, J. Appl. Polym. Sci. 13

(1968) 1741-1747,

[38] M. Zenkiewicz, Methods for the calculation of SFE of solids, Polymer Testing 26; (2007)

14-19.

[39] A. Rudawska, E. Jacniacka, Analysis of Determining SFE Uncertainty with the Owens-

Wendt method. Inter. J. Adhes. Adhes. 29 (2009) 451-457.

[40] https://www.kruss.de/fileadmin/user_upload/website/literature/kruss-tn306-en.pdf.

[41] L. Holysz, E. Chibowski, K. Terpilowsi. Contact Angle, Wettability and Adhesion, Vol. 5,

K. L. Mittal (Ed.), VSP/Brill, Leiden 2008.

[42] L. Mazzola, L. Coriand, N. Felde, G. Bruno, B Galasso, V Quaranta, F Albano, A Auletta,

Development and Technological Characterization of Multi-functional Aeronautical

Coating From Lab-Scale to the Relevant Environment, Journal of Aeronautics &

Aerospace Engineering 6 (1) (2017) 1-16

31
[43] L. Mazzola, Aeronautical livery coating with icephobic property, International Journal of

Surface Engineering 32 (10) (2016), 733-744.

[44] Airworthiness Standars FAA FAR Part 25.

http://www.engineerstoolkit.com/Airworthiness%20Standards%20%20FAA%20FAR%20

Part%2025.pdf, 2007 (accessed 31/8/2017).

[45] MIL-STD-810G, Department of defense test method standard: environmental engineering

considerations and laboratory tests (31 OCT 2008), (http://everyspec.com/MIL-STD/MIL-

STD-0800-0899/MIL-STD-810G_12306/).

[46] I. Yilgor, S. Bilgin, M. Isik, E. Yilgor, Tunable wetting of polymer surfaces, La,gmuir, 28

(2012) 14808-14814.

[47] S. Kojevnikova, A. Marmur, Multi-scale roughness and the Lotus effect: Discontinuous

liquid-air interfaces, Colloids and Surfaces A: Physicochemical and Engineering Aspects,

521 (2017) 78-85.

[48] T. Bharathidasan, S.V. Kumar, M.S. Bobji, R.P.S. Chakradhar, B.J. Basu, Effect of

wettability and surface roughness on ice-adhesion strength of hydrophilic, hydrophobic and

superhydrophobic surfaces, Applied Surface Science 314 (2014) 241–250.

[49] J. Conti, J. De Coninck, M.N. Ghazzal, Design of water-repellant coating using dual scale

size of hybrid silica nanoparticles on polymer surface, Applied Surface Science 436 (2018)

234–241.

[50] J. De Coninck, F. Dunlop, T. Huillet, Is superhydrophobicity robust withrespect to

disorder? Eur. Phys. J. E 36 (2013) 104–112.

[51] A.B.D. Cassie, S. Baxter, Wettability of porous surfaces, Trans. Faraday Soc. 40(1944)

546–551.

32
[52] A Handbook of Spectroscopic data Chemistry, B.D. Mistry, Oxford Book company,

Jaipur, India, ISBN: 978-81-89473-86-0, Edition 2009.

33
 It has been developed a simplified and not-expensive method to prepare effective

superhydrophobic coatings for aeronautical applications.

 The new developed coatings has contact angle at room temperature up to 154° instead of 73°

of the reference.

 The hydrophobicity resists to aging able to simulate its durability in flight condition.

 The new coating can be applied to both opaque and transparent, rough and smooth

substrates achieving good results.

 The new coating modifies both the chemistry and the morphology of the surface, by

substituting hydrophilic –OH groups, typical of the commercial reference, with hydrophobic

groups, along with aggregation of silica nanoparticles in porous structures, which allow the

trapping of small-scale air pockets.

34

You might also like