Sunteți pe pagina 1din 9

Article

pubs.acs.org/IECR

A Facile Electrodeposition Process for the Fabrication of


Superhydrophobic and Superoleophilic Copper Mesh for Efficient
Oil−Water Separation
Yan Liu,†,‡ Kaiteng Zhang,† Wenguang Yao,† Chengchun Zhang,*,† Zhiwu Han,† and Luquan Ren†

Ministry of Education, Key Laboratory of Bionic Engineering, Jilin University, Changchun 130022, China

State Key Laboratory of Automotive Simulation and Control, Jilin University, Changchun 130022, China

ABSTRACT: A superhydrophobic and superoleophilic copper mesh with excellent oil−water separation efficiency was
successfully fabricated via electrodeposition and then surface modification with lauric acid. The surface morphologies, chemical
composition, and wettability were characterized by means of scanning electron microscopy (SEM), atomic force microscopy
(AFM), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared (FT-IR) spectroscopy, and water contact angle
(WCA) measurements. It was found that the as-prepared surface is both superhydrophobic and superoleophilic, with static WCA
values as high as 155.5° ± 3° and an oil contact angle (OCA) of 0°. Furthermore, the as-prepared surface exhibited excellent oil−
water separation efficiency including petroleum, toluene, hexane, gasoline and diesel, even after being recycled 10 times. In
addition, the as-prepared copper mesh shows self-cleaning character with water and chemical stability. This study provides a
facile, inexpensive, and environmentally friendly route to fabricate large-scale and excellent oil−water separation surface with high
separation efficiency for a great number of potential applications.

1. INTRODUCTION In the past several years, numerous methods have been


In the past few years, with the increase of oil spill accidents and reported for the manufacture of the superhydrophobic surface
for oil−water separation on metal meshes such as chemical
industrial oily wastewater, oil−water separation has become a
etching,24 electro-spinning,25,26 layer by layer,27,28 self-assem-
worldwide challenge and has aroused much attention.1−4 To
bly,29,30 sol−gel method,31−36 ARGET-ATRP,37 zeolite-coated
resolve these water pollution issues, many approaches have
method,38 self-assembly of block co-polymers,39 chemical vapor
been used for the treatment of oil−water pollution, including deposition (CVD) method,40−42 electrodeposition, and so
filtration, oil skimmers, centrifugal machine, precipitation tanks, on.43,44 Meanwhile, many scholars have focused on the research
magnetic separations, flotation technologies, oil-absorbing of underwater superoleophobic surface. Dudchenko et al.45
materials, and combustion.5−11 However, most of the methods have demonstrated a membrane-based and fouling-free oil/
involve harsh conditions, such as expensive equipment, water separation method that couples carbon nanotube
complex devices, complicated processing steps, high processing poly(vinyl alcohol) underwater superoleophobic ultrafiltration
cost, long processing time, and so on. Taking all factors into membranes with magnetic Pickering emulsions. Dunderdale et
account, separation treatment stands out from the rest as a al.46 highlighted the pH-responsive superoleophobicity of
simple, universal, scalable approach for valid removal of oil pDMAEMA brush surfaces, which gives excellent oil drop
from water. mobility and low adhesion underwater.
In addition, surfaces with superhydrophobic and super- However, to better resolve oil−water separation in industry,
oleophilic wettability have triggered unprecedented research the method of facile and large-scale preparation must be
excitement in the field of filtration.12−15 Many plants and developed.
insects exhibit excellent superhydrophobicity, such as lotus Copper mesh is widely used for the separation of oil and
leaves, rose petals, marigold petals, water striders, butterfly water, because of its porous structure, which is composed of
wings, rice leaves, mosquito eyes, and so on. Taking inspiration microfibers. Based on the above-mentioned discussions, we
from nature, superhydrophobic surfaces are fabricated by reported a simple, highly efficient, and low-cost route to
simulating typical structures of plant surface and considering preparing superhydrophobic copper meshes. In this research,
chemical composition simultaneously.16−20 Also, materials with copper meshes with superhydrophobic and superoleophilic
special wettability and micro-nanoscale structures have surfaces were fabricated by electrodeposition and then modified
attracted increasing attention for oil−water separation. This with lauric acid. The electrodeposition process takes place at
filtration process needs excellent wettability materials, which are atmospheric pressure, using a solution of 0.2 M sulfuric acid
both superhydrophobic (water contact angle (WCA) of >150°) and 0.15 M copper sulfate that were cured under ambient
and superoleophilic (oil contact angle (OCA) of <5°). Oil can
selectively penetrate through the mesh while water is blocked. Received: September 21, 2015
Designing and fabricating coatings with special wettability Revised: February 5, 2016
would be a facile and effective way to solve the problem of oil− Accepted: February 21, 2016
water separation.21−23 Published: February 22, 2016

© 2016 American Chemical Society 2704 DOI: 10.1021/acs.iecr.5b03503


Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

Figure 1. SEM images of newly prepared copper mesh samples, showing the surface morphology of the copper mesh at different electrodeposition
times in an aqueous solution of 0.2 M sulfuric acid and 0.15 M copper sulfate: (a) 0 s, (b) 100 s, (c) 150 s, (d) 250 s, (e) 300 s, and (f) 400 s. The
insets show magnified images of the corresponding surfaces.

conditions. The superhydrophobic copper meshes can be easily remove the oxide film and organic matter on the surface of the
combined with an independent device, which acts as a filter, copper mesh, respectively. The obtained mesh then was used as
forming a barrier that allows oils to pass through while repelling the cathode and a copper plate was used as the anode. The
water. reaction solution was contained 0.2 M sulfuric acid and 0.15 M
copper sulfate. Finally a direct current (DC) voltage of 4.0 V
2. EXPERIMENTAL METHODS was applied to the two electrodes for different times (100, 150,
2.1. Materials. The 40 mm × 10 mm × 2 mm copper plate 200, 250, 300, 350, and 400 s) with a distance of 3 cm. At the
and copper mesh (120, 150, and 180 mesh woven from 0.08, end of the experiment, the sample was washed with a sufficient
0.07, and 0.05 mm diameter wire, respectively) were obtained amount of distilled water. The resulting copper meshes then
from Hebei Shijiazhuang Yuanpeng Metal Manufacturing were dried in a vacuum oven at 60 °C for 20 min.
Works. The meshes had pore diameters of 124, 100, and 80 2.3. Surface Modification. The prepared copper meshes
μm, respectively. Lauric acid (C12H24O2, 97.5%, Tianjin were immersed in a 1.5 × 10−3 M ethanol solution of lauric acid
Guangfu Fine Chemical Research Reagent Institute), copper and for 12 h under ambient conditions. The −COOH groups
sulfate (CuSO4, 99%, Tianjin Zhiyuan Chemical Reagent can be chemically adsorbed onto the native oxide surface of
Institute). All other chemicals were purchased from Beijing copper clusters and the surface-active molecules form a self-
Chemical Works, were of the highest commercially available assembled monolayer.39,40 The resulting copper meshes then
grade, and were used without further purification. This includes were dried in a vacuum oven at 60 °C for 10 min.
hydrochloric acid, anhydrous ethanol, chloroform, Methylene 2.4. Sample Characterization. The water/oil contact
Blue, toluene, petroleum ether, and hexane. angle (WCA/OCA) values were measured by using a contact
2.2. Sample Preparation. Copper plate and meshes were angle meter (Model OCA 20, Dataphysics, Germany) at room
ultrasonically rinsed for 5 min in a solution of diluted temperature. Water droplets (3 μL) were carefully dropped
hydrochloric acid, anhydrous ethanol, and distilled water to onto the surfaces, and the average static contact angle (CA)
2705 DOI: 10.1021/acs.iecr.5b03503
Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

value was obtained by measuring five different positions of each the surface morphology of the copper meshes, we deduce that
surface. The surface morphologies of as-prepared samples were the ideal electrodeposition time is 250 s.
examined via field-emission scanning electron microscopy Furthermore, the three-dimensional (3-D) morphology of
(SEM) (Model EVO 18, Zeiss) and atomic force microscopy the as-prepared copper mesh has been characterized by AFM.
(AFM) (Model Dimension Icom, Bruker). The chemical Figure 2 shows the AFM images and surface roughness values
composition of the samples was characterized using an X-ray
photoelectron spectroscopy (XPS) system (Model XR50,
SPECS, Japan) and Fourier transform infrared (FT-IR)
spectrophotometer (JASCO, Japan). Oil concentrations were
measured by an automatic infrared oil analyzer (Model JLBG-
158, Xi’an Yima Opto-Electrical Technology Co., Ltd., China).
2.5. Oil−Water Separation Apparatus. The device
consists of two quartz tubes, which are 100 mm and 150 mm
in length, respectively, and 30 mm in diameter. The as-prepared
copper mesh was fastened between the quartz flange fixtures by
four screws, and the device was positioned an a tilt angle of
∼20°. The oil/water mixture was poured onto the mesh. The
separation was achieved driven by gravity.

3. RESULTS AND DISCUSSION


3.1. Surface Morphology. The wettability of solid surfaces
is strongly dependent on both the geometrical structure and the
chemical composition.47 Figures 1a−e show SEM images of the Figure 2. (a) SEM image of the as-prepared copper mesh sample; (b)
woven copper mesh formed using ∼100-μm-diameter wires as two-dimensional (2-D) AFM image of the area bounded by the red
substrates. The typical image of a copper mesh substrate that square in panel (a); (c) three-dimensional (3-D) AFM image of the
area bounded by the red square in panel (a); and (d) fluctuation of the
has been knitted by copper wires with a pore diameter of ∼100 height for the red line shown in panel (b).
μm is shown in Figure 1a, which indicates that the pristine
mesh has a smooth surface. The morphology of the copper
mesh was greatly influenced by the electrodeposition time. of the copper-coated mesh. Taking into account the fact that
Figure 1b shows the SEM image of the copper mesh with an the copper wire is a curved surface, we can only roughly
electrodeposition time at 100 s. It can be clearly seen that the calculate the root-mean-square surface roughness (RMS). The
surface was relatively smooth, with only some small spherical height of the micro-nanoparticles is concentrated in the range
particles. No obvious submicrometer and nanoscale particles of 0−1.1 μm. The surface roughness of the coating is 993 nm
are observed in the magnified images. As shown in Figure 1c, as (RMS value).
the electrodeposition time increased to 150 s, the microscale 3.2. Chemical Characterization. XPS is used to test the
particles on the surface of the copper mesh became rougher, chemical composition of the thin film of the as-prepared
and the spherical particles were increased. Figure 1d shows the surfaces modified after electrodeposition. Figure 3 shows the
SEM image of the copper mesh after an electrodeposition time XPS spectrum of the as-prepared superhydrophobic surface
at 250 s, spherical particles several hundred nanometers in size after modification with lauric acid. It reveals the presence of
were uniformly dispersed on the surface of the wires. It can be carbon, oxygen, and copper on the as-prepared surfaces. One
seen that the pristine mesh wires has been completely covered can see the strong peak of C 1s at 284.78 eV and O 1s at 531.2
by circular particles. There is a large amount of micro-nanoscale eV. Figure 3c presents the Cu 2p XPS spectrum of the as-
circular particles, which can capture a large amount of air. prepared copper mesh surface. This region includes one
Therefore, it provides the geometric conditions for the highlight peak at 933.5 eV; therefore, elemental copper is
formation of hydrophobicity. However, as the time increases expected to exist on the as-prepared mesh surface after
to 300 s (Figure 1e), the size of the particles was much larger electrodeposition. Figures 3b and 3d reveal the presence of
than that observed at 250 s, and the rough micro-nanoscale carbon and oxygen on the as-prepared surfaces. One can see
structure, which was slightly damaged because of spherical the strong peak of C 1s at 284.78 eV and O 1s at 531.2 eV.
submicrometer structure, is replaced by a larger crystal Compared to the untreated copper mesh, the contents of
structure with diameters of 5−10 μm. Longer electrodeposition elemental carbon and oxygen were obviously increased on the
times encouraged the formation of larger surface features that, as-prepared surfaces, which were treated with lauric acid, which
in some locations, resembled broccoli structures. It can be seen indicated that the as-prepared coating on the as-prepared
from Figure 1f that these broccoli structures grew orthogonally copper mesh surface was copper, and the surface was
out of the mesh wire walls. These findings show that an successfully modified with lauric acid.
increase in electrodeposition time leads to a change in The sample was modified to reduce the surface energy after
morphology, resulting in a transition of the wetting state the electrodeposition of copper. In order to confirm the
from hydrophobic to superhydrophobic. The microsized presence of lauric acid on the coated surface of the copper
spherical particles, in combination with the nanoscale granular mesh, FT-IR spectra was used. Figure 4 shows FT-IR spectra of
protuberance, endowed a hierarchical composite structure to the electrodeposition mesh before and after being modified by
the mesh that possessed superhydrophobic behaviors. Such lauric acid. Compared to the sample before modification, many
micro-nanoscale hierarchical structures can help support water absorption bands are detected on modified sample surface,
droplets and are necessary for superhydrophobic surfaces. From which indicates that the copper mesh surface has been
2706 DOI: 10.1021/acs.iecr.5b03503
Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

Figure 3. XPS spectra of copper mesh fabricated by electrodeposition and then surface modification with lauric acid coatings.

were successfully prepared on the surface of copper mesh


through this method. It was expected that the micro-nanoscale
hierarchical composite structure and chemical component
made the as-prepared mesh possess superhydrophobicity, and
the surface wettability was evaluated by CA measurement.
Figure 5a shows that the water contact angle (WCA) of the
pristine copper mesh was 79°. For the copper mesh after an
electrodeposition time of 250 s only and the sample modified
by lauric acid only, the WCAs were measured to be 118° and
120°, respectively, as shown in Figures 5b and 5c. In contrast,
an oil droplet (toluene) with a low surface tension quickly
spreads on the surface and permeates through it, exhibiting a
highly oleophilic property. Figure 5d shows the image of a
water droplet (3 μL) on the as-prepared copper mesh. Its WCA
approaches 155.5° ± 3°, indicating that the surface has
superhydrophobicity.
Figure 4. FT-IR spectra of the copper mesh treated by electro- The resulting meshes transitioned to a dull red color from
deposition only (black trace) and copper mesh treated by electro- their pristine purplish-red color and showed superhydrophi-
deposition then modified by lauric acid (red trace).
licity, as shown in Figure 6a. When the as-prepared super-
hydrophobic copper mesh and original hydrophilic copper
successfully modified. The FT-IR spectrum exhibits typical mesh were immersed in water, a plastron (air pockets) that was
absorption features of the coupled −COO− stretching indicative of a robust Cassie−Baxter state was formed on the
vibrations at 1701 cm−1, CH3− stretching vibrations at 2920 as-prepared mesh surface only. The submerged as-prepared
cm−1, and CH2 stretching vibrations at 2855 cm−1. It is known copper mesh looked like a silver mirror when viewed at a
that the CH3 group had a surface energy of 24 mJ/m2 and the glancing angle, as shown in Figure 6b. To further illustrate that
CH2 group had a surface energy of 31 mJ/m2, both of which the as-prepared mesh has better reflectance, two samples were
were widely used to reduce the free energy. The FT-IR spectra immersed in water again. The reflectance of these two samples
indicated that the lauric acid was bound to the copper mesh was measured in the visible light range. As shown in Figure 6c,
surfaces. the as-prepared sample had a reflectance of ∼40%. However,
3.3. Surface Wettability. 3.3.1. Formation of Super- the original sample had a reflectance of ∼8%, because of the
hydrophobicity. A combination of SEM, XPS, and FT-IR reflectance of light at the air layer trapped on the surface. This
characterizations shows that the micro-nanoscale hierarchical trapped air can effectively prevent wetting on the as-prepared
structures were created and that low-surface-energy materials copper mesh surface under water. After several minutes of
2707 DOI: 10.1021/acs.iecr.5b03503
Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

property. Thus, the as-prepared copper mesh possessed


superhydrophobic with low adhesion character. According to
the test results, we deduce that the ideal electrodeposition time
is 250 s.
In addition, we find that the pore size is also important,
because different pore sizes can result in different wetting
performances. Therefore, we chose the above-mentioned 150
mesh as a contrast test and carried out another two groups of
experiments to investigate the relationship between pore size
and wetting performance. Figure 7 displays the WCAs as a
function of the electrodeposition time of the electrochemical
deposition (all of three copper meshes were obtained with the
same immersion time and were modified with the same
concentration solution of the lauric acid). As shown in Figure 8,
for the maximum pore size (120 mesh) of copper mesh, an
increase in plating time results in a gradual increase in CA
values. However, the hydrophilic state remained (WCA < 65°)
after an electroplating time of 400 s. The resulting electro-
deposition is deposited on the copper wirethe larger the pore
size, the smaller the effective deposition area will beso a large
pore size is not conducive to building superhydrophobic micro-
nanostructures. To get the superhydrophobic state, a longer
Figure 5. Surface wettability of the pristine and as-prepared mesh: (a)
plating time may be required. In contrast, for the copper mesh
the photograph of water droplet on the surface of pristine copper (180 mesh) with a minimum pore size, an opposite trend in CA
mesh; (b) the photograph of water droplet on the surface of copper is observed: an increase in plating time results in a gradual
mesh after 250 s of deposition only; (c) the photograph of water decrease in CA values. This is because a smaller aperture will
droplet on the surface of copper mesh modified by lauric acid only; provide better microscale roughness for the copper mesh,
and (d) the photograph of water droplet on the surface of as-prepared which is an essential necessity to obtain the superhydropho-
copper mesh. bicity. Compared with copper mesh (180 mesh) with a pore
size of 80 μm, the copper mesh (150 mesh) with a pore size of
immersion testing, the material surfaces were completely dry 100 μm had a larger pore size. Thus, copper mesh with a larger
and still displayed a large CA (>150°) with water droplets. pore size is more favorable for the permeation of water. Taking
3.3.2. Effect of the Electrodeposition Time on Wettability. into account the properties of velocity, plating time, and
The fabrication of nanostructures on a copper mesh by wetting, we can determine that the optimal pore size is ∼100
electrodeposition is an inexpensive and facile technique; μm.
however, this method always needs to accurately control the 3.4. Self-Cleaning Effect. As a comparison, self-cleaning
deposition time and solution concentration, which play tests were carried out with two types of copper mesh, including
important roles in crystal growth. Because the roughness and pristine and as-prepared, as shown in Figure 9. Here, the
crystallinity are positively related. Thus, we explored the effect graphite is used to represent the conventional pollutant.
of different electrodeposition times on wettability. The samples Removal of fine graphite powder on the superhydrophobic
were prepared under 100, 150, 200, 250, 300, 350, and 400 s of surface is documented in the above images (see Figures 9a−c).
deposition time. Figure 7a shows the relationship between For graphite particles, excellent self-cleaning properties were
electrodeposition times and static CAs on the as-prepared observed. The graphite particles are wetted by the water and
surface. The CA value reached 134° after a deposition time of imbibed into the droplet as it rolls off the surface. Super-
100 s. With the time was extended to 250 s, the CA value hydrophobicity is essential for such self-cleaning particles.
reached 155.5° ± 3° and the sliding angle reached 13° ± 4°. When the pristine copper mesh was used, water droplets
Meanwhile, the advancing and receding angles shown in Figure adhered to the surface and could neither imbibe the particles
7b were measured, still showing very good hydrophobic nor roll off the surface, as shown in Figure 9b. Thus, no self-

Figure 6. (a) Macroscopic images of the pristine and as-prepared samples; (b) mirror-like phenomenon can be observed on the resulting
superhydrophobic mesh submerged in water. (c) reflectance of the original mesh (red trace) and as-prepared superhydrophobic copper mesh (black
trace).

2708 DOI: 10.1021/acs.iecr.5b03503


Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

Figure 7. (a) Contact angles and (b) sliding angles of the different electrodeposition time on simple surfaces for 100, 150, 200, 250, 300, 350, and
400 s after modification by lauric acid.

Figure 8. Influence of pore sizes of copper mesh on water contact


angle (WCA).

cleaning properties were observed on the pristine copper mesh


surface.
3.5. Chemical Stability. It is well-known that Cassie-like
wetting is temporary under water. We have explored temporal
changes to the CA value over extended periods. We put the as-
prepared sample into the water for different time periods and
then removed it and immediately measured the CA. Figure 10
shows the relationship between the soaking time and static CA.
After 2 h of soaking, the sample can keep a CA value of 140° ±
3° and has good hydrophobicity. However, after the sample was Figure 9. Demonstration of the self-cleaning effect of the resultant
superhydrophobic mesh through the removal of graphite particles
dried after a certain period of time, its original super-
from the surface using water droplets.
hydrophobicity was again restored. Therefore, the as-prepared
sample could exist stably in the water and lauric acid does not
dissolve into the water phase. 4. SEPARATION OF OIL AND WATER
In order to further illustrate the stability of samples prepared
under corrosive environment, the sample was estimated by Because of a combination of superhydrophobicity and super-
oleophilicity, the as-prepared mesh has great potential to be
measuring the static WCA values for samples that were
applied to oil−water separations. The oil−water separation
immersed in aqueous solutions for 10 min at various pH values experiment was performed as shown in Figure 12 The as-
(ranging from 1 to 14). Figure 11 shows the relationship prepared copper mesh was fixed between two quartz tubes. In
between the pH values and the static CA; the measured static order to ensure the sealing performance, there are two sealing
CA ranged from 143° ± 3° to 152° ± 2°. After immersion in rings in the junction of the two quartz tubes. Compared to
the corrosive solution, the samples still have excellent water, oil has a lower density. Therefore, the measured oil floats
hydrophobic proprities. The results indicate that the as- on the water, with no contact with the copper mesh.
prepared surface has good chemical stability in aqueous The short glass tube was replaced by a longer bent one, and
solutions of most acidic, alkali, and some aqueous salts. the device was placed at a tilt angle of ∼30°. In this experiment,
2709 DOI: 10.1021/acs.iecr.5b03503
Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

Figure 10. Relationship between the soaking time and the static
contact angle (CA) on the as-prepared copper mesh surface.

Figure 11. Relationship between the pH values and the static CA on Figure 12. Oil−water separation device and process: (a) before mixing
the as-prepared copper mesh surface. (petroleum ether and water), (b) after mixing, (c) after separation.
Water was colored by Methylene Blue, and oil was colored by Oil Red
O.
25 mL of petroleum ether (dyed with oil red) was poured into
25 mL of water (dyed with Methylene Blue) contained in a respectively. The oil concentration were measured using an
beaker, and the petroleum ether floated on top of the aqueous automatic infrared oil analyzer. As shown in Figure 13, different
phase, because of its lower density. When the mixture of
petroleum ether and water with a total volume of 50 mL was
poured into the quartz tube, the petroleum ether penetrated the
mesh and flowed down the beaker underneath, while the water
was retained in the upper glass tube. During the oil−water
separation process, no artificial external force was employed,
indicating its easy operation and low energy costs. Note that
the ubiquitous water loss induced by the water adhesion on the
surface of glass vessels during the oil−water separation process
was not considered in the efficiency calculations.
Oil−water separation efficiency was used to quantitatively
describe the oil−water separation ability of the as-prepared
copper mesh. Because the superhydrophobic layer is robust, the
as-prepared mesh can be reused. Therefore, we investigated the
suitability of a copper mesh for the separation of other oils from
water. The recycling ability was also investigated by taking the Figure 13. Oil collection efficiency of superhydrophobic mesh capped
petroleum ether/water mixture as an example. We measured container for five oils after 10 cycles.
the separation efficiency of the as-prepared copper mesh with
different types of oil, such as petroleum, toluene, hexane,
gasoline, and diesel. The oil separation efficiency (R, expressed types of oils such as gasoline, diesel oil, petroleum, toluene, and
as a percentage) was then calculated based on the following hexane, can be separated using the above apparatus at an
equation: efficiency of more than 93%. The stability of the as-prepared
mesh differs only slightly, even after 10 cycles, indicating its
⎛ C ⎞ general suitability for various oil−water separations. After 10
R (%) = ⎜1 − P ⎟ × 100 separate experiments, the sample was rinsed with sufficient
⎝ CO ⎠
alcohol and distilled water to remove the surface of the oil, and
Herein, CO and CP are the oil concentrations in the pristine then it was dried under atmospheric conditions for 10 min.
oil−water mixture and after the separation of water, Measuring the CA again shows that the CA approached 148° ±
2710 DOI: 10.1021/acs.iecr.5b03503
Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

3°, which indicates that the organic solvent will not damage the (8) Zouboulis, A. I.; Avranas, A. Treatment of Oil-in-Water
structure and chemical composition of the hydrophobic surface. Emulsions by Coagulation and Dissolved-Air Flotation. Colloids Surf.,
There is a good combination of the laurel acid and the rough A 2000, 172, 153.
surface; therefore, the lauric acid does not dissolve into the oil (9) Al-Shamrani, A. A.; James, A.; Xiao, H. Separation of Oil from
phase. Water by Dissolved Air Flotation. Colloids Surf., A 2002, 209, 15.
(10) Zhou, X.; Zhang, Z.; Xu, X.; Men, X.; Zhu, X. Facile Fabrication
of Superhydrophobic Sponge with Selective Absorption and
5. CONCLUSIONS Collection of Oil from Water. Ind. Eng. Chem. Res. 2013, 52, 9411.
A facile method to fabricate a copper mesh with selective (11) Lin, K. A.; Yang, H.; Petit, C.; Hsu, F. Removing Oil Droplets
wettability has been demonstrated. Superhydrophobic and from Water Using a Copper-Based Metal Organic Frameworks. Chem.
superoleophilic meshes were fabricated by an electrodeposition Eng. J. 2014, 249, 293.
method and then modified on lauric acid coatings. The as- (12) Gu, C. D.; Xu, X. J.; Tu, J. P. Fabrication and Wettability of
prepared copper mesh surface not only exhibits super- Nanoporous Silver Film on Copper from Choline Chloride-Based
Deep Eutectic Solvents. J. Phys. Chem. C 2010, 114, 13614.
hydrophobicity, with a water contact angle (WCA) of 155.5°
(13) Wang, B.; Liang, W.; Guo, Z.; Liu, W. Biomimetic Super-
± 3°, but also superoleophilicity with an oil contact angle lyophobic and Super-lyophilic Materials Applied for Oil/Water
(OCA) of 0°. In addition, a series of oil/water mixtures, such as Separation: A New Strategy Beyond Nature. Chem. Soc. Rev. 2015,
petroleum ether/water, hexane/water, toluene/water, gasoline/ 44, 336.
water, and diesel/water, were observed to be separated by the (14) Islam, Md. S.; Choi, W. S.; Kim, S. H.; Han, O. H.; Lee, H.-J.
mesh, with a separation efficiency of >93%; it remained high Inorganic Micelles (Hydrophilic Core@Amphiprotic Shell) for
even after 10 cycles. Furthermore, the as-prepared super- Multiple Applications. Adv. Funct. Mater. 2015, 25, 6061.
hydrophobic copper mesh shows self-cleaning character and (15) Dunderdale, G. J.; Urata, C.; Sato, T.; England, M. W.; Hozumi,
chemical stability. The fabrication process is facile, inexpensive, A. Continuous, High-Speed, and Efficient Oil/Water Separation Using
and environmentally friendly, and suited to large-scale Meshes with Antagonistic Wetting Properties. ACS Appl. Mater.
applications. Because of the excellent oil−water performance, Interfaces 2015, 7, 18915.
the as-prepared copper meshes have potential applications in (16) Yin, X. Y.; Liu, Z. L.; Wang, D. A.; Pei, X. W.; Yu, B.; Zhou, F.
treating industrial oil−water mixtures and environmental oil Bioinspired Self-Healing Organic Materials: Chemical Mechanisms
spills. and Fabrications. J. Bionic Eng. 2015, 12, 1.


(17) Wang, G. Y.; Guo, Z. G.; Liu, W. M. Interfacial Effects of
Superhydrophobic Plant Surfaces: A Review. J. Bionic. Eng. 2014, 11,
AUTHOR INFORMATION 325.
Corresponding Author (18) Yan, Y. Y.; Gao, N.; Barthlott, W. Mimicking Natural
*Tel.: +86 431 85095760. Fax: +86 431 85095575. E-mail: Superhydrophobic Surfaces and Grasping the Wetting Process: A
address:sdzcc@jlu.edu.cn. Review on Recent Progress in Preparing Superhydrophobic Surfaces.
Adv. Colloid Interface Sci. 2011, 169, 80.
Notes (19) Jiang, T.; Guo, Z.; Liu, W. Biomimetic Superoleophobic
The authors declare no competing financial interest. Surfaces: Focusing on Their Fabrication and Applications. J. Mater.


Chem. A 2015, 3, 1811.
ACKNOWLEDGMENTS (20) Feng, L.; Zhang, Y.; Xi, J.; Zhu, Y.; Wang, N.; Xia, F.; Jiang, L.
Petal Effect: A Superhydrophobic State with High Adhesion Force.
The authors thank the National Natural Science Foundation of Langmuir 2008, 24, 4114.
China (Nos. 51275555, 51475200, and 51325501), Science and (21) Lahann, J. Environmental Nanotechnology: Nanomaterials
Technology Development Project of Jilin Province (No. Clean Up. Nat. Nanotechnol. 2008, 3, 320.
20150519007JH). (22) Yao, X.; Song, Y.; Jiang, L. Applications of Bio-inspired Special

■ REFERENCES
(1) Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.;
Wettable Surfaces. Adv. Mater. 2011, 23, 719.
(23) Zhang, F.; Zhang, W. B.; Shi, Z.; Wang, D.; Jin, J.; Jiang, L.
Nanowire-Haired Inorganic Membranes with Superhydrophilicity and
Underwater Ultralow Adhesive Superoleophobicity for High-Efficiency
Marinas, B. J.; Mayes, A. M. science and technology for water
Oil/Water Separation. Adv. Mater. 2013, 25, 4192.
purification in the coming decades. Nature 2008, 452, 301.
(24) Guo, W.; Zhang, Q.; Xiao, H.; Xu, J.; Li, Q.; Pan, X.; Huang, Z.
(2) Schwarzenbach, R. P.; Escher, B. I.; Fenner, K.; Hofstetter, T. B.;
Johnson, C.; Gunten, A. U.; Wehrli, B. The Challenge of Micro- Cu Mesh’s Super-hydrophobic and Oleophobic Properties with
pollutants in Aquatic Systems. Science 2006, 313, 1072. Variations in Gravitational Pressure and Surface Components for
(3) Dalton, T.; Jin, D. Extent and Frequency of Vessel Oil Spills in Oil/Water Separation Applications. Appl. Surf. Sci. 2014, 314, 408.
U.S. Marine Protected Areas. Mar. Pollut. Bull. 2010, 60, 1939. (25) Wang, L.; Yang, S.; Wang, J.; Wang, C.; Chen, L. Fabrication of
(4) Montgomery, M. A.; Elimelech, M. Water and Sanitation in Superhydrophobic TPU Film for Oil−Water Separation Based on
Developing Countries: Including Health in the Equation. Environ. Sci. Electrospinning Route. Mater. Lett. 2011, 65, 869.
Technol. 2007, 41, 17. (26) Wu, J.; Wang, N.; Wang, L.; Dong, H.; Zhao, Y.; Jiang, L.
(5) Zhou, S.; Jiang, W.; Wang, T.; Lu, Y. Highly Hydrophobic, Electrospun Porous Structure Fibrous Film with High Oil Adsorption
Compressible, and Magnetic Polystyrene/Fe3O4/Graphene Aerogel Capacity. ACS Appl. Mater. Interfaces 2012, 4, 3207.
Composite for Oil−Water Separation. Ind. Eng. Chem. Res. 2015, 54, (27) Zhang, L.; Zhong, Y.; Cha, D.; Wang, P. A Self-Cleaning
5460. Underwater Superoleophobic Mesh for Oil−Water Separation. Sci.
(6) Hu, B.; Scott, K. Influence of Membrane Material and Rep. 2013, 3, 2326.
Corrugation and Process Conditions on Emulsion Microfiltration. J. (28) Wang, J.; Shi, Z.; Fan, J.; Ge, Y.; Yin, J.; Hu, G. Self-Assembly of
Membr. Sci. 2007, 294, 30. Graphene into Three-Dimensional Structures Promoted by Natural
(7) Elias, E.; Costa, R.; Marques, F.; Oliveira, G.; Guo, Q.; Thomas, Phenolic Acids. J. Mater. Chem. 2012, 22, 22459.
S.; Souza, F. G., Jr. Oil-spill cleanup: The influence of acetylated (29) Liu, N.; Cao, Y.; Lin, X.; Chen, Y.; Feng, L.; Wei, Y. A Facile
curaua fibers on the oil-removal capability of magnetic composites. J. Solvent-Manipulated Mesh for Reversible Oil/Water Separation. ACS
Appl. Polym. Sci. 2015, 132, 41732. Appl. Mater. Interfaces 2014, 6, 12821.

2711 DOI: 10.1021/acs.iecr.5b03503


Ind. Eng. Chem. Res. 2016, 55, 2704−2712
Industrial & Engineering Chemistry Research Article

(30) Wang, L.; Pan, K.; Li, L.; Cao, B. Surface Hydrophilicity and
Structure of Hydrophilic Modified PVDF Membrane by Nonsolvent
Induced Phase Separation and Their Effect on Oil/Water Separation
Performance. Ind. Eng. Chem. Res. 2014, 53, 6401.
(31) Lu, S.; Chen, Y.; Xu, W.; Liu, W. Controlled Growth of
Superhydrophobic Films by Sol−Gel Method on Aluminum Substrate.
Appl. Surf. Sci. 2010, 256, 6072.
(32) Yang, H.; Pi, P.; Cai, Z.; Wen, X.; Wang, X.; Cheng, J.; Yang, Z.
Facile Preparation of Super-hydrophobic and Super-oleophilic Silica
Film on Stainless Steel Mesh via Sol−Gel Process. Appl. Surf. Sci.
2010, 256, 4095.
(33) Su, C.; Xu, Y.; Zhang, W.; Liu, Y.; Li, J. Porous Ceramic
Membrane with Superhydrophobic and Superoleophilic Surface for
Reclaiming Oil from Oily Water. Appl. Surf. Sci. 2012, 258, 2319.
(34) Leventis, N.; Chidambareswarapattar, C.; Bang, A.; Sotiriou-
Leventis, C. Cocoon-in-Web-like Superhydrophobic Aerogels from
Hydrophilic Polyurea and Use in Environmental Remediation. ACS
Appl. Mater. Interfaces 2014, 6, 6872.
(35) Li, R.; Chen, C.; Li, J.; Xu, L.; Xiao, G.; Yan, D. Facile Approach
to Superhydrophobic and Superoleophilic Graphene/Polymer Aero-
gels. J. Mater. Chem. A 2014, 2, 3057.
(36) Zhang, G.; Li, M.; Zhang, B.; Huang, Y.; Su, Z. A Switchable
Mesh for On-Demand Oil−Water Separation. J. Mater. Chem. A 2014,
2, 15284.
(37) Dunderdale, G. J.; England, M. W.; Urata, C.; Hozumi, A.
Polymer Brush Surfaces Showing Superhydrophobicity and Air Bubble
Repellency in a Variety of Organic Liquids. ACS Appl. Mater. Interfaces
2015, 7, 12220.
(38) Wen, Q.; Di, J.; Jiang, L.; Yu, J.; Xu, R. Zeolite-Coated Mesh
Film for Efficient Oil−Water Separation. Chem. Sci. 2013, 4, 591.
(39) Mansur, H. S.; Orefice, R. L.; Mansur, A. Characterization of
Poly(vinyl alcohol)/poly(ethylene glycol) Hydrogels and PVA-
Derived Hybrids by Small-Angle X-ray Scattering and FTIR
Spectroscopy. Polymer 2004, 45, 7193.
(40) Crick, C. R.; Gibbins, J. A.; Parkin, I. P. Superhydrophobic
Polymer-Coated Copper-Mesh; Membranes for Highly Efficient Oil−
Water Separation. J. Mater. Chem. A 2013, 1, 5943.
(41) Zhang, J.; Seeger, S. Polyester Materials with Superwetting
Silicone Nanofi Laments for Oil/Water Separation and Selective Oil
Absorption. Adv. Funct. Mater. 2011, 21, 4699.
(42) Zhou, X.; Zhang, Z.; Xu, X.; Men, X.; Zhu, X. Facile Fabrication
of Superhydrophobic Sponge with Selective Absorption and
Collection of Oil from Water. Ind. Eng. Chem. Res. 2013, 52, 9411.
(43) Wang, S.; Song, Y.; Jiang, L. Microscale and Nanoscale
Hierarchical Structured Mesh Films with Superhydrophobic and
Superoleophilic Properties Induced by Long-Chain Fatty Acids.
Nanotechnology 2007, 18, 015103.
(44) Wang, B.; Guo, Z. Superhydrophobic Copper Mesh Films with
Rapid Oil/Water Separation Properties by Electrochemical Deposition
Inspired from Butterfly Wing. Appl. Phys. Lett. 2013, 103, 063704.
(45) Dudchenko, A. V.; Rolf, J.; Shi, L.; Olivas, L.; Duan, W.; Jassby,
D. Coupling Underwater Superoleophobic Membranes with Magnetic
Pickering Emulsions for Fouling-Free Separation of Crude Oil/Water
Mixtures: An Experimental and Theoretical Study. ACS Nano 2015, 9,
9930.
(46) Dunderdale, G. J.; Urata, C.; Miranda, D. F.; Hozumi, A. Large-
Scale and Environmentally Friendly Synthesis of pH Responsive Oil-
Repellent Polymer Brush Surfaces under Ambient Conditions. ACS
Appl. Mater. Interfaces 2014, 6, 11864.
(47) Gao, N.; Yan, Y. Y. Characterisation of surface wettability based
on nanoparticles. Nanoscale 2012, 4, 2202.

2712 DOI: 10.1021/acs.iecr.5b03503


Ind. Eng. Chem. Res. 2016, 55, 2704−2712

S-ar putea să vă placă și