Vous êtes sur la page 1sur 10

Applied Surface Science 258 (2012) 7872–7881

Contents lists available at SciVerse ScienceDirect

Applied Surface Science


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

Effects of exposure time on variations in the structure and hydrophobicity of


polyvinylidene fluoride membranes prepared via vapor-induced phase
separation
Yuelian Peng ∗ , Hongwei Fan, Yajun Dong, Yanna Song, Hua Han
College of Environmental and Energy Engineering, Beijing University of Technology, Beijing 100124, PR China

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

Article history: The present investigation revealed how the surface morphology and hydrophobicity of polyvinylidene
Received 26 October 2011 fluoride (PVDF) membranes, which were prepared via a vapor-induced phase separation (VIPS) method,
Received in revised form 15 February 2012 were affected by the exposure time. The mass variation of the cast film was recorded. Membrane mor-
Accepted 16 April 2012
phologies were observed by scanning electron microscopy (SEM) and thermal behaviors of membranes
Available online 23 April 2012
were examined by differential scanning calorimetry (DSC). Wide angle X-ray diffraction (WAXD) was
employed to analyze the crystalline structures of the overall membranes and the surface layers. The
Keywords:
results showed that different membrane morphologies and hydrophobicities could be obtained by chang-
Surface morphology
Hydrophobicity
ing the exposure time. A long exposure time facilitated the crystallization process, resulting in the
Polyvinylidene fluoride membranes formation of a porous skin and particle morphology, which increased the hydrophobicity of the sur-
Vapor-induced phase separation face. A short exposure time favored the formation of a digitate macrovoid and dense skin resulting from
Vacuum membrane distillation liquid–liquid phase separation in the immersion process, which reduced surface hydrophobicity. The
water permeate flux in vacuum membrane distillation was greatly affected by the membrane porosity
and surface hydrophobicity.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction for these systems, water is used as the non-solvent, so that the
water transfer associated with a humid air exposure step leads
Phase separation is a widely used method for making microp- to phase separation. The extent and rate of water transfer can be
orous polymer membranes. There are two main phase separations controlled by adjusting the air velocity, relative humidity (RH) and
involved in this process, i.e., non-solvent induced phase separation temperature of the air, as well as the exposure time. By a judicious
(NIPS) and thermally induced phase separation (TIPS), depending combination of these variables, symmetric membranes with large
on the driving force for the phase separation. NIPS is the most pores at the top surface can be obtained [1–7].
important method in the preparation of reverse osmosis, nanofil- This unique hierarchical micro- and nanostructuring on the
tration, ultrafiltration and microfiltration membranes. Typically, a leg’s surface of water striders therefore appears to be responsible
clear solution (dope) consisting of a polymer, a solvent and usu- for its water resistance and the strong supporting force [8]. VIPS
ally a non-solvent is formed into the desired shape (e.g., a flat has also been used to construct a superhydrophobic surface. Xie
sheet or hollow fiber) and is then contacted with a non-solvent et al. [9] used water-vapor-induced phase separation in an ambi-
that is miscible with the solvent. The ensuing transfer of solvent ent environment to prepare a fluorine-end-capped polyurethane
and non-solvent leads to phase inversion and the formation of a (FPU)/poly (methyl methacrylate) (PMMA) film that possessed a
three-dimensional porous polymer network. Wet casting is char- micro- and nanoscale binary surface structure similar to the sur-
acterized by rapid mass-transfer rates and typically, the formation face of a lotus leaf and a sponge-like bulk structure. The water
of a dense skin at the top surface of the membrane. contact angle (CA) of the film was 166◦ , whereas the oil CA was
The non-solvent vapor-induced phase-separation (VIPS) pro- 140◦ . This film is thus classified as a superamphiphobic surface.
cess, which is a NIPS process, has attracted an increasing amount Peng et al. [10] prepared a porous and highly hydrophobic sur-
of attention in recent years. In VIPS, the cast film is first exposed face of polyvinylidene fluoride (PVDF) by the gelation of a PVDF/N,
to humid air prior to immersion in the coagulation bath. Typically, N-dimethylacetamide (DMAc) casting solution in open air instead
of its immersion into a precipitation bath. It was found that the
membranes exhibited certain micro- and nanoscale hierarchical
∗ Corresponding author. Tel.: +86 10 67391090; fax: +86 10 67391983. surface roughnesses, and the water CA of the samples was as high
E-mail address: pyl@bjut.edu.cn (Y. Peng). as 150◦ .

0169-4332/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.apsusc.2012.04.108
Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881 7873

Membrane distillation (MD) processes or membrane contactors


have many industrial applications, mostly in water desalination
[11–16] and wastewater treatment [17–20]. In the MD process,
the liquid entry pressure (LEP), defined as the minimum trans-
membrane pressure that is required for process liquid to enter the
pores of the membrane, should be as high as possible. A high LEP
reduces the occurrence of pore wetting that can lead to the deteri-
oration of the quality and rate of permeate production. The LEP is
a characteristic of each membrane, which is calculated as follows:

−2BL cos 
LEP = (1)
rmax

where B is a membrane tortuosity factor with a value that is


typically assumed to be 2,  L is the surface tension of the solu-
tion with a value of 69.63 mN/m for water at 40 ◦ C,  is the CA
between the solution and the membrane surface, and rmax is the
largest membrane pore radius. Eq. (1) indicates that a high LEP
can be achieved by using materials of high hydrophobicity (i.e.,
Fig. 1. Diagram of the saturated air generator.
large contact angles with feed solutions) and small maximum pore
radius. If the maximum pore radius of a membrane is 1 ␮m [2],
the LEPs of such a membrane with contact angles of 95◦ and 150◦ 2. Experimental
calculated from Eq. (1) are 24 and 241 kPa, respectively, which
would be the upper application pressure limits. These calculations 2.1. Materials
demonstrate the large effect of the CA on the LEP and subse-
quently on the maximum operating pressures. Therefore, to avoid PVDF was obtained from ShangHai New Materials Company
wetting, the hydrophobicity of a membrane is very important in Limited (3FNM, FR-904, intrinsic viscosity1.4–1.9 dL/g). DMAc
commercial design [15]. Thus, the preparation of a hydrophobic (analytical reagent) was purchased from Tianjin Yongda Chemical
microporous membrane surface is an actively pursued goal in the Reagent Co. (China). A commercial PVDF membrane with pore size
current development of membrane science. El-Bourawi et al. [21] 0.1 ␮m was provided by Beijng Institute of Plastic Research.
and Khayet [22] have reviewed the characteristics of most of the
commercially available membranes used in MD processes. These 2.2. Membrane preparation
membranes are mainly made of polypropylene (PP), polyethy-
lene (PE), PVDF and polytetrafluoroethylene (PTFE). PVDF is a PVDF membranes were prepared via a VIPS method. 15 wt.%
commercially available fluoropolymer with a low surface energy PVDF (predried in an oven at 120 ◦ C) was dissolved in DMAc at
(25 dynes/cm), and microporous hydrophobic PVDF membranes room temperatures. After the solution became homogeneous, it
prepared by the phase inversion method have been used success- was degassed overnight at room temperature. The degassed poly-
fully for many years in MD and membrane contactors. An increasing mer solution was cast onto a glass plate to a thickness of 250 ␮m,
number of researchers are focusing on how to prepare microporous and the cast film was exposed to humid air at 25 ◦ C over the planned
and hydrophobic PVDF membranes [23–30]. Superhydrophobic time interval; the cast film was then immersed into a water bath to
surfaces reported in the literature are mainly found in materials in allow for solidification of the resulting morphology. Thirty minutes
which both hydrophobicity and surface roughness have been com- later, the solidified film was transferred to another water bath for
bined. Xie et al. [9] and Peng et al. [10] prepared porous and highly 24 h to remove residual DMAc. The membrane was then dried in
hydrophobic PVDF surfaces with certain micro- and nanoscale hier- air at room temperature. The time that the cast films were exposed
archical roughnesses via VIPS. Unfortunately, the ambient RH was to the humid air was varied between 0 min and 10 min. The bath
not controlled in these studies, and the formation mechanism of the temperature was kept constant at 25 ◦ C.
hydrophobic surfaces was not further discussed. Therefore, more In VIPS, the RH of the humid air is very important; in this study,
research is needed to determine how the RH of the environment it was controlled at 100% in the device shown in Fig. 1 during mem-
and other conditions influence the micro- and nanostructures and brane formation. Briefly, a Pyrex© glass chamber (custom fabricated
hydrophobicity. in our laboratory) was placed in a thermostatic water bath (HH-I,
To date, most studies on VIPS have focused on the phase sepa- RongHua Apparatus Manufacturer), and a grille with open area of
ration mechanism and the mass transformation during membrane 60% was placed between the chamber and the water bath. There
formation. To the best of the authors’ knowledge, few studies was a movable drawer in the chamber that was used to support the
have concentrated on the membrane surface structure changes glass plate and the cast film. Before the glass plate with the film
occurring in VIPS, especially the hydrophobicity changes in PVDF. was placed into the drawer, the temperature in the water bath was
This study therefore investigated the membrane surface hydropho- set to 40 ◦ C. Water continuously evaporated, and thus, the RH in
bicity changes in VIPS, in which a wide range of membrane the chamber increased. The temperature in the chamber was mon-
morphologies was produced by controlling the phase-separation itored with the temperature probe and was shown on the panel,
process. The major factor affecting the morphology of the PVDF which was 25 ◦ C. When water droplets were observed on the roof,
membrane, namely, exposure time and the mechanisms that gov- suggesting that the air in the chamber was saturated with water,
ern their formation are discussed. The membrane morphology, the RH value was assumed to be 100%. The glass plate with the cast
hydrophobicity and crystallinity were investigated. The membrane film was then placed in the drawer, and the drawer was immedi-
performance was investigated in vacuum membrane distillation ately pushed back into the chamber. After a planned exposure time
(VMD). period, the glass plate was removed and placed in the water bath.
7874 Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881

2.3. Mass transfer 2.4.8. Porosity


The total porosity (Ak , %) was calculated as a function of the
The glass plate with the cast film was placed on a balance scale membrane weight as follows:
(AR5120, Ohaus, USA) in the Pyrex© glass chamber described above
W1 − W2
with a 100% RH value. The mass variation of the cast film was Ak = × 100% (3)
A × ı × H2 O
recorded in situ every 10 min, and the data were saved automati-
cally by a computer connected to the balance. where W1 and W2 are the equilibrium weights of a membrane in the
swollen and dry states, respectively, A is the area of the membrane,
2.4. Membrane characterization ı is the average thickness of the membrane, and H2 O is the density
of water.
2.4.1. Morphological characterization
The morphology of the top surface (defined as the surface in con- 2.4.9. VMD process
tact with the coagulation bath or humid air), the bottom surface (the A flat-sheet VMD configuration with an area of 26.4 × 10−4 m2
surface in contact with the glass plate), and the cross-section of the was used to evaluate the performance of membranes prepared in
sample films was characterized by scanning electron microscopy this experiment and a purchased PTFE membrane. Fig. 2 shows a
(SEM, JSM-5510LV, JEOL, Japan; FEI Quanta 200, Holland; Hitachi, schematic diagram of the VMD experimental apparatus. The hot
S-4300, Japan) operated at 5–20 kV. Prior to the observations, all feed flow was controlled by a peristaltic pump (BT300-1J, Baod-
samples were coated with gold in vacuum. ing Longer Precision Pump Co, Ltd.). The temperature of the feed
was controlled by a thermostat water bath. The hot water temper-
2.4.2. Static contact angle ature inside the configuration was measured, after steady state was
The static CA of the PVDF membrane was measured using the reached, by a sensor connected to a digital temperature gauge with
sessile drop method with a contact angle meter (DSA100, Krüss, an accuracy of ±0.1 ◦ C. A vacuum pump (ZXZ-2, Linhai Tan Vac-
Germany). The water droplets used for the measurements were uum Equipment Co, Ltd.) was connected to the permeate side of the
5 ␮L in volume. The CA values of each sample were measured at configuration to remove the vapor. The vacuum in the downstream
five different positions of one sample and then averaged. was measured with a vacuum gauge with an accuracy of ±0.1 kPa.
The feed temperature, feed flow and vacuum were maintained at
2.4.3. Mechanical properties 50 ◦ C, 48 L/h, and 40 kPa, respectively. A filter flask connected to a
Wet porous membranes were cut into the standard dumbbell chiller was used to collect the permeate water. The VMD flux was
shape for the measurements of tensile strength and modulus. The calculated in every case by weighing the permeate water collected
tensile strengths and elastic moduli at the breaking point were in the filter flask for a predetermined period. Practically all VMD
measured using an Instron 4302 universal tensile testing machine experiments were repeated twice to ensure reproducibility of the
under ambient conditions for at least three samples, and the aver- measurements.
age values were recorded. A 3.5 wt.% NaCl aqueous solution was employed as feed, to assess
the membrane flux and NaCl rejection rate. A conductivity indica-
2.4.4. Differential scanning calorimeter (DSC) tor was used to measure NaCl concentration in the feed solution
Thermal behaviors of the membranes were determined by a and permeate solutions. Because of the good linear relationship
differential scanning calorimeter (PerkinElmer Pyris-1 DSC, USA) between concentration and conductivity, the permeate concentra-
calibrated with indium. The membrane crystallization was char- tion can be calculated. The NaCl rejection rate (R) was determined
acterized by DSC at 10 ◦ C/min increasing temperature rate. About as follows
10 mg sample was used. Melting heat Hm was determined from Cf − Cp
R= × 100 (4)
the melting peak area. Crystallinity was evaluated by: Cf
H where Cf , Cp are the NaCl concentration in the feed and permeate
DSC = (2)
H100 solution respectively.

H100 = 104.7 J/g is the melting heat for a 100% crystalline sample
3. Results
of PVDF [31].
3.1. Membrane structure
2.4.5. Wide-angle X-ray diffraction (WXRD)
Crystalline structures of the overall membranes were investi-
The cast films were exposed to humid air (25 ◦ C, 100% RH) over
gated by WXRD measurements. WAXD patterns with 2 angles
the planned time interval and then immersed in a water bath. The
ranging from 10◦ to 50◦ were obtained on a rotating-anode X-ray
exposure time was varied between 0 min and 10 min. The depen-
diffractometer (Bruker D8 Advance, Germany). All the X-ray diffrac-
dence of the final membrane cross-section, top and bottom surface
tometer data given below were recorded by with voltage 40 kV and
morphologies on the exposure time are shown in Figs. 3–5, respec-
current 40 mA applied to the X-ray rotating-anode.
tively. To clearly show the structures formed by VIPS, the walls
of the microvoids in the cross-section shown on the left side of
2.4.6. X-ray photoelectron spectroscopy (XPS) Fig. 3 are magnified and displayed on the right side of Fig. 3; great
The XPS analyses were carried out with a PHI Quantera SXM differences in the structures of the films prepared with different
(ULVAC-PH INC) spectrometer using a monochromatic Al K␣ source exposure times were readily observed by SEM.
(20 mA, 2.0 kV). Binding energy accuracy is ±0.5 eV and an analysis When the exposure time was close to 0 min, the typical large,
area of 1 mm × 1 mm. digitate macrovoids were formed (Fig. 3a-1), the top surface was
dense and smooth, and no micropores were observed by SEM
2.4.7. Mean pore diameter and pore size distribution (Fig. 4a), which is the characteristic structure obtained with the wet
The pore size distribution of all membrane samples was deter- process (i.e., direct immersion of the cast film into a water bath).
mined by a modified bubble point method as described elsewhere Furthermore, the walls of the microvoids showed a cellular struc-
[32]. ture (Fig. 3a-2). As the exposure time was increased to 3 min, the
Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881 7875

Fig. 2. Schematics of the experimental setup used for the VMD experiments.

digitate macrovoid structures remained but became shorter; the change when the exposure time was increased, but the porosity
bottom of the macrovoids still had a cellular structure, although the increased greatly. Membranes Nos. 1, 2, 3 and 4 had similar CAs,
cells were slightly larger with the exposure time of 3 min than those but the permeate flux increased with increasing porosity. For the
with 0 min. With an exposure time of 5 min, the digitate macrovoids three membranes Nos. 4–6, the flux increased greatly; for these
completely disappeared and the cross-section showed a spherulitic membranes, the porosity only increased from 83.4% to 85.4%, a
morphology (Fig. 3c-1), but a large amount of cellular structure net increase of 2.0%, so the hydrophobicity increase played the
remained among the particles (Fig. 3c-2). The aggregate spheres major role. Because membranes Nos. 7 and 8 possessed similar
were mostly 1–2 ␮m in diameter. With an exposure time of 6 min, morphologies (Fig. 3d and e), porosities and CA values (Table 1),
the cellular structure again decreased but still persisted (Fig. 3d-2). they had approximately the same permeate fluxes. The permeate
When the exposure time was further increased to 10 min, the cellu- flux increases were due to the combined porosity increases and
lar structure completely disappeared (Fig. 3e-2). No obvious change hydrophobicity improvements.
was observed in the cross-sectional membrane structure when the Membrane No. 8 possesses the best performance; the highest
exposure time was increased from 5 min to 10 min, as shown in permeate flux and about 100% NaCl rejection rate, and which was
Figs. 3c-1, d-1 and e-1, except that the particles were larger at the chosen to undergo long time VMD test. It was found that the per-
longer exposure time. meate flux and NaCl rejection rate nearly did not change in 4 h,
It was clearly observed that the large crystals were composed in other words, no wetting happened in membrane No. 8. But the
of many smaller crystals (Fig. 4f). On the surface of the microscale NaCl rejection rate decreased sharply for the commercial PVDF
aggregate spheres, we observed nanoscale multimolecular micellar membrane in the second hour, which is the result of wetting. The
spheres. The microstructured surface of these micellar aggregates difference of membrane NO. 8 and commercial membrane NO. 9 is
of PVDF is very similar to the micro-nanoscale binary structure of no hydrophilic pore-forming agents like LiCl and PEG were added
the lotus leaf [33]. This endowed the PVDF surfaces produced via in the casting solution when membrane No. 8 was prepared, and
VIPS with a bionic “lotus effect”. All of the bottom surfaces showed the pores in membrane were formed in VIPS process via PVDF crys-
the particulate structure illustrated in Fig. 5, which was beneficial tallization.
to hydrophobicity. The CAs were between 113◦ and 132◦ .
4. Discussions
3.2. Performance in membrane distillation
Hydrophobic PVDF membranes with water CAs greater than
The membranes described in Section 3.1 were used in VMD 140◦ were produced by controlling the exposure time in VIPS. The
experiments. The salt rejection rate was between 99.6% and 99.8% micro-nano bionic spheres and the low surface energy of the nano-
for all of the tested membranes. Fig. 6 shows how the trends of structure together contributed to the high CA on the PVDF surfaces.
changing permeate flux depended on the CA and the porosity of the Detail discussions were presented in this section.
membranes, indicating that the permeate flux generally increased
with increasing membrane hydrophobicity and porosity. When 4.1. Mass transfer and phase separation in exposure stage
the CA was increased from 84.5◦ to 142.5◦ and the porosity was
increased from 73.4% to 87.3%, the flux increased from 1800 g/m2 h The morphology change can attribute to the slow membrane
to 7000 g/m2 h. formation process in VIPS. For such a membrane formation process,
The permeate flux was not only affected by the hydrophobicity the membrane structure could form either in the humid air by
and porosity of the membranes but also by the pore diameter of the VIPS or in the coagulation bath via a phase separation induced by
membranes. Table 1 presents the pore diameters, porosities and liquid water. During the exposure to humid air, at a constant RH,
CAs of membranes when the cast film was exposed to the humid the chemical potential of water in the gaseous phase is constant;
air for different period time. The average pore diameter did not therefore, the water intake rates were similar for all casting
7876 Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881

Fig. 3. Dependence of the membrane cross-sectional structure on the exposure time that the PVDF/DMAC (15 wt.% PVDF) cast film was exposed to humid air (room
temperature, 100% RH). Exposure time: (a) 0 min, (b) 3 min, (c) 5 min, (d) 6 min, (e) 10 min. Magnification: 1k× (left column); 10k× (a-2 to d-2), 30k× (e-2) (right column).
Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881 7877

Fig. 4. Dependence of the membrane top surface structure on the exposure time that the PVDF/DMAc (15 wt.% PVDF) cast film was exposed to humid air (room temperature,
100% RH). Exposure time: (a) 0 min, (b) 3 min, (c) 5 min, (d) 6 min, and (e) 10 min. Magnification: 10k× (a). f is the magnification of the particle in e (100k×) The up-right
insert in each photos is the profile of a water droplet on the surface with a CA of (a) 84.5◦ ; (b)88.3◦ ; (c)107.3◦ ; (d)138.2◦ ; (e)142.5◦ .

solutions at the various exposure time. The content of absorbed reported as percentages of the initial mass of the casting solution.
water in the cast film thus depends only on the exposure time, At the beginning of the exposure step, it is clear that the predom-
which controls the phase separation mechanism. inant phenomenon was water absorption, leading to the observed
During the exposure stage, there are two types of mass transfer: mass increase. After the film reached the maximum mass, solvent
water absorption into the film by the hygroscopic solvent (DMAc) and water evaporation prevailed over the water sorption, thereby
and the loss of DMAc from the casting film due to evaporation. causing a decrease in weight. Finally, when all of the DMAc and
These two mass fluxes are evidenced by the differing mass varia- water that could be evaporated under the applied drying condi-
tions in the cast films with time (Fig. 7). The mass variations are tions were removed, the film mass was constant. We observed that

Table 1
Pore diameters and porosities of membranes when the film was exposed to humid air for different times (with a casting solution of 15% PVDF and 85% DMAc).

Membrane Exposure time to Water contact Average pore Porosity


number the humid air (min) angle (◦ ) diameter (␮m) (%)

1 0 84.5 0.11 73.4


2 1 86.2 0.11 74.1
3 2 88.3 0.11 82.7
4 3 90.8 0.11 83.4
5 4 107.3 0.11 84.2
6 5 138.2 0.11 85.4
7 6 140.1 0.11 87.1
8 10 142.5 0.11 87.3
9a 140.3 0.13 80.5
a
Commercial PVDF membrane.
7878 Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881

Fig. 5. Dependence of the membrane bottom surface structure on the exposure time that the PVDF/DMAC (15 wt.% PVDF) cast film was exposed to humid air (room
temperature, 100%RH). Exposure time: (a) 0 min, (b) 3 min, (c) 5 min, (d) 6 min, and (e) 10 min. Magnification: 10k×. The up-right insert in each photos is the profile of a
water droplet on the surface with a CA of 115.8◦ (a); 113.3◦ (b); 115.7◦ (c); 126.9◦ (d); 132.1◦ (e).

20
10
0
-10
-20
Mass variation(%)

-30
-40
-50
-60
-70
-80
-90
-100
0 5 10 15 20 25
time(h)

Fig. 7. Mass variation of the PVDF forming system as a function of exposure time.
The inset shows the mass variation during the first five hours.
Fig. 6. Influence of contact angle and porosity on the permeate flux in VMD.
Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881 7879

the final mass of the cast film was approximately the same when
the polymer content in the casting solution was constant.
The measured value for this absorption rate at 100% RH was
approximately 4.66 × 10−6 g/cm2 s. This value represents the dif-
ferences between the water intake and solvent evaporation fluxes.
During the mass decrease, the rate of mass variation became nearly
constant. The measured value for this rate 100% RH was approxi-
mately 1.20 × 10−6 g/cm2 s. Because the water content in the films
was low, the mass loss due to water evaporation is negligible com-
pared with that due to DMAc evaporation. The above absorption
rate plus the evaporation rate is the water intake rate, which was
5.86 × 10−6 g/cm2 s at 100% RH. From these values, we calculated
that the water intake was approximately 1.41 × 10−3 g/cm2 after
4 min of contact with the humid air at a RH of 100%. We know
that less than 5 wt.% water is required to induce phase separation
when the PVDF content is 10% in a casting solution at 25 ◦ C. A water Fig. 8. The dependence of the tensile strengths of the membranes on exposure time.
intake of 1.41 × 10−3 g/cm2 can thus cause phase separation in
2.82 × 10−2 g of solution per square centimeter of film, correspond-
ing to a region with a thickness of approximately 262 ␮m, based on
a casting solution density (PVDF 15%, DMAc 85%) of 1.085 g/cm3 . On
the basis of the above calculations, we concluded that with a 4 min
exposure and a RH of 100%, the water intake from the humid air
was sufficient to generate a phase separation over the entire film
cross-section.
When the exposure time was not more than 4 min, for example,
3 min, the absorbed water was insufficient to cause phase sepa-
ration in the entire film; the composition was in the immiscible
region. Once the film was immersed into the water bath, NG was the
predominant mechanism, resulting in a dense skin (Fig. 4b) and the
typical digitate bulk and cellular macrovoid walls (Fig. 3b-1). Once
the exposure time was more than 4 min, for example, 5–10 min,
the absorbed water was enough to cause phase separation in the
entire film. We did observe the film turned opaque when the expo-
sure time was longer than 4 min. Because mass transfer was slow
in exposure stage, with the water intake from air, PVDF had suf-
ficient time to crystallize. Crystallization demixing dominated in
VIPS, and the phase separation (crystallization) occurred due to the
Fig. 9. Effect of exposure time on melting curve pattern.
water absorbed from the humid air. This led to a porous skin and
small crystalline particle morphology (Fig. 3c-1, d-1 and e-1). With
an exposure time of 5 min, the PVDF crystals did not have suffi- for each membrane. It can be clearly seen that the crystallinity of
cient time to grow before the film was immersed in the water bath, membranes with longer exposure time show higher crystallinity
so they remained small (Figs. 3c-2 and 4c). When the exposure except the membrane with 3 min exposure time. This indicates that
time was increased, the crystal size also increased. Besides VIPS, long exposure time is favorable to crystallization in VIPS. Because
PVDF can also be crystallized in a “soft” coagulation bath, such as slower mass exchange between solvent and non-solvent will delay
water-DMAc [10], 1-octanol [34], or ethanol [35].
The residual cellular structure was formed of those PVDF
domains which had not crystallized by NG phase separation in the
immersion stage (Fig. 3c-2). Thus, a longer exposure time results
in less uncrystallized PVDF and therefore a less cellular structure
(Fig. 3d-2). A porous skin and small crystallite particle morphology
is highly detrimental to the mechanical strength of a membrane.
The tensile strengths of the membranes decreased sharply with
increasing exposure time, as shown in Fig. 8.

4.2. Crystallization and micro-nano bionic spheres

Crystallization plays a very important role in the membrane


morphology formation. Because the casting solution was prepared
at room temperature and the casting environment was also main-
tained at room temperature, crystallization by quenching was
impossible; in other words, TIPS did not occur. The membrane crys-
tallization behaviors were demonstrated by DSC (Fig. 9). It can be
seen that the membranes exhibit a similar melting behavior with
one broad melting peak at 155–158 ◦ C.
The total crystallinity was calculated by melt heat (Fig. 10). The
obtained data (Fig. 10) were average values of triplicate samples Fig. 10. Total crystallinities of PVDF membranes.
7880 Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881

Fig. 11. Effect of exposure time on WXRD.

liquid–liquid demixing, which might provide more time for crys- separation mechanism and the fast L–L demixing existed, the crys-
tallization process to yield the membrane, and crystallization is an tallinity was low.
important phase separation mechanism in VIPS.
WAXD measurements were performed to examine the crys- 4.3. Low surface energy
talline structures of the overall membranes. As shown in Fig. 11, the
membranes formed with different exposure time exhibit a similar During the exposure stage, with the evaporation of DMAc and
diffraction pattern except the membrane with 3 min exposure time. the adsorption of water, phase separation begins, and a lot of micro-
The predominant presence of ␣-phase is conformed by the occur- nano bionic spheres formed. Because of the low surface energy of
rence of two most distinctive diffraction peaks at around 18.3◦ and fluorocarbons, the fluorine component in PVDF molecules tends to
19.9◦ [36]. self-enrich on the outer surface of the micro-nano bionic spheres.
The total crystallinity was determined by deconvolution of Because the area of air–liquid intersurface on the top surface of
diffraction peaks into contributions from amorphous and crys- membrane is larger than that in the bulk, more fluorine com-
talline regions, following a method described in the literature ponent self-enrich on the top of the polymer film, resulting in
[37]. It is noted that the total crystallinity values measured by the low surface energy of the bionic polymeric film. We mea-
WAXD are lower than those from DSC measurements. The dis- sured the chemical composition near the PVDF membrane surface
crepancy between WAXD and DSC results may be associated with with X-ray photoelectron spectroscopy (XPS) at different depths
the presence of some imperfect crystals and part of the crystalline (Fig. 12), and the results showed that the ratio of the number of
amorphous interphase, which contribute to the DSC estimate but fluorine atoms ( CF2 ) to the total number of fluorocarbon atoms
are not probed by WAXD. This discrepancy may also because WXRD
is a tool analyzing the crystalline structure at the membrane sur-
face, while DSC is for the whole membrane; it may be possible that
for the membranes the crystallinities are different at the surface
and in the bulk; and the crystallinity at the surface is smaller than
in the bulk. Nevertheless, the trends of WAXD and DSC crystallinity
are in good agreement.
In addition, we can find that the membrane with 3 min exposure
time is quite different from the others in DSC and WXRD patterns
and the crystallinity determined from DSC and WXRD results. The
following may be the reason. For the membrane with 3 min expo-
sure time, the membrane forming mechanism is different from the
membranes without exposure time (0 min, NG mechanism) and
long exposure time (5–10 min, crystallization). When the film was
exposed to the humid air, at 3 min exposure time, its composition
was in the immiscible region, no phase separation happened. But
the water percentage in the film was higher than that of the film
without exposure time, that is to say the composition was closer to
phase separation line, and less time was needed to start NG phase
separation in coagulation bath. Fast L–L demixing led to a dense top
surface (Fig. 4b) and a typical digitate bulk and cellular macrovoid Fig. 12. Plot of the ratio of the number of fluorine atoms to the total number of
walls (Fig. 3b-1). Because crystallization was not the main phase fluorocarbon atoms against XPS detection angle.
Y. Peng et al. / Applied Surface Science 258 (2012) 7872–7881 7881

( CF2 and CH2 ) increased near top surface when the exposure [13] M. Khayet, J.I. Mengual, T. Matsuura, Porous hydrophobic/hydrophilic com-
time was 10 min, which showed that during exposure stage, fluo- posite membranes Application in desalination using direct contact membrane
distillation, Journal of Membrane Science 252 (2005) 101.
rine atoms had enough time to move to the top surface to low the [14] Z. Wang, F. Zheng, S. Wang, Experimental study of membrane distillation with
surface energy of membrane. brine circulated in the cold side, Journal of Membrane Science 183 (2001)
171.
[15] J. Zhang, J.D. Li, M. Duke, Z. Xie, S. Graya, Performance of asymmetric hollow
5. Conclusions fibre membranes in membrane distillation under various configurations and
vacuum enhancement, Journal of Membrane Science 362 (2010) 517.
The porous and spherulitic morphology resulting from PVDF [16] G.W. Meindersma, C.M. Guijt, A.B. de Haan, Water Recycling, Desalination
by air gap membrane distillation, Environmental Progress 24 (4) (2005)
crystallization favored a hydrophobic surface in VIPS. With the 434.
increase in exposure time, the phase separation process varied [17] J. Phattaranawik, A.G. Fane, A.C.S. Pasquier, W. Bing, F.S. Wong, Experimental
from L–L demixing to crystallization. The membrane morphology Study and design of a submerged membrane distillation bioreactor, Chemical
Engineering and Technology 32 (1) (2009) 38.
changed from typical large digitate macrovoids and dense and
[18] Y. Wu, Y. Kong, J. Liu, J. Zhang, J. Xu, An experimental study on membrane
smooth skin with a CA of approximately 90◦ (at less than 3 min of distillation -crystallization for treating waste water in taurine production,
exposure time) to a spherulitic morphology (at longer than 5 min Desalination 80 (1991) 235.
of exposure time). The CA reached 142.5◦ when the exposure time [19] Z. Wang, Z. Gu, S. Feng, Y. Li, Application of vacuum membrane distillation
to lithium bromide absorption refrigeration system, International Journal of
was 10 min. The micro-nano bionic spheres and the low surface Refrigeration 32 (2009) 1587.
energy of the nanostructure together contributed to the high CA [20] P.P. Zolotarev, V.V. Ugrozov, I.B. Volkina, V.M. Nikulin, Treatment of waste water
on the PVDF surfaces. The permeate flux increase in VMD was due for removing heavy metals by membrane distillation, Journal of Hazardous
Materials 37 (1994) 77.
to the combined effects of the increased porosity and improved [21] M.S. El-Bourawi, Z. Ding, R. Ma, M. Khayet, A framework for better understand-
hydrophobicity ing membrane distillation separation process, Journal of Membrane Science
285 (2006) 4.
[22] M. Khayet, Membranes and theoretical modeling of membrane distillation: a
Acknowledgments review, Advances in Colloid Interface Science, in press.
[23] B. Li, K.K. Sirkar, Novel membrane and device for direct contact membrane
distillation-based desalination process, Industrial and Engineering Chemistry
The authors would like to thank Professor Liu Zhongzhou of Research 43 (2004) 5300.
the Research Center for Eco-Environmental Science, the Chinese [24] B. Li, K.K. Sirkar, Novel membrane and device for vacuum membrane
Academy of Sciences for his valuable discussions and suggestions. distillation-based desalination process, Journal of Membrane Science 257
(2005) 60.
The authors also thank National Natural Science Foundation of
[25] C. Feng, R. Wang, B. Shi, G. Li, Y. Wu, Factors affecting pore structure and per-
China for financial support (21176008). formance of poly(vinylidene fluoride-co-hexafluoro propylene) asymmetric
porous membrane, Journal of Membrane Science 277 (2006) 55.
[26] N. Tang, Q. Jia, H. Zhang, J. Li, S. Cao, Preparation and morphological characteri-
References zation of narrow pore size distributed polypropylene hydrophobic membranes
for vacuum membrane distillation via thermally induced phase separation,
[1] H.C. Park, Y.P. Kim, H.Y. Kim, Y.S. Kang, Membrane formation by water vapor Desalination 256 (2010) 27.
induced phase inversion, Journal of Membrane Science 156 (1999) 169. [27] M. Khayet, T. Matsuura, J.I. Mengual, Porous hydrophobic/hydrophilic com-
[2] H.J. Lee, B. Jung, Y.S. Kang, H. Lee, Phase separation of polymer casting solution posite membranes: estimation of the hydrophobic layer thickness, Journal of
by non-solvent vapor, Journal of Membrane Science 245 (2004) 103. Membrane Science 266 (2005) 68.
[3] Y.S. Su, C.Y. Kuo, D.M. Wang, J.Y. Lai, A. Deratani, C. Pochat, D. Bouyer, Interplay [28] A. Gugliuzza, F. Ricca, E. Drioli, Controlled pore size, thickness and surface free
of mass transfer, phase separation, and membrane morphology in vapor- energy of super-hydrophobic PVDF and Hyflon AD membranes, Desalination
induced phase separation, Journal of Membrane Science 338 (2009) 17. 200 (2006) 26.
[4] C.L. Li, D.M. Wang, A. Deratani, D. Quémener, D. Bouyer, J.Y. Lai, Insight into the [29] M. Khayet, Membrane surface modification and characterization by X-ray
preparation of poly(vinylidene fluoride) membranes by vapor-induced phase photoelectron spectroscopy, atomic force microscopy and contact angle mea-
separation, Journal of Membrane Science 361 (2010) 154. surements, Applied Surface Science 238 (2004) 269.
[5] J.T. Tsai, Y.S. Su, D.M. Wang, J.L. Kuo, J.Y. Lai, A. Deratani, Retainment of pore [30] S. Bonyadi, T.S. Chung, Flux enhancement in membrane distillation by fabrica-
connectivity in membranes prepared with vapor-induced phase separation, tion of dual layer hydrophilic–hydrophobic hollow fiber membranes, Journal
Journal of Membrane Science 362 (2010) 360. of Membrane Science 306 (2007) 134.
[6] A. Ripoche, P. Menut, C. Dupuy, H. Caquineau, A. Deratant, Poly(ether imide) [31] C. Marega, A. Marigo, Influence of annealing and chain defects on the melting
membrane formation by water vapour induced phase inversion, Macromolec- behaviour of poly(vinylidene fluoride), European Polymer Journal 39 (2003)
ular Symposium 188 (2002) 37. 1713.
[7] H. Matsuyama, M. Teramoto, R. Nakatani, T. Maki, Membrane formation via [32] G. Capannelli, F. Vigo, S. Munari, Ultrafiltration membranes: characterization
phase separation induced by penetration of non-solvent from vapor phase. I. methods, Journal of Membrane Science 15 (1983) 289.
Phase diagram and mass transfer process, Journal of Applied Polymer Science [33] L. Feng, S. Li, Y. Li, H. Li, L. Zhang, J. Zhai, Y. Song, B. Liu, L. Jiang, D. Zhu, Super-
74 (1999) 159. hydrophobic surfaces from natural to artificial, Advanced Materials 14 (2002)
[8] X. Gao, L. Jiang, Water-repellent legs of water striders, Nature 432 (2004) 36. 1857.
[9] Q. Xie, J. Xu, L. Feng, L. Jiang, W. Tang, X. Luo, C. Han, Facile creation of a super- [34] T.H. Young, L.P. Cheng, D.J. Lin, L. Fane, W.Y. Chuang, Mechanisms of PVDF
amphiphobic coating surface with bionic microstructure, Advanced Materials membrane formation by immersion-precipitation in soft (1-octanol) and harsh
16 (2004) 302. (water) non-solvents, Polymer 40 (1999) 5315.
[10] M. Peng, H. Li, L. Wu, Q. Zheng, Y. Chen, W. Gu, Porous poly(vinylidene fluo- [35] L. Qian, Z.L. Xu, J. Lu, Effect of coagulation bath composition and temperature
ride) membrane with highly hydrophobic surface, Journal of Applied Polymer on morphology and property of microporous PVDF hydrophobic membrane,
Science 98 (2005) 1358. Journal of Chemical Engineering of Chinese Universities (Chinese) 24 (2) (2010)
[11] J. Zhang, N. Dow, M. Duke, E. Ostarcevic, J.D. Li, S. Gray, Identification of material 336.
and physical features of membrane distillation membranes for high perfor- [36] M. Zhang, A.Q. Zhang, B.K. Zhu, C.H. Du, Y.Y. Xu, Polymorphism in porous
mance desalination, Journal of Membrane Science 349 (2010) 295. poly(vinylidene fluoride) membranes formed via immersion precipitation pro-
[12] S. Al-Obaidani, E. Curcio, F. Macedonio, G.D. Profio, H. Al-Hinai, E. Drioli, Poten- cess, Journal of Membrane Science 319 (2008) 169.
tial of membrane distillation in seawater desalination: Thermal efficiency, [37] Y.S. Park, T. Hatae, H. Itoh, M.Y. Jang, Y. Yamazaki, High proton-conducting
sensitivity study and cost estimation, Journal of Membrane Science 323 (2008) Nafion/calcium hydroxyphosphate composite membranes for fuel cells, Elec-
85. trochimica Acta 50 (2004) 595.

Vous aimerez peut-être aussi