Sei sulla pagina 1di 7

Ultrasonics Sonochemistry 20 (2013) 232238

Contents lists available at SciVerse ScienceDirect

Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultson

Poly(vinylidene uoride) membranes by an ultrasound assisted phase inversion method


Mi-mi Tao, Fu Liu , Li-xin Xue
Ningbo Institute of Materials Technology & Engineering, Chinese Academy of Sciences, 519 Zhuangshi Road, Ningbo 315201, China

a r t i c l e

i n f o

a b s t r a c t
Poly(vinylidene uoride) (PVDF) membranes were prepared by an ultrasound assisted phase inversion process. The effect of ultrasonic intensity on the evolution of membrane morphology with and without the addition of pore former LiCl during precipitation process was comprehensively investigated. Besides the inter-diffusion between the solvent and nonsolvent, the ultrasonic cavitation was thought to have signicant inuences on phase inversion and the resultant membrane morphology. The mutual diffusion between water and solvent during the ultrasound assisted phase inversion process was measured. The crystalline structure was detected by wide angle X-ray diffractometer (WAXD). The thermal behavior was studied by differential scanning calorimeter (DSC). The mechanical strength, forward and reverse water ux, rejection to bovine serum albumin (BSA) and pepsin were also investigated. By the ultrasound assisted phase inversion method, ultra-ltration membrane was successfully prepared, which exhibited more preferable morphology, better mechanical property and more favorable permeability without sacricing the rejection and thermal stability. 2012 Elsevier B.V. All rights reserved.

Article history: Received 23 February 2012 Received in revised form 21 July 2012 Accepted 28 August 2012 Available online 5 September 2012 Keywords: PVDF membrane Ultrasound Morphology Phase inversion Cavitation

1. Introduction Poly(vinylidene uoride) membrane usually prepared by phase inversion is widely used in micro-ltration and ultra-ltration process due to its superior physical and chemical properties. Depending on the precipitation rate, PVDF membranes can be: (a) asymmetric, with a selective skin or large voids on a spongy sublayer or (b) symmetric, with an almost even porosity along the thickness. It has been investigated that many parameters inuence the precipitation process and ultimately the morphologies and the ltration performance of the membranes, including the composition of casting solution [14], precursor preparation [5,6], evaporation time [7], the harshness and temperature of the coagulation bath [811], etc. The phase separation behavior of PVDF/solvent/ H2O system is far more complex than amorphous polymer system due to the semi-crystal nature. Both liquidliquid demixing and liquidsolid (crystallization) demixing are usually occurred during the phase inversion process of PVDF membrane. These two mechanisms are developed to different extents according to the thermodynamics and mass transfer properties. Bottino [12] prepared PVDF membranes by casting and coagulating solutions of the polymer in eight solvents and found a good correlation between solventnonsolvent diffusivity and the membrane structure. Numerous researches also revealed that thermodynamics have less
Corresponding authors. Tel.: +86 574 86685256; fax: +86 574 86685186 (F. Liu), tel.: +86 574 86685831; fax: +86 574 86685186 (L.-x. Xue). E-mail addresses: fu.liu@nimte.ac.cn (F. Liu), xuelx@nimte.ac.cn (L.-x. Xue).
1350-4177/$ - see front matter 2012 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.ultsonch.2012.08.013

inuence on the structures and performance of the nal PVDF membranes, while the kinetic mass transfer is the key factor to control the membrane structure and formation [1315]. It is known that ultrasound can induce a wide range of chemical and physical consequences. In liquids irradiated with ultrasound, ultrasonic cavitation serves as the primary mechanism for sonochemical effects, where bubble collapse produces intense local heating, high pressures, microjet, turbulence, acoustic streaming, etc. [1618]. Physical effects of ultrasound include enhanced mass transfer, emulsication, bulk thermal heating and various effects on solids. Diverse applications of ultrasound have been explored. Nanostructured materials, nanometer colloids and protein microspheres can all be prepared by this general route [19,20]. Besides, ultrasound is evidenced as a powerful tool to polymerize [21], degrade polymer [22,23] and disperse nanoparticles [24,25], and it is widely used in the bulk or surface modication of polymers [26]. With respect to membranes for separation, ultrasound is usually utilized in membrane cleaning [27,28]. In consideration of the facilitation of ultrasound during physical or chemical process, it has the potential to assist the mass transfer between the solvent and nonsolvent, crystallization growth and structure evolution during phase separation of polymeric membranes. To the best of our knowledge, the role of ultrasound in phase inversion process of PVDF membrane has not been studied. Here, PVDF membranes were prepared by an ultrasound assisted phase inversion technique. The effect of ultrasonic intensity on membrane morphology was investigated. The crystalline structure, thermal stability,

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238

233

mechanical property and permeability of membranes were also reported in this paper. 2. Experimental 2.1. Materials Commercial PVDF (FR904) was purchased from 3F company (Shanghai, China) and was dried at 80 C in the oven for 24 h prior to use. N,N-dimethylfoemamide (DMF, P99.0%) and bovine serum albumin (BSA Fraction V, Mn = 67,000) were purchased from Sinopharm Chemical Reagent Co. Ltd., China and Sigma Aldrich, respectively. Lithium chloride (LiCl, P99.0%) and pepsin (Mn = 35,000) were purchased from Aladdin Reagent Co. Ltd., China. All materials were used as received. 2.2. Membrane preparation Fifteen percent by weight PVDF was dissolved in DMF at 80 C with a constant stirring rate for 24 h to obtain a homogenous casting solution. Moreover, six percent by weight LiCl based on the solution weight was added to the same PVDF solution and dissolved under the same condition. The casting solutions were kept still at 80 C for another 24 h and went through vacuum deaeration to remove air bubbles. The casting solutions were uniformly spread onto a glass plate by a casting knife with the thickness of 200 lm. The glass plate and casting knife were heated to 80 C before casting in order to minimize the effect of surrounding temperature on phase inversion. After the nascent lm was evaporated in the air for 10 s, it was transferred into the coagulation bath composed of deionized water in the ultrasonic instrument (40 Hz, KQ-300DB, Kunshan, China) and irradiated by ultrasound. The temperature of coagulation bath was kept constant at 28 C. In one minute, the membranes were automatically peeled off from the glass plate and taken out from the ultrasonic bath. The solidied membrane was nally transferred to the fresh deionized water for 24 h to remove the residual solvents. Therefore, the thermal effect of ultrasonic irradiation on the coagulation bath could be ignored in one minute. The solidied membrane was nally transferred to the fresh deionized water for 24 h to remove the residual solvents. Subsequently, the membranes were dried at room temperature. The preparation conditions for various membranes are summarized in Table 1. 2.3. Membrane characterization Morphological structures of the prepared PVDF membranes were examined by scanning electron microscope (SEM, S-4800, Hitachi, Japan). The cross section samples were fractured in liquid nitrogen. Both the surface and cross section of the samples were sputtered gold for 2 min for observation. To determine the mutual diffusion between water and solvent during the membrane formation process via phase inversion, the water and DMF contents in the membranes at a xed time after

immersion into the ultrasonic bath were measured as follows: the nascent membrane was immersed in the coagulation bath for 5 s and then taken out to wipe off water drop adhered to the membrane surface, afterwards, the wet membranes were transferred to specic amount (mmethanolwater ) of methanol (to extract water from the membrane) and deionized water (to extract DMF from the membrane), respectively and then preserved under seal. The lixivium was analyzed by coulometer (KF831, Metrohm, Switzerland) and UVVISNIR spectrometer (Lambda 950, Perkin Elmer, US) at 197 nm, respectively to get the concentration of water (cH2 O ) and DMF (cDMF ). Then the wet membranes should be dried to obtain dry weight (mdry ) and the contents of water and DMF (M waterDMF ) can be calculated using the following equation:

MwaterDMF

C H2 ODMF mmethanolwater mdry

Crystalline structure of the membrane was analyzed by wide angle X-ray diffractometer (WAXD, D8 Advance, Bruker, Germany). All samples were detected in a continuous scan mold between 5 and 60 2h with the scanning speed of 0.2 s/step. The thermal behavior of the membrane was determined by differential scanning calorimeter (DSC) (Pyris Diamond, Perkin Elmer, US). The temperature was raised from 25 to 250 C at a rate of 10 C/min at a nitrogen atmosphere. Crystallinity of the membrane was calculated by melting heat based on the fusion enthalpy of ideal PVDF crystal DHf 0 105 J=g:

Crystallinity

DHsample DH 0 f

where, DHsample is the melting heat of the sample obtained by DSC curve. Mechanical properties of the membranes were determined using a tensile tester (5567, Instron, US) with the stretching rate of 5 mm/min at room temperature. Each sample was cut into 10 1 cm, and the thickness of the sample was measured according to the SEM pictures. Pure water ux of PVDF membranes were measured by a cross ow ltration system (Saifei company, China) (the effective membrane area = 24 cm2). The measuring protocol was as follows: In the rst 40 min, the membrane was compacted at 0.15 MPa to diminish the compaction effect and obtain the stable ux, and then the ux was measured at 0.1 MPa. Water ux reported in this work was the nal steady ux. Afterwards, pure water was changed to 1 g/L BSA or pepsin solution. The protein concentration of both feed and permeate solutions were examined by UVVISNIR spectrometer (Lambda 950, Perkin Elmer, US) at 280 nm (BSA) or 270 nm (pepsin). The rejection to protein (R (%)) of the membranes was calculated using the following equation [29]:

  A1 100 R% 1 A0

Where A1 is the absorption value of the permeate solution, and A0 is the absorption value of the feed solution.

Table 1 Preparation conditions for PVDF membranes and diffusion content of water and DMF at xed time 5s. Code M1 M2 M3 MA1 MA2 MA3
a b

LiCl/PVDF/DMF (wt/wt/wt) 0/15/85 0/15/85 0/15/85 6/15/85 6/15/85 6/15/85

Ultrasonic intensity (W) 0 180 300 0 180 300

Water contenta (g/g) 0.392 0.673 0.545

DMF contenta (g/g) 3.263 2.515 1.675

DMF diffusion amounta,b (g/g) 2.404 3.152 3.992

Water or DMF content in one gram of PVDF membrane. Calculated based on 15 wt.% solid content: 85 15 M DMF .

234

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238

3. Results and discussion 3.1. Membrane morphologies 3.1.1. Morphologies of membranes prepared without additives Fig. 1 shows the morphologies of M1, M2 and M3. All membranes exhibit spongy sublayer, which can be seen in the cross section in Fig. 1a and b. However, the membrane cross section close to the top surface displays different structure with the variation of ultrasonic intensity. Membrane M1 without ultrasonic assistance

has a dense skin layer with the thickness about 8 lm, the possible reason could be that the faster outow of DMF and slower inow of water result in the higher PVDF concentration on the membrane top layer during the phase inversion process. Increasing the ultrasonic intensity to 180 W, membrane M2 exhibited regular teardrop-like voids which develop from the top surface. When the ultrasonic intensity was further raised to 300 W, cellular pores with a graded distribution through the whole cross section are clearly presented and the dense skin layer or teardrop-like voids are totally eliminated in case of membrane M3.

Fig. 1. SEM images of cross section, bottom surface and top surface morphology of M1, M2 and M3: (a) the cross section with magnication 400; (b) the enlarged cross section with magnication 2000; (c) the bottom surface with magnication 5000; (d) the top surface with magnication 20,000.

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238

235

Usually, pseudo binary (solvent and nonsolvent) diffusion theory is adopted to describe the early phase separation stage of membrane formation [30], which reproduces with good agreement the experimental data and observations. There is evidence that a high diffusion rate of nonsolvent gives rise to membrane with large voids beneath the top surface [15]. Based on this and the relevant morphology variation of the membrane cross section close to the top skin layer, it can be inferred that the water diffusion rate into the PVDF/DMF nascent lm during the incipient phase separation moments following the order: M2 > M3 > M1. The contents of water and DMF in the membranes at a xed time 5 s after immersion were measured to predict the mutual diffusion rate between water and DMF. As depicted in Table 1, water content in the membrane, representing the water diffuse rate from the coagulation bath to the membrane, reaches a maximum (0.673 g/g) when the ultrasonic intensity is 180 W corresponding to M2, thus leading to the formation of teardrop-like voids on the top skin layer. The diffusion amount of DMF into the coagulation bath calculated based on the initial composition of the casting solution and the residual quantity progressively increases with increasing the ultrasonic intensity. It is well known that ultrasound can promote the molecule movement and therefore increase the diffusion rate of DMF. However, ultrasonic irradiation with 300 W intensity does not further accelerate water diffusion as expected (water content: 0.545 g/g). The exceptional diffusion behavior may be the result of stronger hydrogen bonding between the water molecules and the carbonyl group of the DMF molecules. Petersen has pointed out that hydrogen bonding interaction between water molecules and carbonyl oxygen are stronger than between water molecules [31]. In case of DMF, a special amide in view of the lack of hydrogen bonding in the pure solvent, this effect may be strengthened by the nitrogen atom due to the resonance forms [32]. Ultrasonic irradiation promotes the mass transfer of DMF from higher concentration zone to lower concentration zone, while the diffusion of water was not only affected by the ultrasonic intensity, but also affected by the hydrogen bonding interaction between water and DMF. Therefore, despite the high-intensity ultrasound (300 W), water diffusion rate in the early stage of M3 is still lower than that of M2, leading to the formation of a uniform structure with cellular pores. The morphology of the bottom surface of the membrane also varies with the variation of ultrasonic intensity used. As Fig. 1c presents, M1 has a dense bottom surface composed of syncretic spherulites, while M2 and M3 exhibit porous surface with massive inter-connected pores. As we all know, the bottom surface contacted with the glass plate is far from the membrane/water interface, therefore, the water/DMF diffusion has less effect on structure evolution during phase inversion. The main factor is the ultrasonic cavitation, which produces high-energy phenomenon and is responsible for the generation of surface damage [33]. We believe that the impingement of shockwaves and microjets on the surface creates the localized erosion. With the increase of ultrasonic intensity, erosion becomes more severe and nally leads to the formation of completely porous surface. All membranes present a dense top surface as can be shown in Fig. 1d. The delayed phase separation leads to the evaporation of solvent and higher polymer concentration. After the casted nascent lm on the glass plate is immersed into deionized water, the fast mass transfer at the interface between water and lm results in instantaneous solidication of top surface, therefore, a dense top surface was formed in all cases. 3.1.2. Morphologies of membranes prepared with LiCl The effect of ultrasonic irradiation on morphologies of membranes prepared with pore former can also be identied. When 6% by weight LiCl is used, all the membranes exhibit highly inhomogeneous structures due to the presence of large voids of differ-

ent size and shape through the cross section as shown in Fig. 2a and b. In membrane MA1, arrestive nger-like pores and compressed cavities simultaneously appear in the porous sublayer. For MA2, the compressed cavities are evolved into isolated pores some of which become interlinked with the nger-like pores close to the top surface owing to the turbulence of ultrasonic irradiation of 180 W. The images of enlarged cross section in Fig. 2b demonstrate that the nger-like pores in MA2 are suppressed to some extent compared with those in MA1, caused by the turbulence of ultrasound during phase inversion. Unlike the membrane without additives, the presence of LiCl in the dope solution promotes the diffusion of water to the PVDF dope due to the good afnity between them. Moreover, the increased dope viscosity hinders the outow of DMF from the polymer dope. Therefore, the evolution of nger-like pores cannot be totally eliminated. In case of MA3, nger-like pores are seen to grow from the top surface almost to the bottom surface. On the one hand, the high intensity ultrasonic irradiation may exacerbate the formation of nger-like pores by enhancing the water diffusion rate. On the other hand, with increasing the intensity of ultrasonic irradiation, the smaller isolated pores conquered the interface tension barrier and were gradually connected together to form the long nger-like pores through the cross section. The large cavities beneath nger-like pores as shown in MA1 derive from the polymer lean phase favorable in growth. With the assistance of ultrasound, the coalescence between these polymer lean phases and surrounding small-scale polymer lean phase gets easier and consequently forms the fully developed large isolated cavities in MA2. With further improving the ultrasonic intensity, the interface barrier between nascent nger-like pores and large cavities below was eliminated due to the turbulence of the ultrasound, resulting in the formation of long nger-like pores. The bottom and top surface of the membranes are illustrated in Fig. 2c and d, respectively. As expected, the ultrasonic irradiation made the bottom surface more porous. There are few discrete circular pores appearing on the mainly dense surface of MA1, while MA2 presents the most porous bottom surface with uniform pore size distribution when the ultrasonic intensity is 180 W. Unexpectedly, the highest ultrasonic intensity (300 W) does not necessarily improve the surface porosity further. As discussed above, numerous bubbles are formed in the viscous dope solution and then collapse to produce microscopic jets, which produces localized erosion in the bottom surface and nally leads to the formation of porous morphology. However, solidication rate may also have strong impact on the resulting surface morphology when ultrasound is used during phase inversion. The abundant nger-like cavities are displayed across most of the cross section in MA3, which reduces water transfer resistance and subsequently accelerates solidication. Due to the limitation of ultrasonic energy, the bottom surface undergoing faster solidication cannot be badly eroded. Resultantly, MA3 has a less porous bottom surface than MA2. The three membranes exhibit relatively dense top surface because of the air exposure and high polymer concentration in the membrane surface. In all cases nanoscale pores arise as LiCl is used as a pore former, which can be clearly seen from the high resolution image inset in MA3-d in Fig. 2. Based on the morphological observation, it is concluded that ultrasonic irradiation may facilitate the elimination and formation of macrovoids, which depends on the ultrasonic intensity applied during phase inversion. By this simple way, completely uniform structure (M3) without additives and typical asymmetric structure with pore formers in the bulk (MA3) can be obtained. Ultrasound contributes a lot to make the bottom surface porous regardless of the addition of pore former. However, the top surface was little inuenced by the ultrasound due to the air exposure and fast solidication.

236

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238

Fig. 2. SEM images of cross section, bottom surface and top surface morphology of MA1, MA2 and MA3: (a) the cross section with magnication 400; (b) the enlarged cross section with magnication 1300; (c) the bottom surface with magnication 5000; (d) the top surface with magnication 20,000; image inset in MA3-d: top surface of MA3 with magnication 100,000.

3.2. Membrane properties Fig. 3 presents the WAXD patterns of the PVDF membranes. Membranes prepared from the same dope solution have the similar crystalline structure. The occurrence of the distinctive peaks at 18.5 and 26.7, corresponding to the reections of a(0 2 0) and a(0 2 1), conrms the predominant presence of a phase in M1, M2 and M3. Despite the small shoulder at 18.5 accompanying the superimposed peaks b(2 0 0) and b(1 1 0) at 20.8, MA1, MA2 and MA3 are dominated by b phase due to the added LiCl.

Thermal stability of PVDF membranes were characterized by DSC measurements. Despite the signicant morphological difference in these membranes, the corresponding thermograms present minute difference as can be seen in Fig. 4. Crystallinity calculated by melting heat based on the fusion enthalpy of ideal PVDF crystal DH0 f 105 J=g and melting temperature are listed in Table 2. All membranes exhibit almost the same melting temperature around 162 C and membranes cast from the same dope solution have similar crystallinity. Nevertheless, the addition of LiCl to the dope solution visibly promotes crystallization, for example, the crystallinity

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238

237

240 forward flux reverse flux


MA3

200

Intensity (a.u.)

MA2 MA1 M3 M2 M1

Pure water flux (L/m2h)


40 50

160

120

80

10

20

30

40 MA1 MA2 MA3

2 (degree)
Fig. 3. X-ray diffractograms for different membranes: M1, M2, M3, MA1, MA2, and MA3.

Membrane
Fig. 5. Permeability of PVDF membranes: forward ux (water ows from top surface to bottom surface), reverse ux (water ows from bottom surface to top surface).

MA3 MA2

MA1 M3 M2 M1

80

120

160

200
O

240

Temperature ( C)
Fig. 4. DSC thermograms for different membranes: M1, M2, M3, MA1, MA2, and MA3.

has the most excellent tensile strength of 7.21 MPa. The tensile strength of M2 and M3 are 6.21 and 6.19 MPa, respectively due to the employment of ultrasound. When LiCl is used as the pore former, membranes exhibit much lower tensile strength than those without LiCl. Ultrasound assisted phase inversion generates membranes with a little higher tensile strength, opposite to the situation in non-additive system. Tensile strength may be related to the morphological structure which is varying with ultrasonic intensity used. M1 has dense skin layers and is less porous than M2 and M3, resulting in the highest tensile strength. The addition of LiCl remarkably promotes the formation of large voids in membranes MA, which consequently conduces to the reduction of tensile strength to less than 4 MPa. The elongation may mainly depend on the crystallinity of the membranes. M1, M2 and M3 have lower crystallinity and therefore have higher elongation (in the range of 83% to 90%), while higher crystallinity of MA1, MA2 and MA3 leads to lower elongation (around 60%). 3.3. Membrane performances Permeability of membranes was measured by pure water ux as can be seen in Fig. 5. Water ux reects the pore size, pore size distribution and morphologies of the membranes. The permeate ux of M1, M2 and M3 was too low to be measured at 0.1 MPa due to the dense top surface of the membrane. While, the permeability of MA2 and MA3 prepared with the assistance of ultrasound is superior to that of MA1. The steady ux of MA2 and MA3 reaches up to 115 L/m2 h and 120 L/m2 h, respectively while MA1 shows the lower ux of 80 L/m2 h. It is because that MA2 and MA3 have sufcient nger-like pores through the cross section and porous bottom surface. The reverse ux of the three membranes was tested (water ows from bottom surface to top surface). As also

60.7% of MA1 is higher than that of M1 51%. LiCl may act as a nucleating agent besides pore former and hence is benecial to the crystallization of PVDF chains during phase inversion. It can be seen that the effect of ultrasonic intensity on crystalline behavior is faintness, which can be ascribed to the concentrated dope solution. The rearrangement and crystallization of PVDF are almost not inuenced due to the high entanglement of macromolecules chains. Therefore, crystalline structure and thermal stability of the membranes are practically unchanged with the variation of ultrasonic intensity. The mechanical properties were also investigated (listed in Table 2). Among the membranes prepared without additive, M1

Table 2 Properties of PVDF membranes. Code M1 M2 M3 MA1 MA2 MA3 Crystallinity (%) 51.0 53.6 51.8 60.7 60.1 61.2 Tm (oC) 161.99 162.85 161.98 162.18 162.34 162.19 Tensile strength (MPa) 7.17 0.22 6.21 0.39 6.19 0.12 3.04 0.37 3.88 0.34 3.86 0.24 Elongation (%) 90 12 88 15 83 13 63 16 66 11 67 12 BSA rejection (%) 98.9 99.1 98.9 Pepsin rejection (%) 81.4 79.4 79.8

Exotherm

238

M.-m. Tao et al. / Ultrasonics Sonochemistry 20 (2013) 232238 [6] X. Wang, X.Y. Wang, L. Zhang, Q.F. An, H.L. Chen, Morphology and formation mechanism of Poly(Vinylidene Fluoride) membranes prepared with immerse precipitation: effect of dissolving temperature, J. Macromol. Sci. Part B: Phys. 48 (2009) 696709. [7] S. Munari, A. Bottino, G. Capannelli, Casting and performance of polyvinylidene uoride based membranes, J. Membr. Sci. 16 (1983) 181193. [8] L.-P. Cheng, Effect of temperature on the formation of microporous PVDF Membranes by precipitation from 1-octanol/DMF/PVDF and water/DMF/PVDF systems, Macromolecules 32 (1999) 66686674. [9] P. Sukitpaneenit, T.-S. Chung, Molecular elucidation of morphology and mechanical properties of PVDF hollow ber membranes from aspects of phase inversion, crystallization and rheology, J. Membr. Sci. 340 (2009) 192 205. [10] M.G. Buonomenna, P. Macchi, M. Davoli, E. Drioli, Poly(vinylidene uoride) membranes by phase inversion: the role the casting and coagulation conditions play in their morphology, crystalline structure and properties, Eur. Polymer J. 43 (2007) 15571572. [11] X. Wang, L. Zhang, D. Sun, Q. An, H. Chen, Formation mechanism and crystallization of poly(vinylidene uoride) membrane via immersion precipitation method, Desalination 236 (2009) 170178. [12] A. Bottino, G. Camera-Roda, G. Capannelli, S. Munari, The formation of microporous polyvinylidene diuoride membranes by phase separation, J. Membr. Sci. 57 (1991) 120. [13] D.-J. Lin, C.-L. Chang, F.-M. Huang, L.-P. Cheng, Effect of salt additive on the formation of microporous poly(vinylidene uoride) membranes by phase inversion from LiClO4/Water/DMF/PVDF system, Polymer 44 (2003) 413422. [14] Y.S. Soh, J.H. Kim, C.C. Gryte, Phase behaviour of polymer/solvent/non-solvent systems, Polymer 36 (1995) 37113717. [15] T.H. Young, L.P. Cheng, D.J. Lin, L. Fane, W.Y. Chuang, Mechanisms of PVDF membrane formation by immersion-precipitation in soft (1-octanol) and harsh (water) nonsolvents, Polymer 40 (1999) 53155323. [16] S.J. Doktycz, K.S. Suslick, Interparticle collisions driven by ultrasound, Science 247 (1990) 10671069. [17] E.B. Flint, K.S. Suslick, The temperature of cavitation, Science 253 (1991) 1397 1399. [18] F.R. Young, Cavitation, Imperial College Press, London, 1999. [19] K.S. Suslick, M.W. Grinstaff, Protein microencapsulation of nonaqueous liquids, J. Am. Chem. Soc. 112 (1990) 78077809. [20] J.H. Bang, K.S. Suslick, Applications of ultrasound to the synthesis of nanostructured materials, Adv. Mater. 22 (2010) 10391059. [21] H.X. Xu, K.S. Suslick, Sonochemical preparation of functionalized graphenes, J. Am. Chem. Soc. 133 (2011) 91489151. [22] H. Kim, J.W. Lee, Effect of ultrasonic wave on the degradation of polypropylene melt and morphology of its blend with polystyrene, Polymer 43 (2002) 2585 2589. [23] Q.Y. Li, G.Z. Wu, Y.L. Ma, C.F. Wu, Grafting modication of carbon black by trapping macroradicals formed by sonochemical degradation, Carbon 45 (2007) 24112416. [24] J. Wang, S.J. Severtson, A. Stein, Signicant and concurrent enhancement of stiffness, strength, and toughness for parafn wax through organoclay addition, Adv. Mater. 18 (2006) 15851588. [25] S.K. Swain, A.I. Isayev, Effect of ultrasound on HDPE/clay nanocomposites: rheology, structure and properties, Polymer 48 (2007) 281289. [26] G.J. Price, F. Keen, A.A. Clifton, Sonochemically-assisted modication of polyethylene surfaces, Macromolecules 29 (1996) 56645670. [27] M. Cai, S.N. Zhao, H.H. Liang, Mechanisms for the enhancement of ultraltration and membrane cleaning by different ultrasonic frequencies, Desalination 263 (2010) 133138. [28] M. Kallioinen, M. Manttari, Inuence of ultrasonic treatment on various membrane materials: a review, Sep. Sci. Technol. 46 (2011) 1388 1395. [29] M.M. Tao, F. Liu, L.-x. Xue, Hydrophilic poly(vinylidene uoride) (PVDF) membrane by in situ polymerisation of 2-hydroxyethyl methacrylate (HEMA) and micro-phase separation, J. Mater. Chem. 22 (2012) 91319137. [30] L. Yilmaz, A.J. McHugh, Modelling of asymmetric membrane formation. I. critique of evaporation models and development of a diffusion equation formalism for the quench period, J. Membr. Sci. 28 (1986) 287310. [31] R.C. Petersen, Interactions in the binary liquid system n,n-dimethylacetamide water viscosity and density, J. Phys. Chem. 64 (1960) 184185. [32] C. De Visser, G. Perron, J.E. Desnoyers, W.J.M. Heuvelsland, G. Somsen, Volumes and heat capacities of mixtures of N,N-dimethylformamide and water at 298.15 K, J. Chem. Eng. Data 22 (1977) 7479. [33] K.S. Suslick, G.J. Price, Applications of ultrasound to materials chemistry, Annu. Rev. Mater. Sci. 29 (1999) 295326.

shown in Fig. 5, in all cases, reverse ux is higher than forward ux (water ows from top surface to bottom surface) since the bottom surface is more porous than the top surface. MA2 exhibits the highest ux, nearly twice the ux of MA1. The reverse ux decreases with the following order: MA2 > MA3 > MA1, conrming the difference of the bottom surface micro-structure. The rejection to BSA and pepsin of membranes prepared by the ultrasound assisted phase inversion process is listed in Table 2. All of them have almost similar rejection to BSA (about 99%) and pepsin (about 80%), indicating the potential application of membranes in ultra-ltration process. 4. Conclusions Ultrasonic irradiation during phase inversion process had a great effect on morphologies of PVDF membranes. Increasing ultrasonic intensity, morphology of upper layer in membranes cast from non-additive solution changed dramatically from original dense skin to regular large voids and then to uniform cellular pores. When LiCl was used as the pore former, nger-like pores were suppressed and then aggravated. Furthermore, the unfavorable cavities in membrane bulk were eliminated. All results revealed that ultrasonic irradiation plays an important role in the formation of cross section morphology during the phase inversion regardless of the addition of pore formers. The porous structure of bottom surface was supposed to be mainly dominated by the ultrasonic cavitation. However, higher ultrasonic intensity (300 W) will not further improve the porosity. Moreover, little change was observed in the top surface of membranes. The tensile strength and permeability mainly depended on the morphological variation. The ultrasonic irradiation was found to have slight inuence on the crystalline structure, thermal stability and tensile elongation. Of all membranes, MA3 prepared with the assistance of 300 W ultrasonic irradiation showed the highest water ux of 120 L/m2 h and the rejection of 98.9% to BSA and the rejection of 79.8% to pepsin. Acknowledgements We are grateful for the nancial support from the National Natural Science Foundation of China (51273211), an international cooperation project from Ministry of Science and Technology of China (2012DFR50470), the National 863 Foundation of China (2012AA03A605), and the welfare technology application research project (2011C31002). References
[1] A. Mansourizadeh, A.F. Ismail, Effect of LiCl concentration in the polymer dope on the structure and performance of hydrophobic PVDF hollow ber membranes for CO2 absorption, Chem. Eng. J. 165 (2010) 980988. [2] C. Mu, Y. Su, M. Sun, W. Chen, Z. Jiang, Remarkable improvement of the performance of poly(vinylidene uoride) microltration membranes by the additive of cellulose acetate, J. Membr. Sci. 350 (2010) 293300. [3] Z. Wang, J. Ma, Q. Liu, Pure sponge-like membranes bearing both high water permeability and high retention capacity, Desalination 278 (2011) 141149. [4] L. Shi, R. Wang, Y. Cao, D.T. Liang, J.H. Tay, Effect of additives on the fabrication of poly(vinylidene uoride-co-hexauropropylene) (PVDF-HFP) asymmetric microporous hollow ber membranes, J. Membr. Sci. 315 (2008) 195204. [5] D.-J. Lin, K. Beltsios, T.-H. Young, Y.-S. Jeng, L.-P. Cheng, Strong effect of precursor preparation on the morphology of semicrystalline phase inversion poly(vinylidene uoride) membranes, J. Membr. Sci. 274 (2006) 6472.

Potrebbero piacerti anche