You are on page 1of 9

Separation and Purification Technology 211 (2019) 578–586

Contents lists available at ScienceDirect

Separation and Purification Technology


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

Congo red dye removal by direct membrane distillation using PVDF/PTFE T


membrane
N.P. Khumaloa, L.N. Nthunyaa,b, E. De Canckb, S. Dereseb, A.R. Verliefdeb, A.T. Kuvaregaa,
⁎ ⁎
B.B. Mambaa, S.D. Mhlangaa, , D.S. Dlaminia,c,d,
a
Nanotechnology and Water Sustainability (NanoWS) Research Unit, College of Science, Engineering and Technology, University of South Africa, Florida, 1709
Johannesburg, South Africa
b
Particle and Interfacial Technology Group, Department of Applied Analytical and Physical Chemistry, Ghent University, Coupure Links 653, 9000 Ghent, Belgium
c
State Key Laboratory of Separation Membranes and Membrane Processes/National Center for International Joint Research on Membrane Science and Technology, Tianjin
300387, PR China
d
School of Materials Science and Engineering, Tianjin Polytechnic University, Tianjin 300387, PR China

A R T I C LE I N FO A B S T R A C T

Keywords: The ability to use the membrane distillation (MD) technique is envisaged as a promising approach to attain
Congo red dye sustainable and reliable clean water supply. In this work, polytetrafluoroethylene/polyvinylidene fluoride
Methyl functionalized silica nanoparticles (PVDF/PTFE) flat sheet membranes were fabricated via the thermally induced solvent evaporation process. The
PVDF/PTFE nanocomposite membranes PVDF/PTFE membrane surface was modified by incorporating methyl functionalized mesoporous silica nano-
Direct contact membrane distillation
particles (MfSNPs). Prior to application, the membranes were characterized with respect to surface and struc-
tural morphology, hydrophobicity and overall porosity. Clean water flux measurements were conducted using a
direct contact membrane distillation (DCMD) lab-scale experimental setup with deionized water as the feed
solution (50 °C) and permeate solution (20 °C). The highest stable pure water flux for the MfSNPS/PVDF/PTFE
membranes was 0.0041 L/h m2. The incorporation of the MfNPS did not only improve the fluxes but also induced
low wetting properties as shown by the contact angle and LEPw values. The MfSNPs/PVDF/PTFE membranes
were highly efficient in removing Congo red dye from water with 99% removal efficiency achieved.

1. Introduction hydrophobic membranes in MD processes [5,11,12]. However, other


studies implement post-surface modification on these commercially
Membrane distillation (MD) technology has emerged as one of the available membranes to improve their performance [13,14]. In recent
potential membrane separation processes for production of potable studies ceramic membranes have been used in MD [14–16]. Results
water from saline water. This technique can also be integrated in water/ from these studies have shown that the application of ceramic mem-
wastewater treatment processes for recovery of nutrients and metals branes in MD may result in higher separation performance due to their
[3–5], removal of dyes [4,5], and effluent treatment for the production appealing properties such as: high mechanical stability, chemical and
of clean water in industrial processes [1,6]. MD is a non-isothermal thermal resistance and the ability to control pore geometry. However,
membrane-based separation process [8]. This technique is based on the ceramic membranes tend to be hydrophilic due to the existence of hy-
diffusive and convective transportation of vapour across a hydrophobic droxyl groups on membrane surfaces, therefore surface modification is
membrane [9]. The hydrophobic membrane serves as a barrier between required to make them hydrophobic, thus increasing fabrication costs
two non-isothermal solutions. The driving force across the membrane is [17]. As such, polymeric membranes have remained popular for ap-
the partial vapour pressure gradient resulting from the temperature plication in MD processes.
difference between the bulk feed and permeate solutions [1,6,7,9]. In recent years, there has been a growing interest to fabricate
Previous research on MD has established that polytetra- membranes with improved properties for MD applications. Current
fluoroethylene (PTFE), polyvinylidene fluoride (PVDF) and poly- research is focused on fabricating MD membranes with high thermal
propylene (PP) are ideal polymeric materials for this application [8]. efficiency to prevent conductive heat loss across the membrane surface
Some studies have reported the application of commercially available [8,18,19]. The low heat conductivity of polymeric membranes


Corresponding authors at: School of Materials Science and Engineering, Tianjin Polytechnic University, Tianjin 300387, PR China (D.S. Dlamini).
E-mail address: derricksdlamini@gmail.com (D.S. Dlamini).

https://doi.org/10.1016/j.seppur.2018.10.039
Received 28 November 2017; Received in revised form 6 September 2018; Accepted 17 October 2018
Available online 17 October 2018
1383-5866/ © 2018 Elsevier B.V. All rights reserved.
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

compromises the mass transfer across the membrane resulting in low functionalization, the mixture was filtered and washed several times
fluxes [8]. The hydrophobic MD membrane serve as a barrier main- with pentane. The resultant material denoted as MfSNPs, was dispersed
taining a liquid/vapour interface formed across the membrane [16]. in pentane and the mixture was vigorously stirred for 1 h at 30 °C to
The liquid/vapour interface maintenance is a function of both physical wash-off all the unreacted HMDS from the nanoparticles. The methyl
and chemical properties of these popularized polymeric membranes. functionalized silica nanoparticles (MfSNPs) were filtered and vacuum
The most crucial characteristics for determining membrane applic- dried at 120 °C. Infrared spectroscopic measurements and nitrogen
ability and efficiency in MD processes are hydrophobicity, surface sorption were performed to confirm the successful functionalizing
roughness, thickness, porosity and pore geometry [6]. Membrane por- through functional (methyl) group identification.
osity and thickness control the mass transport of water vapor across the
membrane [20]. Enhanced membrane hydrophobicity and controlled 2.4. Membrane fabrication
pore shape and size formation results in membranes with low vapour
resistance therefore achieving a high long-term permeate flux [21]. The The membranes were fabricated following a modified method re-
hydrophobic nature of the membrane prevents penetration of the ported by Dong et al., 2014 [25]. Powdered PVDF was dissolved in N-
aqueous solution into the pores, while pore sizes and shapes determine methyl-2-pyrrolidone (NMP) under vigorous magnetic stirring, then
the diffusion or convection across the membrane. However, membrane different PTFE concentrations (0, 3, 6 wt%) were added into the dis-
flux is also affected by operational conditions such as heat gradient solved PVDF solution to obtain the viscosity that allows casting. The
across membrane and direct contact membrane distillation (DCMD) mixture was vigorously stirred for 48 h, at 50 °C. The casting solution
module design (surface area) [21]. Hydrophobicity and pore geometry was degassed and cast using a casting knife, controlling the thickness of
influence the magnitude of liquid pressure entry. the membranes to 0.20 mm for membranes without nanoparticles. For
Recently, functionalizing mesoporous silica nanoparticles with me- the synthesis of nanocomposite membranes a double casting method
thyl groups has gained interest for use in MD membranes. This is mainly reported by Mahlangu et al., 2017 was used [26]. In this method, so-
because the organic functional groups replace the hydrophilic methyl lution A (solution without the nanoparticles) was cast using a casting
groups thus resulting in more ‘super-hydrophobic’ silica nanoparticles knife with the gap clearance of 0.18 mm then a second solution (solu-
[22,23]. In this study the main objectives were: (1) to evaluate the tion B, containing nanoparticles (0.3 and 0.5 wt%)) was cast on top of
effect of methyl functionalized silica nanoparticles (MfSNPs) on the the first layer using a casting knife with the gap clearance 0.02 mm. The
structural morphologies of the fabricated MfSNPs/PVDF/PTFE flat casting knife was quickly cleaned, and the double casting was done in a
sheet membranes and (2) to evaluate the performance of the fabricated period of 30 s to avoid the coagulation of the pre-cast polymer solution
membranes in Congo Red (CR) dye removal using the DCMD config- from air-moisture. The cast solutions were exposed to air and immer-
uration. The membranes were fabricated via thermally induced solvent sion precipitation (IP) phase inversion method was used to produce thin
evaporation and the membrane surface was modified by incorporating film membranes. The membranes were dried between filter papers and
MfSNPS. annealed at 60 °C for 1 h. The composition of the nanocomposite
membranes is presented in Table 1. The membranes were prepared 3
2. Experimental times and the procedure was found to be reproducible.

2.1. Materials 2.5. Characterization of the nanoparticles and membranes

Poly(ethylene glycol) and polypropylene glycol (P123), tetraethyl The chemical composition and porosity of the nanoparticles are
orthosilicate (TEOS), hydrochloric acid (HCl), hexamethyldi-silazane important parameters that can be used to confirm the functionalization
(HMDS), powdered poly(vinylidene) fluoride (PVDF), powdered poly of the mesoporous silica nanoparticles. Diffuse Reflectance Infrared
(tetrafluoroethylene) (PTFE), N-methyl-2-pyrrolidone (NMP) and Fourier Transform Spectroscopy (DRIFTS) was performed using a
acetone were purchased from Sigma Aldrich and were used as-received. ThermoNicolett 6700 FT-IR spectrometer with a custom-made DRIFT
The materials were used in different experiments discussed in Sections cell at 120 °C under vacuum. Nitrogen sorption experiments were exe-
2.3-2.5. cuted on a Micromeretics Tristar at 77 K to obtain the internal surface
area via the Brunauer-Emmett-Teller (BET) theory (SBET) and pore size
2.2. Synthesis of silica nanoparticles (SNPs) distribution (dp,BJH) using the Barrett-Joyner-Halenda theory. The pore
volume (Vp) was determined at P/P0 = 0.95. AXIS SupraTM is an X-ray
Mesoporous silica nanoparticles (SNPs) were synthesized following photoelectron spectrometer (XPS) with unrivalled automation and ease
a modified procedure first reported by Zhao and co-workers, (1998) of use was used for materials surface characterization. The nano-
[24] (Fig. 1). An amount of 4 g of P123 (structure directing agent) was particles were prepared into a pellet form with the size of 1–1.5 mm
dissolved in 120 mL of 2 M HCl and 30 mL of distilled water (DI water). diameter. The prepared membranes were subjected to Fourier trans-
The mixture was vigorously stirred at ambient conditions until the form infrared-attenuated total reflection (FTIR-ATR) spectroscopic
surfactant was completely dissolved. TEOS (9.1 mL) was added into the measurements (Perkin Elmer FTIR spectrometer (Frontier Optica)) to
mixture and the temperature increased to 45 °C for 5 h under stirring. A obtain a good representation of the surface functional groups on the
white precipitate was formed. Subsequently, the temperature was membrane surface. Water contact angle analysis were measured to as-
raised to 90 °C for 18 h at static conditions. The mixture was allowed to sess the hydrophobic/hydrophilic properties of the membranes using a
cool to room temperature and the solids were filtered and washed 3 DSA3OE Kruss Drop shape analyzer, (GmbH, Germany).
times with deionized water and acetone. Finally the nanoparticles were
calcined at 550 °C for 6 h. The nanoparticles were characterized by 2.6. Membrane testing experimental design
nitrogen sorption measurements to determine their surface area, pore
size and pore volume. For the preparation of the feed solution, a stock solution (1000 mg/
L) was prepared by weighing 1 g of CR dye and dissolving in 1 L
2.3. Functionalization of the SNPs nanoparticles deionized (DI) water under vigorous stirring for 3–4 h. The experi-
mental feed solutions were prepared by diluting the dye stock solutions
Mesoporous silica (0.9 g) was dispersed in hexamethyldisilazane to 100 mg/L in a 2 L volumetric flask. Flux and rejection measurements
(HDMS) and the mixture was vigorously stirred for 5 h at 30 °C to were conducted using a DCMD lab-scale experimental setup with
functionalize the SNPs with hydrophobic methyl groups (Fig. 2). After deionized water. The feed temperature was kept at 50 °C while the

579
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

Fig. 1. Synthesis procedure of the silica (SiO2) nanoparticles.

Fig. 2. Functionalization of the SiO2 network.

Table 1
Casting solution composition of the fabricated membranes.
Membrane type PVDF (wt%) PTFE (wt%) MfSNPs (wt%)

M-1 15 0 0.3
M-2 15 3 0.3
M-3 15 6 0.3
M-4 15 0 0.5
M-5 15 3 0.5
M-6 15 6 0.5

Fig. 4. FTIR Spectra of the synthesized membranes. The compositions of M-1 to


M-6 are presented in Table 1.

Table 2
Specific surface area, pore volume and pore dimeter for SNPs and MfSNPs.
Property SNPS MfSNPs

2
Specific surface area (SBET) 667 m /g 361 m2/g
Pore volume (VP) 0.69 mL/g 0.49 mL/g
Pore diameter (DP) 9.1 nm 5.2 nm

(corresponding to a cross flow velocity of 0.4 m/s). The rejection per-


formance of the membranes was evaluated using CR dye, where the
Fig. 3. DRIFTS spectra for the synthesized nanoparticles. filtrate samples were collected after reaching a steady flux rate and the
feed solution was vigorously stirred to maintain homogenous mixture of
the feed solution. The rejection rate was estimated by the difference in
permeate was kept at 20 °C to create a temperature gradient. The pre-
the concentration of the feed and permeate. CR dye measurements were
pared membranes were inserted in the cross-flow mode module with a
performed using UV–vis spectroscopy.
membrane surface area of 0.0125 m2. The feed solution and permeate
The removal efficiencies of the membranes were calculated using
solutions were circulated at an operational speed of 35 L/h average
the Eq. (1):

580
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

Fig. 5. Nitrogen sorption-desorption isotherms (left) and pore size distribution (right) of the synthesized nanoparticles: SNPS (a and b) and MfSNPS (c and d).

Table 3 3. Results and discussion


Elemental quantification for SBA-15 NPs.
Quantification
3.1. DRIFTS and FTIR analysis

Atomic conc. [%] Error [%] Mass conc. [%] Error [%] Fig. 3 shows the DRIFTS spectra of the silica nanoparticles (SNPs)
and methyl functionalized silica nanoparticles (MfSNPs). The broad
SNPs O 1s 69.16 0.32 56.09 0.37
Si 2p 30.84 0.32 43.91 0.37
band at 3746 cm−1 and the intense peak at 1253 cm−1 on the SNPs
spectra are attributed to hydrogen bonding in molecular H2O. The
MfSNPs C 1s 20.25 0.51 14.28 0.37
broad band between 3500 and 3300 cm−1 originated from internal and
O 1s 58.34 0.42 52.06 0.34
Si 2p 21.41 0.26 33.67 0.32 water related H-bonding. The band at 740 cm−1 corresponds to Si-O2
stretching and deformation vibrations. The silanol peak disappeared
completely in the MfSNPs spectra due to the reaction with the trimethyl
Cf −Cpt ⎞ silyl function. Additionally, some new signals appeared at 1226 cm−1
%Removal = ⎛ ⎜ × 100

⎝ Cf ⎠ (1) originating from the newly introduced methyl groups. These observa-
tions clearly confirmed the successful introduction of methyl functions
where Cf and Cp are the feed and permeate concentrations, respectively. on the SNP materials.
The flux was calculated using Eq. (2): Different FTIR absorption bands of the polymers correspond to
different PVDF polymorphs. For example, the signals at 763 cm−1 and
Jwater = (Mt + 1−−Mt )/(tt + 1−−tt )Am (2) 840 cm−1 correspond to α and β forms of PVDF [28,29]. In Fig. 4, the
signal at 763 cm−1 corresponds to the in-plane bending or vibration of
where the membrane surface is denoted as Am, Mt+1 and Mt are the the α-phase, whereas the band at 840 cm−1corresponds to the
permeate masses at time tt+1 and tt, respectively.

581
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

Fig. 6. (a) XPS survey scan for the SNPs and MfSNPs capped nanoparticles, and (b), deconvoluted C 1s spectrum of the methyl functionalized silica nanoparticles.

nitrogen adsorption (lower curve) and desorption (upper curve) of the


isotherm measured. The lag-loop (hysterisis) observed between the
adsorption and desorption isotherms indicates the existance of meso-
porous pores on both the SNPs and MfSNPs. The pore distribution of the
SNPs and MfSNPs (Fig. 5(b) and (c), respectively) indicate that the
nanoparticles were smaller in size with a small pore size distributoin of
3–10 for MfSNPs and 7–12 SNPs indicating a decrease in the pore size.

3.3. X-ray photoelectron spectroscopy analysis

Surface elemental compositions of the nanoparticles are presented


in Table 3. The XPS results suggest that the atomic concentration of
carbon on the MfSNPs was 20.5% indicating that the signal was not due
to the adsorption of carbon dioxide (CO2) on the surface of the nano-
particles but attributed to the methyl groups. It was observed that the
atomic concentration of oxygen on the MfSNPs was lower than that of
the SNPs.
Fig. 6(a and b) shows the XPS spectra of the raw SNPs and the
Fig. 7. Measured contact angles for nanocomposite membranes. MfSNPs. The peaks at 103.5 eV (Si-O) and 150.5 eV (Si 2p) were at-
tributed to silica and the peak at 532.5 eV was attributed to oxygen (O
vibrations of the CH2 groups and the asymmetric stretching of the β 1s). The signal at 285.4 eV (C-OR clearly indicates that silica nano-
phase. The band at 1153 cm−1 corresponds to the CF2 group symme- particles were modified (Fig. 6b). This signal was attributed to carbon
trical stretching whereas 1482 cm−1 corresponds to γ phase. The most in the form of methyl group (–CH3) which confirms the functionaliza-
interesting band is the wide band at 1263 cm−1 which may be attrib- tion of the mesoporous silica nanoparticles by HMDS. The deconvoluted
uted Si-C stretching and deformation vibrations therefore suggesting spectra (Fig. 6(b)), shows the presence of the CeC, CeH and C 1s
the presence of MfSPNs on the surface of the membranes. The trans- bonding on the surface of the silica nanoparticles. This indicates the
mittance of the α-band at 763 cm−1has noticeably decreased indicating bonding states of the carbon to the silica core. Metal oxides carbonates
chemical changes on the polymorph of the PDVF brought about by may have hydrocarbon or carbonate on their core [27].
inclusion of the nanoparticles.
3.4. Contact angle measurements
3.2. Brunauer-Emmett-Teller (BET) analysis
The interaction of water with the surface of the membrane produces
The BET nitrogen sorption measurements determine the mean sur- surface tension on the membrane surface. Therefore contact angle
face area, pore volume and pore diameter of the nanoparticles. The measurement is an important membrane characterization technique
effect of functionalizing using HMDS on the pore size and surface area is because it the wettability of a membrane [30]. Wettability is an im-
presented in Table 2. After functionalization of the SNPs, a significant portant factor in membrane distillation as it gives an idea of the hy-
decrease in the surface area, pore volume and pore diameter was ob- drophobicity of the membranes. The results of the water contact angle
served. This was attributed to trimethyl-silyl functions decorating the measurements of the nanocomposite membranes are shown in Fig. 7.
pores. Therefore, confirming that the nanoparticles fabricated had a When the hydrophobic nanoparticles (nanofiller) were incorporated,
porous structure. the contact angles were increased. Increasing the concentration of the
Fig. 5(a) and (c) depict a typical 77 K N2sorption isotherm for the nanofiller further increased the contact angles of the nanocomposite
SNPs and MfSNPs, respectively. The sorption isotherms have both membranes. Small water contact angles (< 90°) correspond to high

582
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

Fig. 8. SEM images of the membranes (a) PVDF, (b) M-2, (c) M-4 and (d) M-6.

Fig. 9. AFM micrographs of (a) PVDF, (b) M-2, and (c) M-5.

wettability, while large water contact angles (> 90°) correspond to low 3.5. Scanning electron microscopy analysis
wettability [30]. As it can be observed in Fig. 7, the water contact
angles of the nanocomposite membranes were greater than 90 °C, sug- The surface morphology and chemical composition of the MD
gesting that the membranes were hydrophobic, and were thus expected membranes are the two factors determining the membrane wettability
to exhibit low wettability. during MD application. Scanning electron microscopy (SEM) and
atomic force microscopy (AFM) are two techniques that elaborate on

583
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

relative to PTFE concentration in the blended membranes i.e., the ad-


dition of 3 and 6 wt% PTFE did not affect the surface roughness, hence
images are not shown. It was the incoporation of nanoparticles that
increased the surface roughness of the nanocomposite membranes. M-2
(0.3 wt% MfSNPs) and M-5 (0.5 wt% MfSNPs) in Fig. 9 shows the dif-
ference. Although surface roughness is an undesirable property as it
increases membrane fouling, it is correlated to membrane hydro-
phobicity due to air entrapement in the micro voids of the rougher
membrane [6,9]. Indeed, the membranes with higher amounts of silica
nanoparticles, exhibited higher contact angles (Fig. 7). Hydrophobicity
is a key requirement in MD membranes to avoid membrane wetting
[6,9].

Fig. 10. Pure water flux of the membranes in a DCMD set-up. 3.7. Flux and rejection studies

The incorporation of nanoparticles such as Al2O3, TiO2, ZrO2, SiO2


the surface morphology of the membranes. To investigate the surface
and zinc oxide (ZnO) does not only improve the mechanical strength of
morphologies of the membranes; a typical JSM IT 300LV, JEOL scan-
PVDF membranes [33], but also membrane porosity, which positively
ning electron microscopy (SEM) was used to take the micrographs after
impacts on the membrane permeate fluxes. In this study, MfSNPs were
carbon coating (Fig. 8).
incorporated in the PVDF/PTFE blends and their effect on membrane
It was observed from the SEM images that incorporating MfSNPS
flux and rejection capacity was studied. Fig. 10 shows the effect of the
modified the neat structure of the pristine PVDF and impacted mor-
concentration of MfSNPs on membrane flux in a DCMD set-up. Com-
phology changes on the pore geometry of the membrane. From the top
pared with the fluxes obtained from ‘virgin’ PVDF/PTFE membranes, it
surface of the membranes, it can be observed that the nanocomposite
was observed that the MfSNPs enhanced the pure water (deionized
membranes possessed larger pores but fewer in number as compared to
water) flux of the nanocomposite membranes. The highest stable flux
the PVDF membranes. Membrane pores are formed by the demixing
obtained from the surface modified membrane was approximately
(solvent evaporation) during the phase inversion process [31]. Since the
0.0041 L/h m2 on M-3, which was comparable to other studies [34].
MSNs are hydrophobic due to methyl groups present on the surface of
The most important property of MD membranes is to maintain stable
the nanoparticles, their dispersal was improved in the polymeric solu-
flux as this corresponds to low wetting properties. Additionally, the
tions as shown by the occurrence of the enlarged pores.
water flux was generally lower in M-5 and M-6 due to a possible re-
duction in membrane pore sizes induced by the SNP filling of the micro-
3.6. Atomic force microscopy analysis voids of the membrane.
The CR dye removal efficiency of the MfSNPs/PVDF/PTFE mem-
In addition to SEM micrographs, atomic force microgaphs were branes (M-1, M-4) and (M-2, M-3, M-5 and M-6), exhibited a similar
obtained from atomic force microscopy (AFM), WITec Alpha 300 rejection, which decreased gradually with time (Fig. 11 (a)). This was
atomic force microscope (WITec, GmbH, Germany) to investigate the due to the increase in the CR dye concentration which may have re-
changes in the membrane surface due to the incorporation of fMSNs. sulted in concentration polarization. This makes the membranes prone
Surface roughness parameters were calculated by Control Four soft- to fouling causing the pore blockage and contamination of the permeate
ware, (WITec, GmbH, Germany). The average square roughness (SA) solution. The MfSNPs/PVDF/3% PTFE membranes (M-5) showed a
defines the arimethicaverage of the measured peak heights from a one drastic decrease in the membrane removal efficiency after 3 h possibly
dimensional plane whereas the root mean square roughness (Sq) refers due to membrane wettability induced by fouling. The driving force for
to the standard deviation. The topography micrographs and surface separation in MD is vapour difference across the two interfaces of the
rougness parameters are presented in Fig. 9. hydrophobic membrane [16]. Membrane wetting promotes the passage
It was observed that the surface roughness of the PVDF/PTFE of the solutions in the liquid state. The high porous nature of these
blended membranes was larger than the pristine PVDF membranes. The membranes, allowed the passage of the dye molecules, which conse-
trend of these results corresponds with the surface roughness results of quently reduced the removal efficiency [16,21]. The removal efficiency
commercial PVDF/PTFE flat sheet membranes (Mingle, China) [32]. obtained for CR dye was comparable to the removal efficiency obtained
The pristine PVDF membrane had the smoothest surface with for methylene blue separation using DCMD [5]. Pore wetting in MD is
SA = 73.947 nm. There was no significant variation of roughness another challenge that can result in declining efficiency and

Fig. 11. Removal of CR dye, (a) removal efficiency and (b) flux during the DCMD experiment.

584
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

deterioration of the permeate production rate and quality when using [4] K.C. Chong, S.O. Lai, H.S. Thiam, S.S. Lee, W.J. Lau, N.M. Mokhtar, Reactive blue
hydrophobic membranes. Pore wetting is a result of the feed solution dye removal by membrane distillation using PVDF membrane, Ind. J. Sci. Technol.
9 (2016) 1–5, https://doi.org/10.17485/ijst/2016/v9iS1/110379.
passing through the pores of the hydrophobic membrane due to high [5] A.K. An, J. Guo, S. Jeong, E.J. Lee, S.A.A. Tabatabai, T.O. Leiknes, High flux and
pressure. This cannot be avoided in most cases. antifouling properties of negatively charged membrane for dyeing wastewater
Fig. 11(b) illustrates the permeate flux (J/Jo) of the nanocomposite treatment by membrane distillation, Water Res. 103 (2016) 362–371, https://doi.
org/10.1016/j.watres.2016.07.060.
membranes using 100 ppm CR dye. Generally, the permeate fluxes of [6] A. Alkhudhiri, N. Darwish, N. Hilal, Membrane distillation: a comprehensive re-
the all the membranes decreased with time. This was attributed to the view, Desalination 287 (2012) 2–18, https://doi.org/10.1016/j.desal.2011.08.027.
increase of CR dye concentration in the feed side due to the evaporation [7] P. Wang, T. Chung, Recent advances in membrane distillation processes: mem-
brane, development configuration design and application exploring, J. Membr. Sci.
of water molecules on the membrane surface, through the membrane (2015), https://doi.org/10.1016/j.memsci.2014.09.016.
pores to the permeate side. This resulted in a difference in bulk and [8] L.M. Camacho, L. Dumée, J. Zhang, J. de Li, M. Duke, J. Gomez, S. Gray, Advances
membrane surface concentration, thereby contributing to the reduction in membrane distillation for water desalination and purification applications, Water
(Switzerland) 5 (2013) 94–196, https://doi.org/10.3390/w5010094.
in the partial vapour pressure and therefore, a decline in the fluxes can
[9] M.M. Teoh, T.S. Chung, Y.S. Yeo, Dual-layer PVDF/PTFE composite hollow fibers
be experienced with time [35]. This is known as the concentration with a thin macrovoid-free selective layer for water production via membrane
polarization phenomena that is under investigation in a separate on- distillation, Chem. Eng. J. 171 (2011) 684–691, https://doi.org/10.1016/j.cej.
going study. Another factor that can result in flux decline is the 2011.05.020.
[11] M.M.A. Shirazi, A. Kargari, M. Tabatabaei, Evaluation of commercial PTFE
blockage of membranes pores due to fouling caused by hydrophobic membranes in desalination by direct contact membrane distillation, Chem. Eng.
interactions between the surface of the membrane and CR dye mole- Process. Process Intensif. 76 (2014) 16–25, https://doi.org/10.1016/j.cep.2013.
cules. M-2 had a higher flux as compared to other membranes signaling 11.010.
[12] M. Khayet, Membranes and theoretical modeling of membrane distillation: a re-
better performance. view, Adv. Colloid Interface Sci. 164 (2011) 56–88, https://doi.org/10.1016/j.cis.
2010.09.005.
4. Conclusion [13] B.-H. Jeong, E.M.V. Hoek, Y. Yan, A. Subramani, X. Huang, G. Hurwitz, A.K. Ghosh,
A. Jawor, Interfacial polymerization of thin film nanocomposites: a new concept for
reverse osmosis membranes, J. Membr. Sci. 294 (2007) 1–7, https://doi.org/10.
The MfSNPs were successfully functionalized with methyl groups as 1016/j.memsci.2007.02.025.
confirmed by IR analysis and nitrogen adsorption-desorption mea- [14] Z.D. Hendren, J. Brant, M.R. Wiesner, Surface modification of nanostructured
ceramic membranes for direct contact membrane distillation, J. Membr. Sci. 331
surements. The MfSNPs obtained were hydrophobic as seen from the (2009) 1–10, https://doi.org/10.1016/j.memsci.2008.11.038.
water contact angle measurements of the membranes, which increased [15] S.R. Krajewski, W. Kujawski, M. Bukowska, C. Picard, A. Larbot, Application of
with increasing filler loading. A 0.5 wt% loading significantly improved fluoroalkylsilanes (FAS) grafted ceramic membranes in membrane distillation
process of NaCl solutions, J. Membr. Sci. 281 (2006) 253–259, https://doi.org/10.
the PVDF contact angle from 97° to 120°. The results confirmed that the
1016/j.memsci.2006.03.039.
incorporation of the methyl functionalized MfSNPs did not only im- [16] S. Cerneaux, I. Struzyńska, W.M. Kujawski, M. Persin, A. Larbot, Comparison of
prove the fluxes but also induced low wettability properties of the various membrane distillation methods for desalination using hydrophobic ceramic
membranes. Clean water flux was observed to be stable over a long membranes, J. Membr. Sci. 337 (2009) 55–60, https://doi.org/10.1016/j.memsci.
2009.03.025.
period of time in all the MfSNPs incorporated and non-incorporated [17] Y. Zhuang, F. Yu, J. Chen, J. Ma, Batch and column adsorption of methylene blue by
membranes. However, the feed solution containing the CR dye resulted graphene/alginate nanocomposite: comparison of single-network and double-net-
to the decline in fluxes as a function of time. The flux decline induced work hydrogels, J. Environ. Chem. Eng. 4 (2016) 147–156, https://doi.org/10.
1016/j.jece.2015.11.014.
by the CR dye solution was associated with hydrophobic-hydrophobic [18] C. Cabassud, D. Wirth, membrane distillation for water desalination : how to chose
membrane and dye interactions leading to membrane fouling. Although an appropriate membrane? Desalination 157 (2003) 307–314.
the acid dyes are predominantly hydrophobic, they possess the hydro- [19] L. Eykens, K. De Sitter, C. Dotremont, L. Pinoy, B. Van Der Bruggen, How to opti-
mize the membrane properties for membrane distillation: a review, Ind. Eng. Chem.
philic anionicity which can cause membrane hydrophilicity. The in- Res. 55 (2016) 9333–9343, https://doi.org/10.1021/acs.iecr.6b02226.
duction of the membrane hydrophilicity reduces the vapour pressure [20] M. Khayet, A. Velázquez, J.I. Mengual, Modelling mass transport through a porous
gradient and allows the passage of the liquid across the membrane, partition: effect of pore size distribution, J. Non-Equilibrium Thermodyn. 29 (2004)
279–299, https://doi.org/10.1515/JNETDY.2004.055.
hence reduce rejection efficiency as a function of time. Therefore, fur-
[21] T. Jiricek, M. Komarek, J. Chaloupek, T. Lederer, Flux enhancement in membrane
ther studies are required to maintain the balance between the flux and distillation using nanofiber membranes, J. Nanomater. 2016 (2016), https://doi.
rejection efficiency in water systems characterized by high levels of org/10.1155/2016/9327431.
[22] J.E. Efome, M. Baghbanzadeh, D. Rana, T. Matsuura, C.Q. Lan, Effects of super-
hydrophobic solutes.
hydrophobic SiO nanoparticles on the performance of PVDF flat sheet membranes
for vacuum membrane distillation, Desalination 373 (2015) 47–57, https://doi.org/
Acknowledgements 10.1016/j.desal.2015.07.002.
[23] E. Maria Claesson, A.P. Philipse, Thiol-functionalized silica colloids, grains, and
membranes for irreversible adsorption of metal(oxide) nanoparticles, Colloid Surf.
The authors would like to acknowledge the financial support from A Physicochem. Eng. Asp. 297 (2007) 46–54, https://doi.org/10.1016/j.colsurfa.
the University of South Africa (UNISA) and Ghent University. We are 2006.10.019.
thankful to Particle and Interfacial Technology (PAINT) Group and the [24] D. Zhao, J. Feng, Q. Huo, N. Melosh, G. Fredrickson, B. Chmelka, G. Stucky,
Triblock copolymer syntheses of mesoporous silica with periodic 50 to 300 ang-
Nanotechnology and Water Sustainability (NanoWS) Research Unit for strom pores, Science 279 (1998) 548–552.
providing research facilities. [25] Z. Dong, X. Ma, Z. Xu, W. You, F. Li, Superhydrophobic PVDF – PTFE electrospun
nano fibrous membranes for desalination by vacuum membrane distillation,
Desalination 347 (2014) 175–183, https://doi.org/10.1016/j.desal.2014.05.015.
Appendix A. Supplementary material [26] O.T. Mahlangu, R. Nackaerts, J.M. Thwala, B.B. Mamba, A.R.D. Verliefde,
Hydrophilic fouling-resistant GO-ZnO/PES membranes for wastewater reclamation,
Supplementary data to this article can be found online at https:// J. Membr. Sci. 524 (2017) 43–55, https://doi.org/10.1016/j.memsci.2016.11.018.
[27] S. Pourbeyram, Effective removal of heavy metals from aqueous solutions by gra-
doi.org/10.1016/j.seppur.2018.10.039.
phene oxide−zirconium phosphate (GO−Zr-P) nanocomposite, Ind. Eng. Chem.
Res. (2016), https://doi.org/10.1021/acs.iecr.6b00728.
References [28] J. Liu, X. Lu, C. Wu, Effect of preparation m]methods on crystallization behavior
and tensile strength of poly(vinylidene fluoride) membranes, Membranes (Basel) 3
(2013) 389–405, https://doi.org/10.3390/membranes3040389.
[1] C.A. Quist-Jensen, A. Ali, S. Mondal, F. Macedonio, E. Drioli, A study of membrane [29] R. Gregorio, Determination of the α, β, and γ crystalline phases of poly(vinylidene
distillation and crystallization for lithium recovery from high concentrated aqueous fluoride) films prepared at different conditions, J. Appl. Polym. Sci. 100 (2006)
solutions, J. Membr. Sci. 505 (2016) 167–173, https://doi.org/10.1016/j.memsci. 3272–3279, https://doi.org/10.1002/app.23137.
2016.01.033. [30] Y. Yuan, T.R. Lee, Contact Angle and Wetting Properties, in: G. Bracco, B. Holst,
[3] K.M. Udert, M. Wächter, Complete nutrient recovery from source-separated urine (eds), (Eds.), Surface Science Techniques. Springer Series in Surface Sciences,
by nitrification and distillation, Water Res. 46 (2012) 453–464, https://doi.org/10. Springer, Berlin, Heidelberg, 2013, , https://doi.org/10.1007/978-3-642-34243-
1016/j.watres.2011.11.020. 1_1.

585
N.P. Khumalo et al. Separation and Purification Technology 211 (2019) 578–586

[31] B.S. Lalia, V. Kochkodan, R. Hashaikeh, N. Hilal, A review on membrane fabrica- 10.3390/membranes4010055.
tion: structure, properties and performance relationship, Desalination 326 (2013) [34] K.T. Rashid, S. Binti, A. Rahman, Enhancement the flux of PVDF-Co-HFP hollow
77–95, https://doi.org/10.1016/j.desal.2013.06.016. fiber membranes for direct contact membrane distillation applications, Ind. J. Sci.
[32] K. Xiao, J. Sun, Y. Mo, Z. Fang, P. Liang, X. Huang, J. Ma, B. Ma, Effect of membrane Technol. 11 (2016) 2189–2192, https://doi.org/10.17485/ijst/2017/v10i7/
pore morphology on microfiltration organic fouling: PTFE/PVDF blend membranes 111446.
compared with pvdf membranes, Desalination 343 (2014) 217–225, https://doi. [35] T. Jiracek, M. Komarek, J. Chaloupek, T. Lederer, Flux enhancement in membrane
org/10.1016/j.desal.2013.09.026. distillation using nanofiber membranes, J. Nanomater. 2016 (2016), https://doi.
[33] C.Y. Lai, A. Groth, S. Gray, M. Duke, Nanocomposites for improved physical dur- org/10.1155/2016/9327431.
ability of porous PVDF membranes, Membranes 4 (2014) 55–78, https://doi.org/

586

You might also like