You are on page 1of 11

Journal of Environmental Chemical Engineering 10 (2022) 107454

Contents lists available at ScienceDirect

Journal of Environmental Chemical Engineering


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

High-performance thin-film composite forward osmosis membranes with


hydrophilic PDA@TiO2 nanocomposite substrate for the treatment of oily
wastewater under PRO mode
Parashuram Kallem a, b, *, Ravi P. Pandey a, b, Hanaa M. Hegab a, b, Ruchi Gaur c,
Shadi W. Hasan a, b, Fawzi Banat a, b, *
a
Center for Membranes and Advanced Water Technology (CMAT), Khalifa University of Science and Technology, PO Box 127788, Abu Dhabi, United Arab Emirates
b
Department of Chemical Engineering, Khalifa University of Science and Technology, PO Box 127788, Abu Dhabi, United Arab Emirates
c
Solid State and Structural Chemistry Unit, Indian Institute of Science, Bangalore 560012 Karnataka, India

A R T I C L E I N F O A B S T R A C T

Editor: Dr. Zhang Xiwang The applications of thin-film composite (TFC) forward osmosis (FO) membranes are limited by low flux and
strong internal concentration polarization (ICP). In this study, a simple, facile, and eco-friendly approach to
Keywords: improve the wettability of the FO membrane substrate using hydroxyl- and amino-functionalized TiO2 nano­
Functionalized TiO2 composites by polydopamine (PDA) coating on TiO2 particles is demonstrated. In our research, a composite
Eco-friendly approach
membrane substrate was fabricated by incorporating PDA@TiO2 into polyethersulfone (PES) through the non-
Forward osmosis
solvent induced phase separation (NIPS) process. Then, a selective active polyamide layer was formed by
High flux
Anti-fouling interfacial polymerization on top of the PES-PDA@TiO2 membrane substrate to obtain the resulting TFC-FO
Oily wastewater membrane. The composite membrane substrate exhibited higher hydrophilicity (about 25% lower contact
angle) and higher porosity than the pristine membrane substrate. The water flux (Jw) of 34.3 L/m2h and 60.1 L/
m2h was achieved for the resulting TFC-PDA@TiO2 FO membrane under FO and pressure retarded osmosis (PRO)
modes which was ~174% and ~183% higher than that of the pristine TFC-FO membrane under FO and PRO
modes respectively. Furthermore, the TFC-PDA@TiO2 membrane exhibited enhanced selectivity, as expressed by
the decrease in the specific salt flux (Js/JW) values (from 0.48 g/L and 0.53 g/L of pristine TFC to 0.27 g/L and
0.28 g/L for TFC-PDA@TiO2 under the FO and PRO modes respectively). More importantly, the TFC-PDA@TiO2
membrane exhibited better fouling behavior with the organic foulant solution. These results demonstrate that the
fabricated TFC-FO membranes have significant potential for sustainable water reclamation from oily wastewater
via FO under the PRO mode.

1. Introduction and clean water has led to the development of sustainable processes for
wastewater treatment. However, remedying oily wastewater pollution
Rapid industrial developments in the oil and gas sector have gener­ using conventional approaches, such as absorbent materials, flotation,
ated large volumes of oily wastewater worldwide. Oily wastewater gravity separation, centrifugation, and coagulation, is inefficient and
pollution adversely affects groundwater and drinking water resources generates secondary pollution [3]. Therefore, it is essential to develop
threatening aquatic life and human health, and thus is a problem to be proficient, effective, and low-cost techniques for oil/water separation
addressed and remedied through a recycling process [1]. Generally, oily [4].
wastewater exists in free droplets, dispersed and emulsified organ­ Forward osmosis (FO) is an osmotically-driven membrane separation
ic/inorganic materials, and immersed particulates [2]. The discharge of process that has received intensive attention for treating emulsified oily
such oily wastewater into uncontaminated water sources without prior wastewater due to its low energy input, ease of function, low operating
treatment has become a considerable global concern. Demand for fresh costs, high rejection rate, and high water recovery [5–7]. Several studies

* Corresponding authors at: Center for Membranes and Advanced Water Technology (CMAT), Khalifa University of Science and Technology, PO Box 127788, Abu
Dhabi, United Arab Emirates.
E-mail addresses: parashuram.kallem@ku.ac.ae (P. Kallem), fawzi.banat@ku.ac.ae (F. Banat).

https://doi.org/10.1016/j.jece.2022.107454
Received 28 December 2021; Received in revised form 10 February 2022; Accepted 21 February 2022
Available online 22 February 2022
2213-3437/© 2022 Elsevier Ltd. All rights reserved.
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

on FO have been conducted on polyamide (PA)-based thin film com­ constant room temperature. The color of the dispersed solution changed
posite (TFC) membranes due to their higher water permeability, selec­ to dark brown indicating the formation of PDA@TiO2 composites.
tivity, and good mechanical strength [8,9]. Recently, alternative PDA@TiO2 composites were washed thrice with DI water and kept in a
membrane fabrication approaches have been utilized to further improve drying oven at 80 ◦ C for 12 h. A schematic representation of the syn­
the separation performance of TFC membranes [10–12]. A number of thesis protocol is given in Fig. 1a.
commonly used membrane substrates made of polymers, such as poly­
vinylidene fluoride (PVDF), polyacrylonitrile (PAN), polybenzimidazole 2.3. Fabrication of PES-PDA@TiO2 nanocomposite substrates
(PBI) polysulfone (PSf), and polyethersulfone (PES), have been deployed
[13–17]. PES possesses excellent chemical strength, thermal stability, The nonsolvent-induced phase separation (NIPS) method was used to
and superior mechanical properties. However, PES is not sufficiently fabricate a pristine and composite PES FO membrane. The casting so­
hydrophilic thereby limiting its FO treatments [18]. Further, the lution was prepared by sonication of PDA@TiO2 nanocomposite parti­
incorporation of nanoparticles into typical polymeric membrane mate­ cles in NMP for 1 h to minimize the agglomeration of particles on the
rials, known as nanocomposite membranes, has been presented as a way membrane surface. The PVP and PES were added simultaneously to the
to develop high-performance membranes [19]. Incorporating hydro­ casting solution for 24 h under a vigorous stirring protocol. The homo­
philic nanomaterials into the substrate layer of FO membranes helps geneous casting solution was left for 2 h to remove air bubbles from the
improve hydrophilicity, separation performance, and fouling-resistant solution. The prepared solution was then uniformly cast on a non-woven
properties. Various types of inorganic nanomaterials, such as carbon supported glass plate using a casting knife. The plate was then immersed
nanotubes (CNTs), TiO2, SiO2, graphene oxide (GO), and zeolite, have in an aqueous coagulation bath at room temperature for phase separa­
been introduced to functionalize TFC membranes [20–23]. Among these tion (Fig. 1b). After the membrane solidified on a glass plate, it was
various nanomaterials TiO2 nanoparticles have been seen to have moved to an additional coagulation bath to wash and remove the
particular advantages due to their low cost, high chemical stability, possible residual NMP. Finally, the membrane substrates were immersed
non-toxicity, and photocatalytic activity [24,25]. Studies demonstrate and stored in DI water until required. Three substrates for the FO
the feasibility of improving the FO performance through the incorpo­ membrane, namely, PES, PES-TiO2, and PES-PDA@TiO2, were prepared
ration of TiO2 particles into the substrate of the FO membrane [26,27]. according to the compositions shown in Table S1.
Despite this, the benefits of improving membrane hydrophilicity and
fouling resistance have been limited [8,28]. There is a considerable 2.4. Preparation of the TFC-FO membrane
desire to functionalize TiO2 with hydrophilic groups to improve hy­
drophilicity and dispersibility so that TiO2 can prove its significant For the preparation of thin TFC membranes, the polyamide (PA)
positive effects [29]. selective skin layer was deposited on the surface of the substrates via
Dopamine (DA) is a mussel-inspired biomolecule consisting of interfacial polymerization (IP) [7]. As illustrated in Fig. 1c, the first 2%
abundant functional groups such as imine, amine, and catechols. It can (w/v) MPD in a 100 mL DI solution was introduced on the surface of the
undergo self-polymerization on material surfaces in mild alkaline membrane substrates and allowed for 10 min to ensure complete
media, producing a thin film of polydopamine (PDA). PDA has been dispersion of the MPD solution within the pore. The residual MPD so­
widely used to improve the surface charge, permeability, and anti- lution was drained from the membrane substrate surface using a rubber
fouling properties in various membranes [4,30]. Combining the prop­ roller. The acyl halide solution (0.1% (w/v) TMC in 100-mL hexane) was
erties of TiO2 and PDA, such as hydrophilicity and negative surface cast on the substrate surface with a contact time of 2 min to react with
charge, would in a result in a hybrid material that would enhance the the MPD. To remove the excessive acyl halide solution, the substrate
permeability and fouling resistance of the FO membrane substrates. This surface was rinsed and washed with pure n-hexane. Subsequently, the
study then, describes the use of environmentally friendly PDA-coated substrate surface was dried in an oven at 60 ◦ C for 8 min. The resulting
TiO2 composites (PDA@TiO2) as nanocomposites for the fabrication of TFC-FO membrane was kept in DI water for further characterization.
highly permeable and fouling-resistant FO membrane substrates for the The resulting TFC membranes obtained from the substrates PES, PES-­
treatment of oily wastewater. TiO2 and PES-PDA@TiO2 were named TFC, TFC-TiO2 and TFC-PDA@­
TiO2 respectively.
2. Materials and methods
2.5. Membrane characterization
2.1. Materials
Bruker ALPHA-Platinum ATR FTIR was used to detect the functional
Polyethersulfone (PES pellets, Mw: 58 kDa) was purchased from groups present in the pristine TiO2 structure and functionalized
Good Fellow Co., UK. Polyvinylpyrrolidone (PVP, Mw: 40 kDa), N- PDA@TiO2 nanocomposite in the range of 400–4000 cm− 1. The crys­
methyl-2-pyrrolidone (NMP; 99%), dopamine hydrochloride (DH; 99%), tallinity and size of the samples were measured at 2θ range of 10◦ − 80◦
Trizma®hydrochloride (THCl;>99.0%), and titanium oxide (TiO2; via X-ray diffraction (XRD) PANalytical Empyrean at 0.01◦ step size.
99.8%) were supplied from Sigma-Aldrich. Soy vegetable oil (commer­ Morphology studies were established by FEG QUANTA 250 scanning
cial grade) was purchased from a local supermarket. Sodium alginate, electron microscopy (SEM). Thermal analysis was carried out in a
calcium sulfate (CaSO4; 98%), 1,3-phenylenediamine (MPD; >99.0%) temperature range of 30–600 ◦ C at a rate of 10 ◦ C/minute by using TGA
and acyl chloride monomer i.e. 1,3,5-benzene-tricarbonyl trichloride 4000, Perkin Elmer. FEI Tecnai 20 was used to study the transmission
(TMC; >98.0%) were obtained from Sigma-Aldrich. Deionized water electron microscopy (TEM) analysis at 200 kV. The water contact angle
(DI) (resistivity = 15 Mcm at 25 ◦ C), acquired from a Milli-Q Plus 185 (θ) of each membrane substrate was measured using the Krüss GmbH’
purification system (Millipore Corp., USA), was consistently used Drop Shape Analyzer (DSA) at room temperature 25 ◦ C. The zeta po­
throughout the experimental process, tential (ζ) values of the membrane substrates were obtained using with
Anton Paar SurPASS Electrokinetic Analyzer. The porosity (ε) of the
2.2. Synthesis of PDA-coated TiO2 membrane substrate was calculated by the gravimetric method. First,
the weight of the dried porous substrate was estimated and then the
TiO2 nanocomposites coated with PDA (named PDA@TiO2) were membrane was soaked in DI water. After removing the substrate from
prepared as described in our previous publication [4]. The 5:1 ratio of the DI water, the surface was wiped with tissue paper. Finally, the
fresh TiO2 and dopamine was mixed in the 20 mL Tris-HCl solution weight of the wet membrane surface was measured and which had
adjusted to pH 8.5. The dispersed solution was stirred for 12 h at increased due to the penetration of water between the pores. The bulk

2
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

Fig. 1. Schematic illustration of (a) synthesis of PDA@TiO2 (b) fabrication of PES-PDA@TiO2 substrates and (c) preparation of TFC-FO membrane.

porosity was calculated using Eq. 1 [31]: applied pressure (bar) respectively.
The solute rejection (R) and salt permeability coefficient B (L/m2h)
ε (%) = (w₁ − w₂)
(1) of the membranes were acquired in the presence of 100 mM NaCl feed
Aρ∂ × 100
solution by Eqs. (3) and (4):
( )
where w1 and w2 represent the weights of the wet and dry membrane Cp
substrates respectively in g. A, ρ, and ∂ denote the area of the substrate R(%) = 1 − x100 (3)
Cf
surface (cm2), water density (g/cm3), and thickness of the membrane
( )
(cm) respectively. 1
B = A(ΔP − Δπ ) − 1 (4)
R

2.6. Membrane performance evaluation where Cp, Cf, and Δπ denote the concentrations of permeate (g/L), feed
solution (g/L), and the change in osmotic pressure across the membrane
Intrinsic separation properties, i.e., water permeability coefficient (bar) respectively.
(A, L/m2h bar), salt permeability coefficient (B, L/m2h), and salt The osmotic performance of the TFC membranes was performed
rejection (R) were measured using a lab-scale cross-flow reverse osmosis using a lab-scale FO system with an effective membrane area of 36 cm2
(RO) unit, as described in our previous work [32]. The pure water at room temperature using 1 M NaCl as the draw solution (DS) and DI
permeability coefficient (A) was evaluated at 3 bars and calculated using water as the feed solution (FS) [32]. FO measurements were evaluated in
Eq. 2: FO (AL faces the FS) and PRO (AL faces the DS) modes at 1 M NaCl as DS
ΔV and DI water as FS. The impact of different concentrations of draw so­
A = (2) lution (0.5–2.0 M of aqueous NaCl) on the membrane surface was also
Am Δt ΔP
investigated. The water flux, Jw (L/m2h), was calculated by Eq. (5):
where ΔV, Am, Δt, and ΔP are the volume of obtained permeate (L),
effective area of the membrane (m2), a specific duration of time (h), and

3
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

Fig. 2. (a) FT IR spectra and (b) Raman spectra of TiO2 and PDA@TiO2. (c-d) SEM microimages of TiO2 and PDA@TiO2. TEM micrographs of (e) TiO2 and
(f) PDA@TiO2.

4
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

ΔV unmodified TiO2, the FTIR spectrum of the PDA@TiO2 nanocomposite


Jw = (5)
Δt × Am showed additional high-intensity peaks at 1290 cm− 1, 1670 cm− 1, and
where ΔV, Δt, and Am are volume changes (L), specific duration of 1482 cm− 1 which corresponds to the aromatic ring of dopamine [35].
time (h), and the effective membrane area (m2). The reverse salt flux, Js The characteristic peak at 3300 and 3210 cm− 1 is attributed to -NH and
(g/m2h), was determined using Eq. (6): -OH vibrations suggesting a successful PDA coating on the TiO2 surface.
Furthermore, strong peaks at 1183 cm− 1and 1070 cm− 1 were distinctly
Js =
VtCt − VoCo
(6) noticeable representing C–O–C stretching and C–O–H stretching vibra­
Δt × Am tions. This confirms the formation of the PDA@TiO2 nanocomposite
where Ct and Co are the initial and final concentrations (g/L), while through chemical reactions between the PDA and TiO2 surfaces. In
Vt and Vo are the initial and final volumes (L) of the feed solution. addition, the Raman spectra of TiO2 and the PDA@TiO2 nanocomposite
The fouling behavior of the TFC-FO membrane was determined by (depicted in Fig. 2b) displayed four common characteristic peaks be­
SA polysaccharide as a model foulant agent in AL-DS mode. The con­ tween 133 and 600 cm− 1 in both pristine TiO2 and PDA@TiO2 nano­
centration of the SA solution was 200 mg/L with 1 mM calcium cation. composite confirming the distinctive TiO2 phase. A broad peak at
SA emulsion and NaCl were used in the fouling studies as feed and draw 1580 cm− 1 in the PDA@TiO2 nanocomposite indicated the coating of
solutions [33]. After stabilizing the permeate water flux, the fouling the PDA on the TiO2 [24].
investigation was conducted continuously for 6 h and the normalized The surface morphological structures of the pristine TiO2 and
water flux was measured based on the decline in the water flux. PDA@TiO2 nanocomposite are shown in Fig. 2c-d at a magnification of
Furthermore, the flux recovery capability of the membrane was evalu­ 5 µm. The SEM micrographs of TiO2 and PDA@TiO2 nanocomposites
ated using 1 M NaCl salt as DS and DI water as FS after a simple physical indicate uniform and compact spherical particles. The structures of the
washing. Oil/water separation experiments were performed to assess pristine TiO2 and PDA@TiO2 nanocomposite were further evaluated by
the applicability of pristine TFC and TFC-PDA@TiO2 FO membranes HR-TEM micrographs (Fig. 2e-f). It can be seen from Fig. 2e that the TiO2
under the PRO mode. In this part of the experiment 500 mg/L of particles were unevenly distributed and agglomerated strongly in the
oil-water (O/W) or 0.1 g of sodium dodecyl sulfate (SDS) matrix while the PDA@TiO2 nanocomposites (Fig. 2f) showed good
surfactant-added oil-water (S/O/W) mixed solutions were used as FS dispersion. This could be due to the PDA coating on the TiO2 surface
and 1.0 M NaCl as DS [34]. Oil/water separation experiments were acting as a surfactant and providing reactive sites for the secondary
performed for 7 h and the normalized water flux was measured based on organic reaction [35,36].
the flux decline.

3. Results and discussion 3.2. Characterization of FO membrane substrates and TFC membranes

3.1. Characterization of TiO2 and PDA@TiO2 The surface morphologies of the membrane substrates are shown in
Fig. 3. In the bottom-surface images, the membrane substrate showed a
The chemical structure of the TiO2 and PDA@TiO2 nanocomposite good distribution of pores, while top-surface images revealed a smooth
was examined using FT-IR spectra, as shown in Fig. 2a. Compared to surface without visible pores. Compared to the pristine PES membrane
substrate some particle aggregates were embedded on the surface of the

Fig. 3. SEM micrographs of fabricated pristine PES and nanocomposite PES membrane substrates.

5
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

PES-TiO2 and PES-PDA@TiO2 membrane substrates. Furthermore, the that the membrane substrate with higher surface roughness has a higher
EDS spectrum (Fig. S1) supports the presence of C, O, S, and Ti elements permeability owing to the improved surface area. The surface SEM
in the PES-TiO2 membrane substrate, whereas the N element was also morphology of the prepared TFC membranes was also examined and
observed in the PES-PDA@TiO2 membrane substrate. presented here in Fig. 5. Defect-free films with typical ridge-and-valley
The porosity was influenced by the incorporation of nanoparticles surface structures were observed for all TFC FO membranes.
into the membrane substrate (Fig. S2). The bulk porosity of the pristine
PES, PES-TiO2, and PES-PDA@TiO2 membrane substrate was found to be 3.3. Membrane separation properties
~72, ~75, and ~78%, respectively. The increased porosity in the
composite membrane substrates is due to the immediate exchange of the The important properties of the FO membrane separation properties,
solvent and non-solvent formed by nanoparticles during the NIPS including water permeability (A) and salt permeability (B), are tabu­
process. lated in Table 1. The TFC-PDA@TiO2 membrane showed higher water
Wettability is a key factor affecting the fouling resistance of the permeability with good salt rejection compared to the pristine TFC and
membranes. The contact angle (dynamic and static) of the FO membrane TFC-TiO2 membranes. The B/A ratio is a direct indicator of membrane
supports is shown in Fig. 4a. Throughout the 80 s, the contact angle for selectivity. A minimal B/A ratio is usually preferred to lower reverse
the pristine PES membrane substrate decreased with the addition of diffusion of the solute [17]. In this study, the B/A value of the
TiO2 and PDA@TiO2 particles indicating that the hydrophilicity of the TFC-PDA@TiO2 membrane was less than that of the pristine TFC
membrane substrate is improved. Among all membrane substrates PES membrane.
with PDA@TiO2 had the best hydrophilicity due to the well-dispersed
hydroxyl- and amino-functionalized TiO2 particles in the membrane 3.4. FO performance of TFC membranes
substrate and which is anticipated to improve the permeability anti-
fouling performance of the FO membrane. Furthermore, all membrane The FO performance of fabricated TFC membranes is assessed in
substrates exhibited excellent thermal stability, as shown in Fig. 4b, with terms of water flux (Jw), reverse solute flux (Js), and specific solute flux
a slight enhancement of the thermal stability due to the incorporation of (Js/Jw). Fig. 6a shows the Jw in both the FO (AL-FS) and PRO (AL-DS)
nanoparticles. For instance, the decomposition temperature Td (defined modes at 1 M NaCl as DS and DI water as FS. The results showed that the
at 15% weight loss) improved from 519 ◦ C in the pristine PES membrane composite TFC-TiO2 and TFC-PDA@TiO2 FO membranes exhibited
substrate to ~548 and ~534 ◦ C in PES-TiO2 and PES-PDA@TiO2 higher Jw compared to the control TFC membrane. In the FO mode, the
membrane substrates respectively. The zeta potential values obtained TFC-PDA@TiO2 FO membrane displayed the maximum Jw of ~34 L/
for the pristine PES, PES-TiO2, and PES-PDA@TiO2 substrates (Fig. 4c) m2h corresponding to a ~174% enhancement compared to the TFC-FO
were − 13, − 24, and − 41 mV at a pH of 7. membrane. A similar trend was noticed in the PRO mode, but with a
Adding PDA@TiO2 resulted in the surface of the membrane substrate high water flux and a reverse solute flux generated. To illustrate, the
becoming even more negative because of the negative − OH in PDA TFC-PDA@TiO2 membrane tested in the PRO mode showed around
moiety. An AFM analysis was carried out to study the effect of PDA@­ ~60 L/m2h water flux and 9 g/m2h reverse solute flux corresponding to
TiO2 filler on the PES surface. Changes in the surface roughness pa­ a ~75 and ~79% improvement over Jw and Js achieved in the FO mode
rameters, including average roughness (Ra) and root mean square respectively. The membrane substrates were also investigated for their
roughness (Rq), are presented in Fig. 4d. It can be seen that the surface pure water flux using an ultrafiltration unit with the results presented in
roughness of the PES-PDA@TiO2 substrate Ra = 21.4 nm) is greater Fig. S3. Additionally, the pure water flux of membrane substrates with
than that of pristine PES substrate Ra = 17.0 nm). It is thus established increasing PDA@TiO2 content is shown in Fig. S4. The improved Jw

Fig. 4. The water contact angle (a), TGA (b), Zeta potential (c), and AFM (d) results of the prepared membrane substrates.

6
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

Fig. 5. Surface images of prepared TFC-FO membranes.

Table 1
Separation properties of the membranes.
Membrane Water permeability, A (L/m2h bar) Salt rejection, R (%) Salt permeability, B (L/m2h) B/A (bar)

TFC 1.62 95.2 1.11 0.68


TFC-TiO2 2.21 93.1 1.22 0.55
TFC-PDA@TiO2 3.19 91.1 1.23 0.38

70 18 (b) (c) FO
FO (a) FO
0.6
PRO PRO PRO
60
15
0.5
50
12
0.4
Js (g/m2h)

40

Js/Jw (g/L)
Jw (L/m2h)

9
0.3
30
6
0.2
20

3 0.1
10

0 0 0.0
TFC TFC-TiO2 TFC-PDA@TiO2 TFC TFC-TiO2 TFC-PDA@TiO2
TFC TFC-TiO2 TFC-PDA@TiO2

Fig. 6. FO performance of the fabricated TFC, TFC-TiO2, and TFC-PDA@TiO2 membranes under FO and PRO mode (DS: 1 M NaCl and FS: DI water).

performance can be attributed to the following two reasons: firstly, the the salt leakage, and consequently, the greater the external concentra­
presence of hydrophilic groups such as hydroxyl and amino functional tion polarization (ECP) and ICP effect.
groups on the surface of PDA@TiO2, which improved hydrophilicity and
porosity, thus increasing water flux [30]; secondly, the incorporation of 3.5. Anti-fouling capacity
PDA@TiO2 particles increases the roughness of the membrane substrate
layer (Fig. 4d) contributing to higher surface area and improved water Organic anti-fouling tests were performed to evaluate the fouling
flux. The increase in water flux was consistent with the enhanced A capacity of FO membranes. Fig. 8 shows the normalized membrane
value. When comparing RO performance (Table 1) and FO performance, water flux (Jw/Jw0) of the TFC, TFC-TiO2, and TFC-PDA@TiO2 mem­
Js in FO (Fig. 6b) is well correlated with NaCl rejection in RO where a branes in AL-DS mode with respect to the SA model organic foulant. As
high removal rate corresponds to low Js and vice versa. The specific observed, the normalized water flux progressively decreased for all TFC
solute flux (Js/Jw ratio), an important factor for the selectivity of FO membranes compared to the model foulant as the transport resistance
membranes for both FO and PRO modes, is shown in Fig. 6c. In general, increased due to the accumulation of SA foulant on the membrane
a highly selective FO membrane should possess a low Js/Jw ratio value. substrates ultimately hindering the permeation of water through the
As observed, the improvement in selectivity indicated by the decrease in membranes [8]. The TFC-PDA@TiO2 membrane effectively repelled
the Js/Jw values (from 0.48 g/L of pristine TFC to 0.26 g/L for the negatively charged SA foulants and demonstrated a lower flux decline at
TFC-PDA @TiO2 FO) of the FO membrane by the incorporation of the end of the fouling study compared to the other two TFC membranes
PDA@TiO2 is due to the smaller B/A ratio of the TFC-PDA @TiO2 FO which can be attributed to the high anti-fouling capacity of the devel­
membrane. Taking into account the A, B, B/A values and the Js/Jw ratio oped TFC-PDA@TiO2 membrane. The TFC, TFC-TiO2, and TFC-PDA@­
of all TFC membranes from this study, the TFC-PDA@TiO2 FO mem­ TiO2 membranes reached normalized fluxes of 0.61, 0.67, and 0.72
brane seems the most promising for FO applications. respectively at the end of the 6 h fouling test. The flux recovery capacity
Meanwhile, the performance of the FO membrane as a function of the of the membrane was evaluated after a simple backwash. The pristine
DS concentration (0.5–2.0 M) was measured under both modes and is TFC FO membrane showed a recovery of ~86%, and in the case of
presented in Fig. 7. As shown in Fig. 7a and c, the Jw values rise with the TFC-FO membranes with nanocomposite substrates the recovery was
increasing DS (NaCl) concentration due to the increased osmotic driving found to be a maximum of ~92 and ~96% for TFC- TiO2 and
force released by the higher concentration of DS. Furthermore, an TFC-PDA@TiO2 respectively. As mentioned above, the incorporation of
enhancement in the Js and Js/Jw values with DS (NaCl) concentration hydrophilic PDA@TiO2 particles provides more hydrophilic properties
was also observed. Usually, the higher the DS concentration, the higher and larger pore size in the membrane substrate which leads to better

7
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

Fig. 7. FO performance of the TFC and TFC-PDA@TiO2 membranes at various concentrations of DS under FO (a-b) and PRO mode (c-d).

water permeability. Furthermore, the formation of a hydration layer on


the TFC membrane substrate due to PDA@TiO2 particles decreased the
adhesion of SA foulants to the surface of the membrane substrate which
would ultimately improve the anti-fouling performance of the mem­
brane. Furthermore, the anti-fouling capability may also be ascribed to
the electrostatic repulsion between -COOH in SA and -OH groups in
PDA@TiO2 composites (see the more significantly negatively charged
membrane substrate in Fig. 4c).

3.6. Application of TFC-PDA@TiO2 membrane for oily wastewater


treatment

To evaluate the potential use of FO for the reclamation of water from


emulsified oil/water solutions, 500 mg/L oil/water (O/W) and 0.1 g of
SDS added to S/O/W mixture solutions were used as FS and 1.0 M NaCl
as DS. Fig. 9a shows the normalized evolutionary behavior of the water
flux of the TFC-PDA@TiO2 membranes under PRO mode during 7 h tests
and compares them to the pristine TFC membrane. A gradual decline in
normalized flux over time was noticed for both solution mixtures. The
TFC membrane exhibited a continuous and severe flux reduction
compared to the TFC-PDA@TiO2 membrane and obtained a normalized
Fig. 8. Anti-fouling behavior of the TFC and TFC-PDA-TiO2 membranes (FS:
flux of 0.32 and 0.4 for the treatment of O/W and S/O/W mixtures
sodium alginate (SA) and CaCl2 with a concentration of 200 mg/L and 1 mM,
respectively. The TFC-PDA@TiO2 membrane had a higher normalized
respectively, DS: 1 M NaCl).
flux of 0.63 and 0.73 for the treatment of O/W and S/O/W mixtures
respectively. As discussed in previous sections, the improved hydro­
philicity, tight hydration layer, and enhanced negative charge of the
TFC-PDA@TiO2 membrane substrate layer plays an important role in
normalized flux (Fig. 9b) [34,37]. Furthermore, the S/O/W emulsion
contains an anionic SDS surfactant and may have a strong electrostatic

8
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

Fig. 9. (a) Normalized flux of TFC-FO and TFC-PDA@TiO2 FO with O/W and S/O/W mixture solutions over 7 h cycle operation. (b) Schematic representation of the
TFC-PDA@TiO2 FO membrane to anti-fouling mechanisms in the TFC-PDA@TiO2 FO membrane to avoid oil adhesion on the membrane surface. SEM images of the
pristine TFC (c) and TFC-PDA@TiO2 (d) membrane with S/O/W emulsion.

repulsion with a more negatively charged TFC-PDA@TiO2 membrane TFC-PDA@TiO2 membrane can be efficiently applicable for sustainable
substrate [38]. Therefore, the reduction in the flux of the FO membrane water reclamation of oily wastewater by FO in the PRO mode.
with the S/O/W mixture is less than with the O/W solution. Fouling of Table 2 displays a performance comparison between the results of
the pristine TFC and TFC-PDA@TiO2 membranes used with the S/O/W the present study and other reported works to better indicate fabricated
emulsion was further evaluated by SEM (Fig. 9c-d). As can be seen in the membrane performance.
SEM images, in contrast to the TFC-PDA@ TiO2 membrane (Fig. 9d), the
pristine TFC FO membrane (Fig. 9c) was covered by a fouling layer 4. Conclusions
containing some crystals after filtration of the S/O/W wastewater
demonstrating that both fouling and scaling had formed on the mem­ This work investigated the fabrication of FO TFC-PDA@TiO2 mem­
brane surface [39,40]. These results demonstrate that the prepared branes and their anti-fouling behavior in the treatment of oily

Table 2
Comparison of FO membrane performance.
FO membrane Water flux, Jw (L/m2h), FO/PRO Reverse solute flux, Js (g/m2h) Feed solution Draw solution Ref

CTA-NW (commercial) 4.4/8.19 0.6/2.8 10 mM NaCl 0.5 M NaCl [41]


CTA-HW (commercial) 9.03/15.4 5.3/9.4 10 mM NaCl 0.5 M NaCl [41]
CTA-W (commercial) 5.0/6.55 2.9/4.8 10 mM NaCl 0.5 M NaCl [41]
TFC-PES-GO/Ag 20/- 1.45/- DI water 0.7 NaCl [42]
TFC-PSf 8.9/16.1 1.7/4.0 10 mM NaCl 0.5 M NaCl [26]
TFC-PSf/TiO2 17.1/31.2 1.6/4.0 10 mM NaCl 0.5 M NaCl [26]
29.7/56.2 8.1/14.0 10 mM NaCl 2.0 M NaCl
TFC-PSf/f-CNFs 13.8/18.8 3.3/5.0 DI water 1.0 M NaCl [11]
TFC-PES 12.5/21.2 6.1/11.3 DI water 1.0 M NaCl Present work
TFC-PES/TiO2 26.5/47.5 8.8/14.5 DI water 1.0 M NaCl Present work
TFC-PES/ PDA@TiO2 34.3/60.1 9.2/16.5 DI water 1.0 M NaCl Present work

9
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

wastewater. A simple, facile, and eco-friendly approach was introduced References


to enhance the wettability of the TFC-FO membrane substrate by
incorporating PDA-coated TiO2 particles into the PES membrane sub­ [1] S. Munirasu, M.A. Haija, F. Banat, Use of membrane technology for oil field and
refinery produced water treatment—a review, Process Saf. Environ. Prot. 100
strate. The following are the main findings of this study: (2016) 183–202.
[2] Y. Liang, S. Kim, P. Kallem, H. Choi, Capillary effect in Janus electrospun nanofiber
1. The composite FO membrane substrate exhibited more hydrophi­ membrane for oil/water emulsion separation, Chemosphere 221 (2019) 479–485.
[3] H.J. Tanudjaja, C.A. Hejase, V.V. Tarabara, A.G. Fane, J.W. Chew, Membrane-
licity due to the hydroxyl and amino functional groups in the based separation for oily wastewater: a practical perspective, Water Res. 156
modified TiO2 nanocomposite resulting in an approximately 25% (2019) 347–365.
lower water contact angle and a higher porosity than the pristine [4] P. Kallem, I. Othman, M. Ouda, S.W. Hasan, I. AlNashef, F. Banat, Polyethersulfone
hybrid ultrafiltration membranes fabricated with polydopamine modified ZnFe2O4
membrane substrate. nanocomposites: applications in humic acid removal and oil/water emulsion
2. The water flux (Jw) of 34.3 L/m2h and 60.1 L/m2h was obtained for separation, Process Saf. Environ. Prot. 148 (2021) 813–824.
the resulting TFC-PDA@TiO2 FO membrane under FO and PRO [5] K. Lutchmiah, A. Verliefde, K. Roest, L.C. Rietveld, E. Cornelissen, Forward osmosis
for application in wastewater treatment: a review, Water Res. 58 (2014) 179–197.
modes, which was ~174% and ~183% higher than that of the
[6] J.R. McCutcheon, M. Elimelech, Influence of membrane support layer
pristine TFC-FO membrane under FO and PRO modes respectively, hydrophobicity on water flux in osmotically driven membrane processes,
due to the synergistic effect of hydrophilic TiO2 and PDA resulting in J. Membr. Sci. 318 (2008) 458–466.
the higher hydrophilicity of the hybrid TFC membrane. [7] P. Kallem, F. Banat, L. Yejin, H. Choi, High performance nanofiber-supported thin
film composite forward osmosis membranes based on continuous thermal-rolling
3. An improvement in selectivity, as indicated by the decrease in Js/Jw pretreated electrospun PES/PAN blend substrates, Chemosphere 261 (2020),
values (from 0.48 g/L of pristine TFC to 0.26 g/L for the TFC- 127687.
PDA@TiO2 FO) of the FO membrane by incorporation of PDA@TiO2, [8] G. Blandin, H. Vervoort, P. Le-Clech, A.R.D. Verliefde, Fouling and cleaning of high
permeability forward osmosis membranes, J. Water Process Eng. 9 (2016)
due to the smaller B/A ratio of the TFC-PDA@TiO2 FO membrane. 161–169.
4. More importantly, the TFC-PDA@TiO2 membrane exhibited better [9] N. Widjojo, T.-S. Chung, M. Weber, C. Maletzko, V. Warzelhan, The role of
fouling resistance (recovery was ~96%) with the organic foulant sulphonated polymer and macrovoid-free structure in the support layer for thin-
film composite (TFC) forward osmosis (FO) membranes, J. Membr. Sci. 383 (2011)
solution of the SA model. 214–223.
5. Considering the A, B, B/A values and Js/Jw ratio of all TFC mem­ [10] F. Lotfi, B. Samali, D. Hagare, Cleaning efficiency of the fouled forward osmosis
branes used in this study, the TFC-PDA@TiO2 FO membrane would membranes under different experimental conditions, J. Environ. Chem. Eng. 6
(2018) 4555–4563.
seem the most suitable for FO. [11] I. Tavakol, S. Hadadpour, Z. Shabani, M. Ahmadzadeh Tofighy, T. Mohammadi,
6. At the end of the 7 h test, the TFC and TFC-PDA@TiO2 membranes S. Sahebi, Synthesis of novel thin film composite (TFC) forward osmosis (FO)
achieved normalized fluxes of 0.32 and 0.63 with the O/W, 0.4 and membranes incorporated with carboxylated carbon nanofibers (CNFs), J. Environ.
Chem. Eng. 8 (2020), 104614.
0.73 with the S/O/W mixtures respectively indicating that the
[12] V. Nayak, J. Ms, R.G. Balakrishna, M. Padaki, V.Y. Zadorozhnyy, S.D. Kaloshkin, 4-
resulting TFC FO membrane can be efficiently used for sustainable aminophenyl sulfone (APS) as novel monomer in fabricating paper based TFC
water recovery from oily wastewater via FO in the PRO mode. composite for forward osmosis: selective layer optimization, J. Environ. Chem.
Eng. 8 (2020), 103664.
[13] P. Kallem, Development of polybenzimidazole and ionic liquid based membranes
This study, therefore, provides an alternative perspective for the for high temperature proton exchange membranes (PEMs) and gas separation
design of high-performance membranes with low fouling FO to treat oily applications Développement de membranes à base de polybenzimidazole et de
wastewater. liquides ioniques pour applications à haute température comme membranes
échangeuses de protons (PEMs) et pour la séparation de gaz, in, Université
Montpellier Universidad de Zaragoza (Espagne), 2017.
CRediT authorship contribution statement [14] P. Kallem, M.P. Pina, M. Drobek, A. Julbe, K. Nijmeijer, J. Roziere, D. Jones, R.
Mallada, Polybenzimidazole (PBI) microsieves produced by liquid induced phase
separation microfabrication (LIPSμF): Polymeric ionic liquid/PBI membranes for
Parashuram Kallem: Conceptualization, Validation, Formal anal­ high temperature fuel cell applications, in: Euromembrane 2015, Aachen,
ysis, Investigation, Data curation, Writing – original draft, Writing – Germany, 2015.
review & editing, Visualization Ravi P. Pandey: Validation, Data cura­ [15] P. Kallem, C. Charmette, M. Drobek, A. Julbe, R. Mallada, M.P. Pina, Exploring the
gas-permeation properties of proton-conducting membranes based on protic
tion, Writing – review & editing. Hanaa M. Hegab: Validation, Data imidazolium ionic liquids: application in natural gas processing, Membranes 8
curation, Writing – review & editing. Ruchi Gaur: Validation, Data (2018).
curation, Writing – review & editing. Shadi W. Hasan: Resources, [16] P. Kallem, R. Gaur, R.P. Pandey, S.W. Hasan, H. Choi, F. Banat, Thin film
composite forward osmosis membranes based on thermally treated PAN
Writing – review & editing. Fawzi Banat: Conceptualization, Valida­ hydrophilized PVDF electrospun nanofiber substrates for improved performance,
tion, Resources, Writing – review & editing, Visualization, Project J. Environ. Chem. Eng. 9 (2021), 106240.
administration, Funding acquisition. [17] M. Ghanbari, D. Emadzadeh, W. Lau, H. Riazi, D. Almasi, A. Ismail, Minimizing
structural parameter of thin film composite forward osmosis membranes using
polysulfone/halloysite nanotubes as membrane substrates, Desalination 377
(2016) 152–162.
Declaration of Competing Interest [18] N. Maximous, G. Nakhla, W. Wan, K. Wong, Preparation, characterization and
performance of Al2O3/PES membrane for wastewater filtration, J. Membr. Sci. 341
The authors declare that they have no known competing financial (2009) 67–75.
[19] P. Kallem, Y. Ibrahim, S.W. Hasan, P.L. Show, F. Banat, Fabrication of novel
interests or personal relationships that could have appeared to influence polyethersulfone (PES) hybrid ultrafiltration membranes with superior
the work reported in this paper. permeability and antifouling properties using environmentally friendly sulfonated
functionalized polydopamine nanofillers, Sep. Purif. Technol. 261 (2021), 118311.
[20] M. Sun, Y. Su, C. Mu, Z. Jiang, Improved antifouling property of PES ultrafiltration
Acknowledgment membranes using additive of silica-PVP nanocomposite, Ind. Eng. Chem. Res. 49
(2010) 790–796.
This research was supported by a grant (Award No. RC2- 2018-009) [21] T. Sirinupong, W. Youravong, D. Tirawat, W.J. Lau, G.S. Lai, A.F. Ismail, Synthesis
and characterization of thin film composite membranes made of PSF-TiO2/GO
from the Center for Membranes and Advanced Water Technology nanocomposite substrate for forward osmosis applications, Arab. J. Chem. 11
(CMAT) funded by Khalifa University of Science and Technology in Abu (2018) 1144–1153.
Dhabi (UAE). [22] S. Zarghami, T. Mohammadi, M. Sadrzadeh, B. Van der Bruggen, Bio-inspired
anchoring of amino-functionalized multi-wall carbon nanotubes (N-MWCNTs) onto
PES membrane using polydopamine for oily wastewater treatment, Sci. Total
Appendix A. Supporting information Environ. 711 (2020), 134951.
[23] X.-p Wang, J. Hou, F.-s Chen, X.-m Meng, In-situ growth of metal-organic
framework film on a polydopamine-modified flexible substrate for antibacterial
Supplementary data associated with this article can be found in the and forward osmosis membranes, Sep. Purif. Technol. 236 (2020), 116239.
online version at doi:10.1016/j.jece.2022.107454.

10
P. Kallem et al. Journal of Environmental Chemical Engineering 10 (2022) 107454

[24] Y. Jiang, K. Shi, H. Tang, Y. Wang, Enhanced wettability and wear resistance on [33] H.-g Choi, M. Son, S. Yoon, E. Celik, S. Kang, H. Park, C.H. Park, H. Choi, Alginate
TiO2/PDA thin films prepared by sol-gel dip coating, Surf. Coat. Technol. 375 fouling reduction of functionalized carbon nanotube blended cellulose acetate
(2019) 334–340. membrane in forward osmosis, Chemosphere 136 (2015) 204–210.
[25] S.B. Teli, S. Molina, A. Sotto, E.G. Calvo, Jd Abajob, Fouling resistant [34] W.J. Lee, P.S. Goh, W.J. Lau, C.S. Ong, A.F. Ismail, Antifouling zwitterion
polysulfone–PANI/TiO2 ultrafiltration nanocomposite membranes, Ind. Eng. embedded forward osmosis thin film composite membrane for highly concentrated
Chem. Res. 52 (2013) 9470–9479. oily wastewater treatment, Sep. Purif. Technol. 214 (2019) 40–50.
[26] D. Emadzadeh, W.J. Lau, A.F. Ismail, Synthesis of thin film nanocomposite forward [35] H. Wu, Y. Liu, L. Mao, C. Jiang, J. Ang, X. Lu, Doping polysulfone ultrafiltration
osmosis membrane with enhancement in water flux without sacrificing salt membrane with TiO2-PDA nanohybrid for simultaneous self-cleaning and self-
rejection, Desalination 330 (2013) 90–99. protection, J. Membr. Sci. 532 (2017) 20–29.
[27] D. Emadzadeh, W.J. Lau, T. Matsuura, M. Rahbari-Sisakht, A.F. Ismail, A novel thin [36] D. Yang, S. Huang, Y. Wu, M. Ruan, S. Li, Y. Shang, X. Cui, Y. Wang, W. Guo,
film composite forward osmosis membrane prepared from PSf–TiO2 Enhanced actuated strain of titanium dioxide/nitrile-butadiene rubber composite
nanocomposite substrate for water desalination, Chem. Eng. J. 237 (2014) 70–80. by the biomimetic method, RSC Adv. 5 (2015) 65385–65394.
[28] K. Parashuram, S.K. Maurya, H.H. Rana, P.S. Singh, P. Ray, A.V.R. Reddy, Tailoring [37] P. Li, S.S. Lim, J.G. Neo, R.C. Ong, M. Weber, C. Staudt, N. Widjojo, C. Maletzko, T.
the molecular weight cut off values of polyacrylonitrile based hollow fibre S. Chung, Short- and long-term performance of the thin-film composite forward
ultrafiltration membranes with improved fouling resistance by chemical osmosis (TFC-FO) hollow fiber membranes for oily wastewater purification, Ind.
modification, J. Membr. Sci. 425–426 (2013) 251–261. Eng. Chem. Res. 53 (2014) 14056–14064.
[29] A. Sotto, A. Boromand, R. Zhang, P. Luis, J.M. Arsuaga, J. Kim, B. Van der Bruggen, [38] J. Li, Q. Wang, L. Deng, X. Kou, Q. Tang, Y. Hu, Fabrication and characterization of
Effect of nanoparticle aggregation at low concentrations of TiO2 on the carbon nanotubes-based porous composite forward osmosis membrane: Flux
hydrophilicity, morphology, and fouling resistance of PES–TiO2 membranes, performance, separation mechanism, and potential application, J. Membr. Sci. 604
J. Colloid Interface Sci. 363 (2011) 540–550. (2020), 118050.
[30] G. Han, S. Zhang, X. Li, N. Widjojo, T.-S. Chung, Thin film composite forward [39] R.R. Gonzales, L. Zhang, Y. Sasaki, W. Kushida, H. Matsuyama, H.K. Shon, Facile
osmosis membranes based on polydopamine modified polysulfone substrates with development of comprehensively fouling-resistant reduced polyketone-based thin
enhancements in both water flux and salt rejection, Chem. Eng. Sci. 80 (2012) film composite forward osmosis membrane for treatment of oily wastewater,
219–231. J. Membr. Sci. 626 (2021), 119185.
[31] Z. Xu, J. Zhang, M. Shan, Y. Li, B. Li, J. Niu, B. Zhou, X. Qian, Organosilane- [40] C. Ding, X. Zhang, S. Xiong, L. Shen, M. Yi, B. Liu, Y. Wang, Organophosphonate
functionalized graphene oxide for enhanced antifouling and mechanical properties draw solution for produced water treatment with effectively mitigated membrane
of polyvinylidene fluoride ultrafiltration membranes, J. Membr. Sci. 458 (2014) fouling via forward osmosis, J. Membr. Sci. 593 (2020), 117429.
1–13. [41] J. Wei, C. Qiu, C.Y. Tang, R. Wang, A.G. Fane, Synthesis and characterization of
[32] S. Daer, N. Akther, Q. Wei, H.K. Shon, S.W. Hasan, Influence of silica nanoparticles flat-sheet thin film composite forward osmosis membranes, J. Membr. Sci. 372
on the desalination performance of forward osmosis polybenzimidazole (2011) 292–302.
membranes, Desalination 491 (2020), 114441. [42] A.F. Faria, C. Liu, M. Xie, F. Perreault, L.D. Nghiem, J. Ma, M. Elimelech, Thin-film
composite forward osmosis membranes functionalized with graphene oxide–silver
nanocomposites for biofouling control, J. Membr. Sci. 525 (2017) 146–156.

11

You might also like