You are on page 1of 32

Author’s Accepted Manuscript

Thin-film composite (TFC) hollow fiber membrane


with double-polyamide active layers for internal
concentration polarization and fouling mitigation in
osmotic processes

Gang Han, Zhen Lei Cheng, Tai-Shung Chung


www.elsevier.com/locate/memsci

PII: S0376-7388(16)31597-6
DOI: http://dx.doi.org/10.1016/j.memsci.2016.10.022
Reference: MEMSCI14808
To appear in: Journal of Membrane Science
Received date: 12 September 2016
Revised date: 7 October 2016
Accepted date: 14 October 2016
Cite this article as: Gang Han, Zhen Lei Cheng and Tai-Shung Chung, Thin-film
composite (TFC) hollow fiber membrane with double-polyamide active layers for
internal concentration polarization and fouling mitigation in osmotic processes,
Journal of Membrane Science, http://dx.doi.org/10.1016/j.memsci.2016.10.022
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Thin-film composite (TFC) hollow fiber membrane with double-polyamide active layers for

internal concentration polarization and fouling mitigation in osmotic processes

Gang Han, Zhen Lei Cheng, Tai-Shung Chung*

Department of Chemical and Biomolecular Engineering National University of Singapore,

Singapore 117585

*
Corresponding author. Tel.: +65-65166645; fax.: +65-67791936; chencts@nus.edu.sg

Abstract

Internal concentration polarization (ICP) and severe irreversible fouling occurring within the

porous and tortuous substrates of the forward osmosis (FO) and pressure retarded osmosis (PRO)

membranes have significantly hidden their applications for water purification and osmotic power

generation. This study experimentally demonstrates that designing a double-skin structure in FO

and PRO membranes can effectively control their ICP and fouling propensity. Thin-film

composite (TFC) hollow fiber membranes consist of an inner polyamide selective skin and an

outer polyamide sealing layer were successfully fabricated by double interfacial polymerizations

on a tailored polyethersulfone (PES) fiber substrate (termed as dTFC-PES). Due to the

outstanding rejection of the outer polyamide sealing layer, the penetration of inorganic salts and

foulants into the substrate is sufficiently blocked. As a result, not only ICP and fouling inside the

membrane are effectively minimized, but also sustainable FO and PRO performances are

achieved. By using real wastewater contains multiple inorganic salts and organic foulants as the

1
feed, the dTFC-PES membrane shows a quite low flux decline of 29% in FO operations under

the PRO mode at an ultrahigh feed recovery of 80%. Under PRO tests for power generation, the

membrane power density slightly drops to 90.8% of the initial value after a 12-h test at ΔP=15

bar. In addition, since foulants are primarily accumulated on the surface of the polyamide sealing

layer, physically flushing the fouled membrane surface by either freshwater or commercial

cleaner Genesol 704 can efficiently restore the water flux back to its initial level with a recovery

rate of 87% or 98%, respectively. This study may offer useful insights and meaningful strategies

for the development of effective antifouling FO and PRO membranes.

Keywords: antifouling; double-skin; TFC FO and PRO membrane; water reuse; osmotic power

1. Introduction

Forward osmosis (FO) and pressure retarded osmosis (PRO) have gained renewed attention

recently for water reuse, wastewater treatment and renewable osmotic power generation [1-7].

Both processes utilize the osmotic pressure gradient (Δπ) between the solutions with different

salinities to induce water permeation across a semipermeable membrane. When the applied

hydraulic pressure on the high salinity solution is lower than the osmotic pressure gradient (ΔP <

Δπ) across the semipermeable membrane, water spontaneously flows from the low salinity

solution to the high salinity one (which is referred to the draw solution thereafter). The salinity

gradient energy can be therefore harvested when discharging the inflow high pressure water via a

turbine or energy recovery device [8-10]. Most recent PRO researches focus on the mixing of

brine from reverse osmosis (RO) plants and retentate from municipal wastewater plants because

it can generate much higher osmotic energy than the feed pair of seawater and river water

2
[1,4,11-15]. In addition, it converts two wastes (i.e., RO brine and wastewater) into useful energy

and mitigates their disposal and environmental issues.

Significant progresses have been made in the fabrication of high performance osmotic

membranes with more desirable structure and transport properties recently [16-18]. However,

their FO and PRO performances are severely impeded by internal concentration polarization

(ICP) and membrane fouling particularly when using real wastewater as the feed [19-21]. Under

the PRO mode, the porous substrate of the FO and PRO membranes faces the waste stream. As a

result, the inorganic salts and foulants are easily brought deep inside the substrate by the

convective water flow and get stuck underneath the rejection layer. This would cause severe ICP

and fouling and thus reduce water flux dramatically and rapidly [22-24]. Different from the

fouling on the surface of the selective layer, sophisticated washing techniques and long cleaning

durations are required to remove the foulants underneath the selective layer because of the

substrate tortuosity. Feed pretreatment can reduce fouling, however it would increase energy and

chemical consumptions as well as operation costs. Manipulation of membrane structural

properties with enhanced antifouling ability is a promising strategy to mitigate ICP and fouling

[3,17,25].

By leveraging the momentum of developing effective antifouling membranes for FO and PRO

applications, one novel thin-film composite polyethersulfone (dTFC-PES) hollow fiber

membrane with an advanced double-selective-skin structure has been successfully developed in

the current study. In fact, the double-skin concept was pioneered by Wang et al in order to take

advantages of the high flux nature of FO membranes under the PRO mode and to mitigate the

internal fouling [26]. The original membranes were fabricated by a phase inversion method and

consisted of a dense skin facing the draw solution and a loose dense skin facing the feed to reject

3
foulants. Zhang et al experimentally demonstrated the reduced colloidal fouling propensity [27],

while Tang et al mathematically confirmed the ICP mitigation [28]. Since then, several double-

skin FO membranes with variable barrier layers have been developed [29-34]. However, all of

these reported double-skin membranes have a dense skin made from either phase inversion or

thin-film interfacial polymerization method, but their loose dense skins were from phase

inversion, dip coating or layer-by-layer (LbL) deposition method. As a result, the loose dense

skins normally possess relatively low rejections, particularly to inorganic salts and small foulants.

In addition, these membranes were never tested using real wastewater as the feed, neither were

they tested for osmotic power generation. In contrast, the newly developed dTFC-PES membrane

was prepared from double thin-film interfacial polymerizations on a PES hollow fiber substrate.

The hollow fiber configuration was utilized in this study because of its high surface area per

module and self-mechanical support characteristics [35,36]. It does not need spacers in module

fabrication so that it eliminates the issue of spacer induced membrane deformation observed for

flat-sheet PRO modules under high pressures [37]. The newly developed dTFC-PES membrane

was characterized and tested using real wastewater brine containing multiple inorganic and

organic foulants as the feed. Their FO performance for water reuse, PRO for power generation,

as well as fouling propensity, reversibility and cleaning were systematically investigated. The

current work may open up new perspectives and design strategies for the fabrication of effective

antifouling membranes for FO and PRO applications.

4
2. Materials and methods

2.1 Materials

Radels® A polyethersulfone (PES, Solvay Advanced Polymer, L.L.C., GA), N-methyl-2-

pyrrolidone (NMP, 99.5%, Merck), polyethylene glycol with a molecular weight of 400 Da

(PEG400, Sigma-Aldrich) and deionized water were used as the polymer, solvent, and non-

solvent additive for the preparation of the PES hollow fiber substrate. Polyethylene glycol (PEG)

and polyethylene oxide (PEO) with different molecular weights from Merck were utilized to

characterize the molecular weight cut-off (MWCO), mean pore size and pore size distribution of

the as-spun PES fiber substrate. Ethylene glycol, diethylene glycol, triethylene glycol, and

glucose from Sigma-Aldrich were employed to measure the mean pore size and pore size

distribution of the polyamide modified PES hollow fiber membrane. A 50/50 (wt%) mixture of

glycerol (Industrial grade, Aik Moh Pains & Chemicals Pte. Ltd, Singapore) and water was

applied to post treat the as-spun PES hollow fibers. Trimesoyl chloride (TMC) with 98% purity

and m-phenylenediamine (MPD) with >99% purity ordered from Sigma–Aldrich were acquired

as the monomers for the interfacial polymerization reaction. N-hexane from Merck with >99%

purity was used as the solvent for TMC. Sodium dodecyl sulfate (SDS, >97%) from Fluka was

employed as an additive in the MPD aqueous solution. The commercial membrane cleaner

Genesol 704 was provided by Genesys International Ltd., UK for membrane cleaning. Sodium

chloride (NaCl) from Merck was employed for the preparation of synthetic draw solutions.

Deionized (DI) water with a resistivity of 15 MΩ cm was produced by a Milli-Q unit (Millipore).

5
2.2 Fabrication of dTFC-PES hollow fiber membrane

The PES hollow fiber substrate was firstly fabricated via an optimized dry-jet wet spinning

process [38,39]. After that, a polyamide skin was deposited onto the outer and inner surfaces of

the PES fiber via interfacial polymerization between m-phenylenediamine (MPD) and trimesoyl

chloride (TMC), separately. The detailed procedures and conditions for PES fiber spinning,

interfacial polymerization and module fabrication are described in Supporting Information (SI)

and documented in our previous works [36,38,39].

2.3 Membrane characterizations

Membrane morphology was characterized by a field-emission scanning electron microscope

(FESEM), and the surface chemistry was obtained by X-ray photoelectron spectroscopy (XPS).

Membrane porosity of the PES fiber substrate was measured using a well-established method

descried elsewhere [36,38]. The mean pore size, pore size distribution and molecular weight cut-

off (MWCO) of the fiber substrates were determined by solute rejection experiments as

described in SI [40,41]. Pure water permeability coefficient (A) and salt rejection (R) of the

dTFC-PES membrane were measured by testing the membranes under the reverse osmosis (RO)

mode (solutions flow through the fiber lumen side) via a lab-scale filtration apparatus; while the

membrane salt permeability coefficient (B) and structural parameter (S) were subsequently

obtained following the methods described in SI and elsewhere [39,42]. In order to ensure the

experimental reproducibility, three tests were carried out for each condition and the averaged

value was reported. During each test, one new membrane coupon was used.

6
2.4 FO and PRO tests

The FO and PRO performance tests were carried out via a lab-scale crossflow PRO setup

reported in our previous work [38,39]. For all the tests, the draw solution was flowed against the

inner polyamide selective layer of the dTFC-PES membrane. The flowrate of the draw solution

through the fiber lumen side was kept at 2.2 m/s (Re = 2434) with an adjustable hydraulic

pressure of 0-25 bar. The feed solution was counter-currently circulated through the shell side of

the module at a flowrate of 0.13 m/s. There was no hydraulic pressure difference (ΔP=0 bar)

across the fiber during FO tests; while a hydraulic pressure of 15 bar (ΔP=15 bar) was applied on

the draw solution side in PRO operations by a high pressure pump.

2.5 Fouling experiments and membrane cleaning

To characterize membrane fouling, wastewater brine (termed as WWBr) containing multiple

inorganic and organic foulants from a local water recycling plant was employed as the feed

[20,21]. An 1M NaCl solution was used as the draw solution. The fouling tests were started with

1 L WWBr feed and stopped when the cumulative permeate volume or the testing duration

reached the predetermined values. Baseline experiments were also performed under the same

operating conditions but using deionized water as the feed. After the fouling experiments, the

fouled membrane was immediately cleaned by physical rinse of deionized water along the fouled

surface for 6 h to determine the fouling reversibility. Membrane cleaning was conducted by

flushing the fouled surface with a commercial membrane cleaner Genesol 704 (1 wt%, pH=11)

for 0.5 h. Fouling reversibility and cleaning efficiency were assessed in terms of membrane

initial water flux recovery (i.e., the averaged flux during the first 10 min) using deionized water

as the feed. The detailed descriptions of the FO/PRO setup, the WWBr feed, and procedures for

7
the determination of water flux and power density, and membrane cleaning are disclosed in SI.

Each fouling and cleaning test was carried out three times and the averaged data were reported.

In order to make the figures clearly visible, the error bar was not included in the reported figures.

3. Results and discussion

3.1 Characterizations of the PES hollow fiber substrate

In order to obtain the dTFC-PES membrane with a desirable morphology and properties, the PES

hollow fiber substrate was molecularly tailored via optimizing the dope composition and

spinning conditions. As depicted in Fig. 1, the PES fiber substrate shows a highly centric

structure with an outer diameter of 1125 µm and a cross-section thickness of 300 μm (Table 1).

By employing water as the bore fluid and external coagulant to induce a fast phase inversion,

both the inner and outer surfaces of the PES fiber are relatively smooth and dense without any

large surface holes. Table 1 also shows that the PES fiber possesses a small mean pore size of

11.6 nm in diameter with a relatively narrow pore size distribution as indicated by the small

geometric standard deviation (σp = 1.9). These surface morphology and pore feature are critical

for the subsequent formation of a thin and less defective polyamide selective layer on top and

bottom of the PES substrate via interfacial polymerization [38,43,44]. The highly porous cross-

section morphology consists of two layers of finger-like macrovoids because of instantaneous

demixing and non-solvent intrusion during phase inversion from both sides. This cross-section

morphology together with the great membrane porosity of about 74.4% (Table 1), not only

facilitate the water transports across the substrate but also help reduce ICP. As a result, the PES

fiber substrate has a MWCO of 183.2 kDa and a large PWP of 130 L m-2 h-1 bar-1 as tabulated in

Table 1. However, a relatively thick sublayer with fully sponge-like macrovoid-free structure is

8
observed underneath the fiber inner surface, which is favorable for the formation of a robust TFC

layer [36,43-45].

3.2 Characterizations of the dTFC-PES hollow fiber membrane

Fig. 2 shows the surface and cross-section morphologies of the double-skin dTFC-PES hollow

fiber membrane. A typical “ridge-and-valley” polyamide morphology was observed on both

inner and outer surfaces of the PES fiber substrate with an estimated thickness of around 375 nm

and 470 nm, respectively. The inner polyamide layer is the selective layer while the outer one is

utilized as the sealing layer to mitigate ICP and internal fouling. To understand the pore size and

its distribution of the outer sealing layer, a polyamide layer was purposely interfacial-

polymerized on the outer surface of the PES fiber substrate. Fig. 3 displays its pore size

distribution. Compared to the original PES substrate of 11.6 nm in diameter (Table 1), the

modified PES substrate has a very small mean pore diameter of around 0.7 nm with a narrow

pore size distribution. Thus, inorganic salts and foulants in the feed may be sufficiently

prevented from entering the substrate with the aid of this outer polyamide layer [26-28,33].

Table 2 summarizes the intrinsic transport properties of the dTFC-PES membrane in terms of

pure water permeability (A), NaCl rejection (R), and salt permeability coefficient (B). Due to the

hydrophilic and ultrathin nature of the polyamide layers, a high A of 1.5 L m-2 h-1 bar-1 was

acquired. In addition, the dTFC-PES membrane possesses a great R of 94.2% at a low pressure

of 1 bar and an impressively low B of 0.02 L m-2 h-1, suggesting the outstanding membrane

selectivity. The relatively high A and ultralow B values are crucial to achieve a high water flux in

FO and PRO operations. Owing to the porous finger-like macrovoid structure in the fiber cross-

section, the dTFC-PES membrane has a reasonably small structural parameter S of 996 μm even

9
though it possesses a thick cross-section of 300 μm. Most importantly, the membrane shows

remarkable mechanical strength which can withstand a hydraulic pressure of larger than 20 bar.

The superior membrane structure, transport characteristics and mechanical strength make the

dTFC-PES as a desirable membrane for FO and PRO applications.

3.3 Fouling of dTFC-PES membrane in FO and PRO operations

By using the wastewater brine (WWBr) as the feed and 1M NaCl as the draw solution, fouling

phenomena and propensity of the double-skin dTFC-PES membrane were investigated under the

PRO mode for FO processes and osmotic power generation. Deionized water is also used as a

feed for benchmarking purpose. Fig. 4 (a) shows a comparison of the normalized water flux,

Jw/Jw0, as a function of feed permeate volume (or feed recovery) in FO operations. When using

deionized water as the feed, the water flux slightly drops to 7% of the initial value at a relatively

low permeate volume (i.e., less than 150 ml) followed by a mild decay till to the end of the test

with a total flux reduction of 18% at the feed recovery of 80% (i.e., permeate volume of 800 ml).

This flux decline is mainly attributed to the combined effects of reverse salt flux, concentration

of the feed, and ICP. When WWBr is used as the feed, the newly developed dTFC-PES

membrane also displays a slow and mild flux decline to 8% of the initial value at a relatively low

permeate volume of 150 ml followed by a mild decrease till to the end of the test with a total flux

reduction of 29% at the feed recovery of 80% (i.e., permeate volume of 800 ml). Even though

the normalized flux at 80% feed recovery declines from 18% to 29% when replacing deionized

water by WWBr as the feed, this antifouling performance against WWBr surpasses almost all

TFC FO membranes reported in the open literatures [3-5,17,30,32].

10
Fig. 4 (b) shows the power density (W) of the dTFC-PES membrane as a function of operation

time within a 12-h PRO test at ΔP=15 bar using 1M NaCl as the draw solution. Both deionized

water and WWBr are used as the feeds. The initial power density (W) is 10.7 W/m2 when

deionized water is employed as the feed. W shows negligible decrease with time because of the

great membrane selectivity and reduced ICP by the double-skin structure. When changing the

feed to WWBr, the initial W slightly drops to 9.8 W/m2 because of the increased feed salinity

(i.e., WWBr has a small salinity, thus the osmotic driving force across the membrane is reduced).

Interestingly, the decline of power density with time is very mild where W slightly drops to 8.9

W/m2 during a 12-h test. This corresponds to a small power density reduction of 9.2%. Clearly,

ICP and fouling caused by the WWBr feed in FO and PRO can be effectively mitigated by the

newly designed double-skin dTFC-PES hollow fiber membrane. Thus, sustainable FO and PRO

operations could be achievable.

3.4 Antifouling mechanisms of the dTFC-PES membrane

After the fouling tests, the outer surface of the fouled dTFC-PES membrane was characterized.

As displayed in Fig. S1, a dark brown fouling layer is clearly observed on the membrane outer

surface, suggesting that fouling induced by the WWBr feed is quite significant. The FESEM

images illustrated in Fig. 5 (a) further show that the outer surface is fully covered by a dense

fouling film with a large thickness. Since the pristine outer polyamide layer has a thickness of

470 nm (Fig. 2), the estimated thickness of the attached fouling layer is around 3530 nm. Fig. 6

compares the XPS data between the pristine and fouled membrane surface. The signals of Na, Cl,

Ca, P and Si elements on the fouled membrane surface are quite significant, which suggest the

existence of both inorganic fouling and silica scaling. However, Fig. S2 shows that the

11
membrane inner surface morphology has negligible changes before and after fouling tests. These

indicate that the outer polyamide sealing layer can effectively reject the salts and foulants and

prevent them from entering the substrate layer, making the fouling mainly occur on the

membrane surface.

Fig. 7 elucidates the fouling mechanisms of conventional TFC FO membranes and the newly

developed double-skin dTFC-PES membrane. Due to the outstanding rejection of the polyamide

skin facing the draw solution and the transmembrane water permeation from the feed side,

insignificant fouling is observed on the draw solution side [46,47]. However, the inorganic salts

and foulants of WWBr feed can easily penetrate into the porous and tortuous substrate and

accumulate there because of the water permeation across the membrane [20,21,46]. It not only

results in severe pore blockage, ICP, irreversible fouling, and thus dramatic flux and power

density drops [20,21,48-50], but also brings about difficulties to clean up the membrane. By

introducing a sealing layer with outstanding selectivity on the back side of the TFC membrane,

one can easily block the entry of inorganic salts and foulants, as illustrated in Fig. 7 (b). As a

result, fouling primarily takes place on the polyamide surface. Although external concentration

polarization (ECP) might be enhanced, its effects on FO and PRO performance are much lower

than those induced by ICP and fouling within the substrate. In addition, as the fouling occurs on

membrane surface, the shear force of physical rinsing can effectively flush off the deposited

foulants and satisfactorily recover the membrane flux.

3.5 Fouling reversibility and membrane cleaning

After the fouling tests, fouling reversibility of the dTFC-PES membrane was evaluated by

immediately rinsing the fouled membrane surface with deionized water for 6 h. The initial water

12
fluxes, Jw,DI, of the fouled and rinsed dTFC-PES membranes were measured using deionized

water as the feed and then normalized by the initial flux, Jw0,DI, of a fresh membrane. As shown

in Fig. 8 (a), Jw,DI of the fouled membrane drops to 60% of Jw0,DI after the fouling test in FO at a

high feed recovery of 80%. Compared to the conventional TFC membranes reported in the

literatures [3-7], the dTFC-PES membrane exhibits a much lower flux decline under the same

conditions, indicting its good antifouling properties. Physically rinsing the fouled surface by

deionized water can considerately clean the membrane and restore the water flux back to 87% of

the original flux, suggesting the great fouling reversibility of the double-skin membrane. Figure

S1 shows that the color of the dTFC-PES fibers is restored back to its original gray color after

water rinse. Similarly, Fig. 5 indicates that the thickness of the fouling layer is significantly

reduced from 3530 nm to 230 nm after water rinsing. The XPS spectra displayed in Fig. 8 (b)

further confirm that the inorganic salts are removed from the outer surface by water rinse.

Clearly, the shear force of physical water rinsing can readily remove the foulants.

However, compared to the pristine dTFC-PES membrane (Fig. 2), a certain amount of foulants is

still found on the polyamide sealing layer (Fig. 5). This might be due to the relatively rough

surface of the polyamide skin and the negatively charged carboxylate groups which may enhance

the foulant-surface interactions. In order to further improve the cleaning efficiency and recover

the membrane flux, chemical cleaning was also applied to regenerate the fouled membrane

[46,47]. As shown in Fig. 8, flushing the fouled surface with a commercial membrane cleaner,

Genesol 704 (1 wt%, pH=11), exhibits outstanding effectiveness in removing foulants. The water

flux is restored back to its initial level (i.e., the flux recovery reaches 98%). The intermittent

cycle tests presented in Fig. 9 also confirm the outstanding fouling reversibility and great

effectiveness of chemical cleaning. These demonstrate that the double-skin structure of the

13
newly developed dTFC-PES membrane not only can effectively mitigate fouling in FO and PRO

but also make the fouling reversible and easy to be cleaned.

4. Conclusions

By creating a polyamide sealing layer on the back side of the substrate, one effective approach

has been demonstrated to mitigate ICP and fouling within the porous substrate layer of

asymmetric FO and PRO membranes. The newly developed double-skin TFC polyethersulfone

hollow fiber membrane (dTFC-PES) possesses an inner polyamide selective skin and an outer

polyamide sealing layer with a mean pore size of 0.7 nm in diameter. Due to the outstanding

rejection of the polyamide sealing layer, the entry of inorganic salts and foulants into the

substrate could be effectively blocked, making the fouling primarily happens on the membrane

surface. By using real wastewater brine (WWBr) as the feed, superior antifouling performance

has been demonstrated by the dTFC-PES membrane. The membrane water flux slightly drops to

71% of the initial value at a high feed recovery of 80% in FO under the PRO mode. In PRO, the

membrane also shows a quite stable performance where the power density only decreases to 90.8%

of the initial value after a 12-h test at ΔP=15 bar. Since the fouling primarily happens upon the

surface of the outer polyamide layer instead of within the porous and tortuous substrate, the

fouling layer shows great reversibility. Flushing the fouled surface by deionized water and

commercial cleaner Genesol 704 can effectively restore the membrane flux to its initial level

with a high flux recovery of 87% and 98%, respectively. To the best of our knowledge, the

achieved antifouling performances of the dTFC-PES are superior to other membranes reported in

the literature. In summary, the current study demonstrates the advantages of a double-skin

14
membrane structure made from double interfacial polymerizations to inhibit ICP and fouling in

FO and PRO applications.

Supporting Information

PES hollow fiber spinning, post-treatment and module fabrication; interfacial polymerization;

membrane characterization methods, forward osmosis (FO) and pressure retarded osmosis (PRO)

setup and operating conditions and performance evaluation; optical (Fig. S1) and SEM images

(Fig. S2) of the fouled and cleaned membranes.

Acknowledgments

This research grant is supported by the Singapore National Research Foundation under its

Environment & Water Research Programme and administered by PUB, Singapore’s national

water agency, under the project titled “Membrane development for osmotic power generation,

part 1. Materials development and membrane fabrication” (1102-IRIS-11-01) and the NUS Grant

no. of R-279-000-381-279; and “Membrane development for osmotic power generation, Part 2.

Module fabrication and system integration” (1102-IRIS-11-02) and NUS grant no. of R-279-000-

382-279.

References

[1] T.S. Chung, X. Li, R.C. Ong, Q. Ge, H. Wang, G. Han, Emerging forward osmosis (FO)

technologies and challenges ahead for clean water and clean energy applications, Curr. Opin.

Chem. Eng. 1 (2012) 246–257.

15
[2] B.E. Logan, M. Elimelech, Membrane-based processes for sustainable power generation

using water, Nature 488 (2012) 313–319.

[3] C. Klaysom, T.Y. Cath, T. Depuydt, I.F.J. Vankelecom, Forward and pressure retarded

osmosis: potential solutions for global challenges in energy and water supply, Chem. Soc.

Rev. 42 (2013) 6959–89.

[4] T.S. Chung, L. Luo, C.F. Wan, Y. Cui, G. Amy, What is next for forward osmosis (FO) and

pressure retarded osmosis (PRO), Sep. Purif. Technol. 156 (2015) 856–860.

[5] K. Lutchmiah, A.R.D. Verliefde, K. Roest, L.C. Rietveld, E.R. Cornelissen, Forward osmosis

for application in wastewater treatment: A review, Water Res. 58 (2014) 179–197.

[6] G. Han, C.Z. Liang, T.S. Chung, M. Weber, C. Staudt, C. Maletzko, Combination of forward

osmosis (FO) process with coagulation/flocculation (CF) for potential treatment of textile

wastewater, Water Res. 91 (2016) 361–370.

[7] S. Zhao, L. Zou, C.Y. Tang, D. Mulcahy, Recent developments in forward osmosis:

Opportunities and challenges, J Membr Sci, 396 (2012) 1–21.

[8] S. Loeb, R.S. Norman, Osmotic power plant, Science 189 (1975) 654–655.

[9] K. Gerstandt, K.V. Peinemann, S.E. Skilhagen, T. Thorsen, T. Holt, Membrane processes in

energy supply for an osmotic power plant, Desalination 224 (2008) 64–70.

[10] T. Thorsen, T. Holt, The potential for power production from salinity gradients by

pressure retarded osmosis, J. Membr. Sci. 335 (2009) 103–110.

[11] H. Sakai, T. Ueyama, M. Irie, K. Matsuyama, A. Tanioka, K. Saito, A. Kumano, Energy

recovery by PRO in sea water desalination plant, Desalination 389 (2016) 52–57.

16
[12] A. Achilli, J.L. Prante, N.T. Hancock, E.B. Maxwell, A.E. Childress, Experimental

results from RO-PRO: A next generation system for low-energy desalination, Environ. Sci.

Technol. 48 (2014) 6437–43.

[13] J. Kim, M. Park, S.A. Snyder, J.H. Kim, Reverse osmosis (RO) and pressure retarded

osmosis (PRO) hybrid processes: model-based scenario study, Desalination 322 (2013) 121–

30.

[14] C.F. Wan, T.S. Chung, Energy recovery by pressure retarded osmosis (PRO) in SWRO-

PRO integrated processes, Appl. Energy 162 (2016) 687–698.

[15] G. Han, J. Zuo, C.F. Wan, T.S. Chung, Hybrid pressure retarded osmosis−membrane

distillation (PRO−MD) process for osmotic power and clean water generation, Environ. Sci.:

Water Res. Technol. 1 (2015) 507–515.

[16] I.L. Alsvik, M.B. Hägg, Pressure retarded osmosis and forward osmosis membranes:

materials and methods, Polymers 5 (2013) 303–327.

[17] G. Han, S. Zhang, X. Li, Progress in pressure retarded osmosis (PRO) membranes for

osmotic power generation, Prog. Polym. Sci. 51 (2015) 1–27.

[18] N.N. Bui, J.R. McCutcheon, Nanofiber supported thin-film composite membrane for

pressure-retarded osmosis, Environ. Sci. Technol. 48 (2014) 4129–36.

[19] K.L. Lee, R.W. Baker, H.K. Lonsdale, Membranes for power generation by pressure-

retarded osmosis, J. Membr. Sci. 8 (1981) 141–171.

[20] C.F. Wan, T.S. Chung, Osmotic power generation by pressure retarded osmosis using

seawater brine as the draw solution and wastewater brine as the feed, J. Membr. Sci. 479

(2015) 148–158.

17
[21] S.C. Chen, G. Amy, T.S. Chung, Membrane fouling and anti-fouling strategies using RO

retentate from a municipal water recycling plant as the feed for osmotic power generation,

Water Res. 88 (2016) 144–155.

[22] Y. Liu, B. Mi, Effects of organic macromolecular conditioning on gypsum scaling of

forward osmosis membranes, J. Membr. Sci. 450 (2014) 153–161.

[23] W.R. Thelin, E. Sivertsen, T. Holt, G. Brekke, Natural organic matter fouling in pressure

retarded osmosis, J. Membr. Sci. 438 (2013) 46–56.

[24] Q. She, Y.K.W. Wong, S. Zhao, C.Y. Tang, Organic fouling in pressure retarded osmosis:

Experiments, mechanisms and implications, J. Membr. Sci. 428 (2013) 181–189.

[25] X. Li, T. Cai, T.S. Chung, Anti-fouling behavior of hyper-branched polyglycerol grafted

polyethersulfone hollow fiber membranes for osmotic power generation, Environ. Sci.

Technol. 48 (2014) 9898–9907.

[26] K.Y. Wang, R.C. Ong, T.S. Chung, Double-skinned forward osmosis membranes for

reducing internal concentration polarization within the porous sublayer, Ind. Eng. Chem. Res.

49 (2010) 4824–4831.

[27] S. Zhang, K.Y. Wang, T.S. Chung, H. Chen, Y.C. Jean, G. Amy, Well-constructed

cellulose acetate membranes for forward osmosis: Minimized internal concentration

polarization with an ultra-thin selective layer, J. Membr. Sci. 360 (2010) 522–535.

[28] C.Y. Tang, Q. She, W.C.L. Lay, R. Wang, R. Field, A.G. Fane, Modeling double-skinned

FO membranes, Desalination 283 (2011) 178–186.

[29] J. Su, T.S. Chung, B.J. Helmer, J.S. de Wit, Enhanced double-skinned FO membranes

with inner dense layer for wastewater treatment and macromolecule recycle using Sucrose as

draw solute, J. Membr. Sci. 396 (2012) 92–100.

18
[30] W. Fang, R. Wang, S. Chou, L. Setiawan, A.G. Fane, Composite forward osmosis hollow

fiber membranes: Integration of RO- and NF-like selective layers to enhance membrane

properties of anti-scaling and anti-internal concentration polarization, J. Membr. Sci. 394–

395 (2012) 140–150.

[31] M. Hu, S. Zheng, B. Mi, Organic fouling of graphene oxide membranes and its

implications for membrane fouling control in engineered osmosis, Environ. Sci. Technol. 50

(2016) 685–93.

[32] S. Qi, C.Q. Qiu, Y. Zhao, C.Y. Tang, Double-skinned forward osmosis membranes based

on layer-by-layer assembly-FO performance and fouling behavior, J. Membr. Sci. 405–406

(2012) 20–29.

[33] P.H.H. Duong, T.S. Chung, W, Shawn, I. Lana, Highly permeable double-skinned

forward osmosis membranes for anti-fouling in the emulsified oil-water separation process,

Environ. Sci. Technol. 48 (2014) 4537–4545.

[34] L. Setiawan, R. Wang, K. Li, A.G. Fane, Fabrication and characterization of forward

osmosis hollow fiber membranes with antifouling NF-like selective layer, J. Membr. Sci.

394–395 (2012) 80–88.

[35] E. Sivertsen, T. Holt, W. Thelin, G. Brekke, Pressure retarded osmosis efficiency for

different hollow fibre membrane module flow configurations, Desalination 312 (2013) 107–

123.

[36] G. Han, P. Wang, T.S. Chung, Highly robust thin-film composite pressure retarded

osmosis (PRO) hollow fiber membranes with high power densities for renewable salinity-

gradient energy generation, Environ. Sci. Technol. 47 (2013) 8070–8077.

19
[37] Q. She, D. Hou, J. Liu, K.H. Tan, C.Y. Tang, Effect of feed spacer induced membrane

deformation on the performance of pressure retarded osmosis (PRO): implications for PRO

process operation, J. Membr. Sci. 445 (2013) 170-182.

[38] G. Han, T.S. Chung, Robust and high performance pressure retarded osmosis hollow

fiber membranes for osmotic power generation, AIChE J. 60 (2014) 1107–1119.

[39] Z.L. Cheng, X. Li, Y.D. Liu, T.S. Chung, Robust outer-selective thin-film composite

polyethersulfone hollow fiber membranes with low reverse salt flux for renewable salinity-

gradient energy generation, J. Membr. Sci. 506 (2016) 119–129.

[40] S. Singh, K.C. Khulbe, T. Matsuura, P. Ramamurthy, Membrane characterization by

solute transport and atomic force microscopy, J. Membr. Sci. 142 (1998) 111–127.

[41] B. Van der Bruggen, C. Vandecasteele, Modelling of the retention of uncharged

molecules with nanofiltration, Water Res. 36 (2002) 1360–1368.

[42] Z. Z. Zhou, J.Y. Lee, T.S. Chung, Thin film composite forward-osmosis membranes with

enhanced internal osmotic pressure for internal concentration polarization reduction, Chem.

Eng. J. 249 (2014) 236–245.

[43] A.K. Ghosh, E.M.V. Hoek, Impacts of support membrane structure and chemistry on

polyamide–polysulfone interfacial composite membranes, J. Membr. Sci. 336 (2009) 140–

148.

[44] S. Zhang, P. Sukitpaneenit, T.S. Chung, Design of robust hollow fiber membranes with

high power density for osmotic energy production, Chem. Eng. J. 241 (2014) 457–465.

[45] N.N. Bui, J.R. McCutcheon, Hydrophilic nanofibers as new supports for thin film

composite membranes for engineered osmosis, Environ. Sci. Technol. 47 (2013) 1761–1769.

20
[46] G. Han, J.L. Zhou, C.F. Wan, T.S. Yang, T.S. Chung, Investigations of fouling behaviors,

antifouling and cleaning strategies for pressure retarded osmosis (PRO) membrane using

seawater desalination brine and wastewater, Water Res. 103 (2016) 264–275.

[47] D.I. Kim, J. Kim, H.K. Shon, S. Hong, Pressure retarded osmosis (PRO) for integrating

seawater desalination and wastewater reclamation: Energy consumption and fouling, J.

Membr. Sci. 483 (2015) 34–41.

[48] S. Zhao, L. Zou, D. Mulcahy, Effects of membrane orientation on process performance in

forward osmosis applications, J. Membr. Sci. 382 (2011) 308–315.

[49] M. Zhang, D. Hou, Q. She, C.Y. Tang, Gypsum scaling in pressure retarded osmosis:

experiments, mechanisms and implications, Water Res. 48 (2014) 387–395.

[50] B. Mi, M. Elimelech, Chemical and physical aspects of organic fouling of forward

osmosis membranes, J. Membr. Sci. 320 (2008) 292–302.

Table 1. Membrane mean pore diameter, porosity, molecular weight cut-off (MWCO), PWP and
dimension of the PES hollow fiber substrate
Mean pore Geometric
Membrane PWP
diameter, dp standard Porosity MWCO -2 -1 -1
OD/ID
code variation, σp (%) (kDa) (L m h bar ) (μm/μm)
(nm)
PES hollow
11.6 1.9 74.4 183.2 130 1125/525
fiber substrate

Table 2. Specifications of the dTFC-PES hollow fiber membrane

21
Membrane Water Salt Rejection Structural Burst

code permeability, A permeability, B to NaCl, R parameter, S pressure


-2 -1 -1 −2 −1 -6
(L m h bar ) (L m h ) (%) (×10 m) (bar)

dTFC-PES 1.5±0.2 0.02 94.2±2.5 996 >20

Deionized water and 1000 ppm NaCl solution were used as the feed to measure water
permeability A and salt rejection R of the dTFC-PES membrane at 1 bar, respectively.

Highlights

· Double-skin polyamide TFC-PES (dTFC-PES) hollow fiber membrane was successfully


developed.
· Outstanding effectiveness in mitigation of ICP and fouling in FO and PRO.
· Sustainable FO and PRO performances are achieved even using real wastewater as the
feed.
· W slightly drops to 90.8% of the initial value after a 12-h test in PRO.
· Physically flushing by either freshwater or commercial membrane cleaner can efficiently
restore the water flux.

22
Fig. 1. FESEM images of the PES hollow fiber substrate: (a) inner surface, (b) outer surface, (c) overall
cross-section, and (d) enlarged partial cross-section.
Double-skin polyamide TFC-PES
hollow fiber membrane

Fig. 2. FESEM images of the double-skin thin-film composite polyethersulfone (dTFC-PES) hollow fiber membrane: (a)
overall cross-section of the membrane, (b) surface of the inner polyamide selective skin, (c) cross-section of the inner
polyamide skin, (d) surface of the outer polyamide antifouling layer, and (d) cross-section of the outer polyamide
antifouling layer.
4.0

3.5 0.7
3.0

2.5

2.0

1.5

1.0

0.5

Probability density function (nm-1)


0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6 1.8 2.0
Pore diameter, dp (nm)

Fig. 3. Pore size distribution of the modified PES hollow fiber substrate by interfacially polymerizing
a polyamide layer on the outer surface.
(a) (b) 15
100

12
80

9
60

40 6

Normalized Jw/Jw0 (%)


3

Power density W (W/m2)


20
Deionized water Deionized water
WWBr ΔP=0 bar WWBr ΔP=15 bar
0 0
0 200 400 600 800 1000 0 100 200 300 400 500 600 700 800
Permeate volume (mL) t (min)

Fig. 4. (a) Normalized water flux, Jw/Jw0, of the dTFC-PES hollow fiber membrane as a function of permeate
volume in FO under the PRO mode. (The initial water flux Jw0 was the averaged flux within the first 10 min. The
initial volume of the feed solution was 1 L); (b) variations of membrane power density, W, as a function
operation time at ΔP=15 bar.
4000 nm

700 nm

Fig. 5. Outer surface and cross-section morphology of (a-d) fouled and (e-h) cleaned dTFC-PES hollow fiber membranes.
Membrane cleaning was performed by flushing the fouled outer surface with freshwater for 6 h.
O 1s

Na KL1
dTFC-PES Ca 2s

Fouled dTFC-PES N 1s

Ca 2p

C 1s

Cl 2p
O KL1
Si 2s
O KL2
P 2p

Si 2p

1200 1000 800 600 400 200 0


Binding Energy (eV)

Fig. 6. Outer surface XPS spectra of the pristine and fouled dTFC-PES membranes.
(a) Wastewater feed

PES porous substrate


Js Jw
Polyamide selective skin
Hydration antifouling layer

Draw solution

(b) Wastewater feed

Polyamide sealing layer

Jw
Polyamide selective skin
Hydration antifouling layer

Cations Anions Foulants Draw solution

Fig. 7. Schematic of fouling phenomena: (a) the conventional TFC membrane and (b) the newly developed double-skin
TFC membrane in FO (under the PRO mode) and PRO processes.
(b) O 1s
(a) Na KL1
Ca 2s
Fouled dTFC-PES Fouled dTFC-PES
N 1s
100 Cleaned dTFC-PES Rinsed dTFC-PES
Ca 2p

C 1s
80

Cl 2p
60 O KL1
Si 2s
O KL2
P 2p
40
Si 2p

Jw,DI/Jw0,DI (%)
20

0
Deionized Genesol 704 1200 1000 800 600 400 200 0
water (pH=11)
Binding Energy (eV)

Fig. 8. (a) Normalized initial water fluxes of the fouled and cleaned dTFC-PES membranes and (b) outer surface
XPS spectra of the fouled and cleaned membranes. The membranes were cleaned by flushing with freshwater for 6
h or commercial membrane cleaner Genesol 704 (pH=11.0) for 0.5 h on the fouled surface. Jw,DI and Jw0,DI were
obtained under the PRO mode at ΔP=0 bar using deionized water as the feed.
100 P=0 bar

80

60

40

Normalized Jw/Jw0 (%)


20
1st
2nd
0
0 200 400 600 800 1000
Permeate volume (mL)

Fig. 9. Normalized water flux, Jw/Jw0, of the dTFC-PES hollow fiber membrane as a function of permeate
volume in FO under the PRO mode using WWBr as the feed. After the first test cycle, the membrane was
cleaned by flushing with commercial membrane cleaner Genesol 704 (pH=11.0) on the fouled surface for 0.5
h. (The initial water flux Jw0 was the averaged flux within the first 10 min. The initial volume of the feed
solution was 1 L).

You might also like