Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

RSC Advances

View Article Online


PAPER View Journal | View Issue

Preparation and characterization of the tolerance


Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

to acid/alkaline and anti-oil-fouling of regenerated


Cite this: RSC Adv., 2016, 6, 114750
cellulose membranes for oil–water separation
Wanfa Zhang, Xu Chen, Jiefeng Pan, Congjie Gao and Jiangnan Shen*

Novel regenerated cellulose (RC) membranes were fabricated using an immersion precipitation phase
inversion method from five non-derivative solvents (NaOH/urea/H2O, NaOH/thiourea/H2O, LiOH/urea/
H2O, NaOH/urea/thiourea/H2O and LiOH/urea/thiourea/H2O) for oil–water separation. The prepared RC
membranes exhibited a similar asymmetric and reticulated pore structure, with a pore diameter ranging
from 12.77 to 17.09 nm, which is similar to typical ultrafiltration membranes. The biggest pore diameter
was obtained using a NaOH/urea/H2O solvent and the best mechanical properties were obtained using
a LiOH/urea/H2O solvent, which indicated that by using different mixed solvents, different pore sizes and
mechanical properties can be obtained for fabricated cellulose membranes. Measurements of the flux
and the rejection using edible oil/water in this study showed that our prepared membranes possessed
excellent oil retention ratios of up to 98%; meanwhile the water flux recovery ratios were around 90%
Received 24th July 2016
Accepted 29th November 2016
through simple water flushing. Besides, after immersion in strong acid/alkaline solution for a week, the
pure water flux (J), dimensions and weight showed no obvious change, and there was no change in
DOI: 10.1039/c6ra18766h
filtration performance at room temperature with a pH from 1 to 14, indicating the advantage of high
www.rsc.org/advances tolerance to acid/alkaline.

excellent fouling resistance properties due to their high


1. Introduction hydrophilicity.10
With the rapid growth of industry and the wide application of Cellulose, a renewable chemical resource, is considered as
oil, large amounts of oily wastewater are becoming an inevitable one of the most abundant organic polymers on the earth. It has
challenge to the world; not only bringing many toxic attracted much attention due to its outstanding properties,
compounds to water, but also threatening the ecological envi- such as excellent mechanical strength, biodegradability, high
ronment, and even human beings.1,2 In this context, different hydrophilicity, and thermal and chemical stability. Therefore, it
oil–water separation technologies have emerged, such as is a desirable material source in a future biomaterial-based and
chemical emulsication, pH adjustment, gravity settling, sustainable economy.11–13 However, cellulose applications are
centrifugal settling, heat treatment, membrane ltration, etc., limited because it is very difficult to dissolve in some conven-
each with advantages and disadvantages.3–5 Among these, tional solvents and form into the desired shape.14 The reasons
membrane technology has proved to be one of the best strate- for this are the strong intra-molecular and inter-molecular
gies, possessing higher oil removal efficiency, lower energy hydrogen bonds between cellulose chains.13 Traditionally,
costs, and smaller space requirements, without additives and cellulose is used to produce cellulose derivatives, such as
hence providing a reduction in environmental pollution.6 As the cellulose acetate (CA) and cellulose nitrate (CN), which damages
core component, membranes play an important role in sepa- its ability to endure severe acid, alkaline and organic solvents,
ration. Many studies have been widely developed for oil/water as well as causing serious environmental problems. Conse-
separation with different membranes.7–9 However, membrane quently, cellulose solvents have a profound effect on the utili-
fouling is a key problem affecting the economic and techno- zation of cellulose resources.
logical viability of the membrane separation process. As one of Recently, solvents that can dissolve cellulose have been
the most promising candidates, cellulose membranes exhibit established, including N-methylmorpholine-N-oxide (NMMO),
not only excellent oil/water separation selectivity, but also DMAc/LiCl and ionic liquids, however the cost is too high.15–18
Nowadays, several economical and practical mixed solvents
have been developed, including NaOH/urea/H2O,13,19–28 NaOH/
thiourea/H2O,29–33 LiOH/urea/H2O,34–40 NaOH/urea/thiourea/
Center for Membrane Separation and Water Science & Technology, Ocean College,
Zhejiang University of Technology, Hangzhou 310014, China. E-mail: shenjn@zjut.
H2O41–44 and LiOH/urea/thiourea/H2O.45,46 These non-derivative
edu.cn solvents can dissolve cellulose below 10  C, which makes

114750 | RSC Adv., 2016, 6, 114750–114757 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances

them easy to operate. The dissolution of cellulose at low into a 5 wt% H2SO4 coagulation bath to coagulate and regen-
temperature is a fast dynamic self-assembly process between erate for 5 min at 20  C, and the solutions were coded as RC-NU,
solvent molecules and cellulose macromolecules.47 The disso- RC-NT, RC-LU, RC-LTU and RC-NTU, respectively. The resulting
lution mechanism of cellulose in solvents at low temperature RC membranes were washed with a large amount of running
consists of two steps: (1) NaOH (or LiOH) “hydrates” are water and deionized water until they reached pH ¼ 7, and they
attracted to cellulose chains through the formation of new were nally immersed into deionized water for the
hydrogen-bonded networks, and are relatively stable, existing at characterization.
low temperatures; and (2) the urea (or thiourea, or urea and
thiourea) hydrates act as a shell, surrounding the NaOH (or
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

LiOH) hydrogen-bonded cellulose to form an inclusion complex


(IC) with a sheath-like structure, and dissolve the cellulose.32,40,47 2.3 Characterization
The mixed solvents used in the whole process are 2.3.1 X-ray diffraction (XRD). RC membranes were dried at
environmentally-friendly; meanwhile, the dissolution process is 70  C for 6 h in a vacuum oven and were then ground into
very fast, and a transparent solution is obtained in only a few powder. Samples were tested using a PANalytical (Netherlands)
minutes.48 Regenerated cellulose solutions have been success- X'Pert PRO, for X-ray diffraction (XRD) analysis. The X-ray
fully prepared using the mixed solvent mentioned above, and source is CuKa radiation (l ¼ 0.154056 nm), at a voltage of 40
parts of them have been used for membrane preparation.20,35,49 kV, and a current of 40 mA, using as the detector an X'Celerator
In this work, a series of regenerated cellulose membranes super detector, over a 2q range from 5 to 50 degrees. The crys-
were prepared via a solution casting method using mixed tallinity Xc (%) of the cotton pulp and RC membranes was
solvents. The membrane performance for oil–water separation calculated using the following relationship:
and its tolerance to acidic and alkaline (pH 1 and 14) conditions
Sc
was also studied. In addition, the morphology, and physical and Xc % ¼  100% (1)
Sc þ Sa
mechanical properties of the regenerated membranes were
investigated in detail. We hope to provide some useful infor- where Sc and Sa are the areas of the crystal and amorphous
mation on the structure of cellulose membranes from these diffraction peaks of the samples, respectively.
cellulose solvents at low temperature, and to promote the 2.3.2 X-ray photo-electron spectroscopy (XPS). The N
further application of RC membranes. content of the residual urea (or thiourea, or urea and thiourea)
in the cellulose membranes was measured with an elemental
analyser (XPS, Kratos AXIS Ultra DLD, Japan).
2. Experimental 2.3.3 Scanning electron microscopy (SEM). The
2.1 Materials morphology of the RC membranes was investigated using
The cellulose (cotton pulp) was supplied by Shandong Silver a Hitachi S-4700(II) SEM operating at an acceleration voltage of
Hawk Chemical Fiber Co., Ltd (Gaomi, China) and the average 5 kV. The prepared RC membranes were frozen in liquid
degree of polymerization (DP) was 600. It was dried at 70  C for nitrogen, and immediately snapped and sputtered with gold,
24 h before use. Bovine serum albumin (BSA) with a MW of and then the morphology of the RC membranes was observed
67 000 Da was purchased from Aladdin. Edible oil for the and photographed.
preparation of an oil–water emulsion was supplied by Shanghai 2.3.4 Mechanical properties. The tensile strength (sb) and
Yi Hai Kerry Oil Co., Ltd (Pudong, China). All other chemical elongation at break (3) of the membranes in both dry and wet
reagents of analytical grade were purchased from commercial states were measured using a CTM2050 (Xin Qiang instrument
sources in China. manufacturing Shanghai Co., Ltd) instrument at a strain rate of
5 mm min1 at room temperature. A gauge length of 30 mm was
used. The specimens were cut into strips of 100 mm  10 mm.
2.2 Preparation of regenerated cellulose membranes The wet membranes were measured immediately aer soaking
Five different cellulose directed solvents were employed to in water for 30 min. The experiment was repeated at least three
dissolve the above mentioned cotton pulp. Firstly, aqueous times and average values were taken.
solutions of 7 wt% NaOH/12 wt% urea,20–22,24 9.5 wt% NaOH/4.5 2.3.5 Filtration experiment. Water permeability (J)
wt% thiourea,31,32 5 wt% LiOH/12 wt% urea,34,35 4.6 wt% LiOH/ measurements of the wet membranes were conducted in
10 wt% urea/4.5 wt% thiourea,41–44 and 8 wt% NaOH/8 wt% a cross-ow manner and were carried out using miniature
urea/6.5 wt% thiourea45,46 were precooled to 12  C in a freezer. ultraltration equipment at room temperature. All the
Then, a certain amount of cellulose was added immediately into membrane samples were pre-compacted under a pressure of
a solution and it was stirred vigorously for 5 min at ambient 0.2 MPa for 30 min in order to get steady ltration and were
temperature, and a transparent solution was obtained con- tested at 0.1 MPa pressure. Aer that, an oil-in-water emulsion
taining 4 wt% cellulose. Aer that, the transparent cellulose containing edible oil at 800 mg L1 in deionized water was used
solution was centrifuged at 8000 rpm for 10 min at 10  C to to test the separation performance of the membrane. The oil-in-
remove the slightly insoluble residuals and to degas the solu- water emulsion was generated through mixing edible oil with
tion. The cellulose solution was immediately cast on a glass water at over 3000 rpm for 30 min, which obtained a stable and
plate to a given thickness of 360 mm and was then immersed reproducible emulsion. The pH of the oily wastewater was

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 114750–114757 | 114751
View Article Online

RSC Advances Paper

adjusted using NaOH and H2SO4. The J value was calculated J2


FFR% ¼  100% (7)
using the following equation J1
V
J¼ (2)
St The ux decline ratio (FDR) was calculated using eqn (8):
and the permeability coefficient (P) was calculated using the  
J2
following equation FDR% ¼ 1   100% (8)
J1
J
P¼ (3) where J1 is the pure water ux of the pristine membrane and J2
DP
is the pure water ux of the cleaned membrane aer oil-in-water
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

where J (L h1 m2) is the permeate ux, t (h) is the time, V (L) is emulsion rejection experiments.
the permeated volume of water (or solution) for a given time, S
(m2) is the effective surface area of the membrane, P (L h1 m2
MPa1) is the permeability coefficient, and DP (MPa) is the 3. Results and discussion
operating pressure.
3.1 Structure and morphology of the RC membranes
Solute concentration in the feed and permeated solutions
was determined using ultraviolet visible light spectroscopy (UV- The crystal structure of a cellulose membrane was usually
Vis) methods. The rejection of BSA or oil-in-water emulsion was closely related to its formation conditions and can reect the
dened through: mechanical performance of the membrane. Fig. 1 shows the
  crystalline structure of the RC membranes and cotton pulp. The
Cp
R¼ 1  100% (4) RC membranes displayed three typical diffraction peaks at 2q ¼
Cf 0), (100) and
12.1 , 19.9 and 21.7 , which correspond to the (11
26,30
where Cp and Cf are the concentrations of the permeate and the (200) planes of cellulose II, respectively. These peaks were
feed, respectively. different from the three main peaks of the cotton pulp, whose
The porosity (Pr) of the RC membranes was calculated using special characteristic peaks at 2q ¼ 14.9 , 16.6 , 22.8 were
eqn (5):50 assigned to the (11 0), (110), and (200) planes of cellulose I,
 14
respectively. The XRD results indicate clearly that cellulose I
ðWw  Wd Þ rH2 O
Pr ¼   100% (5) (cotton pulp) is thoroughly converted to cellulose II (regen-
ðWw  Wd Þ rH2 O þ Ww =rc
erated cellulose) through the dissolution and regeneration
where Pr is the porosity of the wet membrane, Ww, and Wd are processes. Furthermore, the crystallinity of all the samples was
the wet and the dry membrane weights respectively, rH2O calculated based on the area under the curves of the diffraction
(0.998 g cm3) is the density of water, and rc (1.528 g cm3) is pattern (see eqn (1)) and this is shown in Table 1. The Xc (%)
the density of cellulose. values of all the RC membranes (32.73–45.32%) are lower than
The mean pore radius, rf (m), was evaluated by means of the that of cotton pulp (67.44%), suggesting that the crystal struc-
water-ow-rate method, using eqn (6):51 ture of the cotton pulp has been destroyed and reconstructed.
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi Fig. 2 shows the whole XPS spectra of ve prepared membranes.
8hH2 OJL It is worth noting that there is no nitrogen element (no N peaks
rf ¼ (6)
Pr DP are present in Fig. 2) existing on the surface of the membranes,
which further suggests that residual urea (or thiourea, or urea
where rf (m) is the mean pore radius of the RC membrane, hH2O
and thiourea) in the membrane has been completely eluted
(N s m2) is the viscosity of water, J (m3 m2 s1) is the pure
through water ushing.
water ux, L (m) is the thickness of the wet membrane, Pr is the
porosity of the wet membrane, and DP (Pa) is the trans-
membrane pressure.
The RC membrane was soaked in sulfuric acid aqueous
solution with a pH of 1 and sodium hydroxide aqueous solution
with a pH of 14, respectively, for a week at room temperature to
evaluate the tolerance to acid and alkaline conditions. The size
and distribution of emulsied oil droplets were measured using
a Malvern Zetasizer 3000 HAS particle size analyzer.
The antifouling properties of the RC membranes were eval-
uated using water–oil–water three stage ltration processes.
Aer oil-in-water emulsion rejection experiments, the
membranes were ushed with deionized water for 30 min, to
clean them. Then the pure water ux of the cleaned membranes
was measured at 0.1 MPa pressure. The ux recovery ratio (FRR)
was calculated using eqn (7):52
Fig. 1 XRD patterns of the RC membranes and cotton pulp.

114752 | RSC Adv., 2016, 6, 114750–114757 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances

Table 1 Porosity (Pr), average pore size (2rf), crystallinity (Xc), and wet Fig. 4 shows SEM images of the RC membranes. All the
membrane thickness (L) for various regenerated cellulose membranes fabricated membranes display a similar asymmetric structure
and cotton pulp
and interconnected networks. The results indicate that the
Sample 2rf (nm) Pr% L (mm) Xc% porous structure of the membrane is well preserved through the
freeze-drying method. The porous structure of the membrane
Cotton pulp — — — 67.44 was ascribed to phase separation of the cellulose solution
RC-NU 15.55 91.46  0.49 0.171 45.32 during regeneration in 5% H2SO4 aqueous solution. When the
RC-NT 13.97 91.11  0.82 0.169 42.60
RC-LU 12.77 90.64  0.20 0.161 32.73
glass plate for casting the cellulose solution was immersed into
RC-LTU 17.09 91.84  0.42 0.183 37.21 the coagulation bath, diffusion between non-solvent and
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

RC-NTU 15.21 90.39  0.06 0.180 33.17 solvent occurred. With a large amount of non-solvent diffusion
in the casting solution, polymer-rich and solvent-rich regions
were formed, and the solvent-rich regions contribute to the
formation of the pores.21 Simultaneously, from the cross-section
of the RC-LTU membrane, we can see the pore size near the free

Fig. 2 XPS whole spectra of the RC membranes.

In this study, ve kinds of regenerated cellulose (RC)


membranes were successfully fabricated using different mixed
solvents, but the same structure was exhibited from a macro-
scopic perspective. To minimize changes to the structure of the
membrane in the wet state, freeze-drying techniques were
adopted to obtain the dry membrane (50  C, 2 Pa for 12 h).
Here, as an example, the RC-LTU membrane (diameter, 47 mm)
in the wet state and dry state (freeze-drying) is shown in Fig. 3.
Clearly, the dry or wet state of the RC membrane has
a tremendous inuence on the transparency; it exhibits
complete transparency under wet conditions, but seems like
a piece of white paper under dry condition. That is attributed to
different llers in the pores of the membrane under wet or dry
conditions. In wet conditions, the porous membrane (as shown
in Fig. 3a) is lled with water, which has a much lower refractive
index compared with air, so the membrane exhibits a trans-
parent state.

Fig. 4 SEM images of the RC membranes: (a–c) RC-NU, (d–f) RC-NT,


(g–i) RC-LU, (j–l) RC-LTU, and (m–o) RC-NTU; (a, d, g, j and m) the
free surface (in contact with the coagulant), and (b, e, h, k and n) the
back surface (in contact with glass); (c, f, i, l and o) a cross-section; and
Fig. 3 Digital photos of an RC membrane: (a) wet state; and (b) dry (p–r) a cross-section of the RC-LTU membrane: (p), near the free
state. surface; (q), the middle section; and (r), near the back surface.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 114750–114757 | 114753
View Article Online

RSC Advances Paper

surface (in contact with the coagulant) is larger than that near stronger cellulose solubility than the NaOH/urea aqueous
the back surface (in contact with glass) from the SEM images, solvent and NaOH/thiourea aqueous solvent.40 The results
and the pore size was gradually reduced from the free surface to indicate that cellulose membranes with different mechanical
the back surface. This could be explained in that coagulant properties can be obtained by using different mixed solvents.
diffusion was much faster on the free surface than on the back The tensile strengths of all the membranes in the dry or wet
surface. As listed in Table 1, the pore diameter (2rf) ranges from state were relatively low, due to the low degree of polymerization
12.77 to 17.09 nm for different RC membranes, measured using of cellulose (DP 600) and low concentration of casting solution
the water-ow-rate method, and the porosity (Pr) was very high, (4 wt%) under the effective operating pressure (below 0.2 MPa),
between 90% and 92%, due to the excellent hydrophilicity of as shown in part 3.3.
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

cellulose. The results indicate that cellulose membranes with


different pore sizes can be obtained through using different
3.3 Permeation and rejection performance
mixed solvents.
In order to evaluate the permeation and rejection performances
of our prepared RC membranes, BSA rejection was taken as an
3.2 Mechanical properties
example. The experimental results, including water ux and
The mechanical properties of membranes, including tensile BSA rejection, are depicted in Fig. 6a. The pure water ux values
strength (sb) and elongation at break (3), are important of the RC membranes are between 11.6 and 18.5 L h1 m2.
parameters in relation to their application in oil/water separa- Interestingly, the pure water ux decreases with a decrease in
tion processes. Fig. 5 displays sb and 3 for the RC membranes the crystallinity of the membranes, which is in contrast with
prepared from the different mixed solvents. It is clear that the sb Table 1, for the RC membranes (RC-NU, RC-NT, and RC-LU)
and 3 values for both the dry and wet prepared RC membranes fabricated in a three component solvent system. That is to
decrease with a decrease in the membrane crystallinity (shown say, for a certain concentration of cellulose, the stronger the
in Table 1). The RC-LU membrane prepared from a 5 wt% LiOH/ dissolution system for membrane preparation, the smaller the
12 wt% urea system exhibits lower mechanical properties than pore size of the membrane structure. The Xc value for RC-LU
the others. For the ternary system (7 wt% NaOH/12 wt% urea,
9.5 wt% NaOH/4.5 wt% thiourea, 5 wt% LiOH/12 wt% urea), the
reason could be the fact that the LiOH/urea aqueous solvent has

Fig. 6 (a) Pure water flux and BSA rejection using the RC membranes;
Fig. 5 The tensile strength (sb) and elongation at break (3) of the five and (b) pure water flux of the RC-LTU membrane under different
prepared membranes under dry (a) and wet (b) conditions. operating pressures.

114754 | RSC Adv., 2016, 6, 114750–114757 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances

Table 2 RC membrane tolerance to acidic and alkaline conditions

Ja SA/SBb WA/WBc
Conditions pH (L h1 m2) (%) (%)

Untreated 7 18.5 100 100


Sulfuric acid 1 18.3 98.6 98.5
treatment
Sodium hydroxide 14 18.4 98.3 98.1
treatment
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

a
Pure water ux. b Area of the membrane aer and before treatment
under acidic or alkaline conditions, respectively. c Weight of the
membrane aer and before treatment under acidic or alkaline
conditions, respectively.

(32.73%) is relatively close to the value for RC-NTU (33.17%),


but there is an obvious gap in the pure water ux between them;
and RC-LTU possesses the highest ux of 18.5 L h1 m2 with
an Xc value of only 37.21%. The reason is that the pores formed
by the four component system are larger than in the three
component system, as conrmed in Fig. 4. The permeate ux
variation under different operating pressure conditions was
also evaluated, as shown in Fig. 6b. With an increase in the
operating pressure, rstly, the permeate ux of the RC-LTU
membrane increased gradually, and then the permeate ux
gradually stabilized aer the operating pressure increased to
0.18 MPa. It is implied that the networked pores of the RC
membrane undergo a continuous compression process, which
reduces the porosity and increases the ux resistance to some
extent with an increase in the operating pressure. Therefore, all
the RC membranes were pre-compacted at 0.2 MPa for 30 min
before testing the ux.

3.4 Tolerance to acid and alkaline conditions


Long term alkaline and acid stability, for adapting to different pH
values of operating environments, is one of the anticipated
properties for membrane application to oil/water separation. As
a representative of the RC membranes, only the RC-LTU
membrane was tested, due to similar chemical properties across
all samples. In this study, the performance losses (J, S and W) were
evaluated aer immersion in a solution with a pH value of 1 and
14, and recorded one week later at room temperature. Table 2 Fig. 7 Size distribution of emulsified oil droplets in oil-in-water
demonstrates that the RC-LTU membrane exhibits only a slight emulsions with different pH values: (a) pH ¼ 1; (b) pH ¼ 7; (c) pH ¼ 14.
degradation in performance, including pure water ux, and area
or weight aer treatment. This implies that the membrane is
stable under both alkaline and acid conditions. droplet size is as small as 0.5 mm. Meanwhile, a similar
phenomenon occurs in the emulsions with different pH values
3.5 Characterization of oil-in-water emulsions (pH ¼ 1, 7 and 14), due to the excellent resistance to acid and
alkaline conditions.
Fig. 7 indicates the size distribution of emulsied oil droplets in
oil-in water emulsions with a pH of 1, 7 or 14. Digital photos are
depicted in Fig. 8. It can be seen that the sizes of oil droplets 3.6 Oil–water separation performance and anti-oil-fouling
range from 0.5 mm to 4 mm, complying with the normal distri- properties
bution law. Aer a separation process using a membrane of RC- Anti-oil-pollution abilities are also very important for oil-in-
LTU or RC-LT, all the solutions present a change from a turbid water emulsion separation, because trace amounts of oil drop-
solution to a clear and transparent one. These excellent sepa- lets can enter into pores of membrane and result in membrane
ration results for oil and water further prove the successful fouling. The RC-LTU and RC-LU membranes were employed to
preparation of oil/water separation membranes, even if the oil evaluate the anti-oil-fouling properties of the RC membranes

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 114750–114757 | 114755
View Article Online

RSC Advances Paper

revealed the excellent anti-oil-fouling properties of the RC


membranes.
RC membranes in acid, neutral, and alkaline oil-in-water
emulsions, with pH values of 1, 7, and 14, were also employed
to evaluate the anti-oil-fouling properties (Fig. 9b). The initial
ux of the RC-LTU membrane was 18.5 L (m2 h)1, which
Fig. 8 Digital photos of the rejection results using RC membranes
declined to 16.4 L (m2 h)1, 17.1 L (m2 h)1, and 15.8 L (m2 h)1
before and after separation (F: feed emulsion, P: permeate solution): for oil-in-water emulsion ltration at a pH of 1, 7, and 14,
(a) different RC membranes; (b) the RC-LTU membrane at different respectively. The ux could recover to 16.6 L (m2 h)1, 16.8 L (m2
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

pHs. h)1, and 16.5 L (m2 h)1 following simple water ushing, with
a ux recovery ratio of 89.7% at pH 1, 91.0% at pH 7, and 89.1%
(Fig. 9a). Oil rejection ratios were all above 98%, with the RC- at pH 14. Although the RC membrane exhibited a slightly lower
LTU membrane showing 98.34%, and the RC-LU membrane ux recovery ratio at pH 1 and 14, the anti-oil-fouling properties
99.02%. Particularly, the initial ux of the RC-LTU membrane of the RC membrane were also competitive under a harsh acidic
was 18.5 L (m2 h)1 and it dropped to 17.1 L (m2 h)1 aer the or basic environment. The rejection ratios of the emulsions at
ltration of the oil-in-water emulsion, while the pure water ux different pH values were higher than 98% at a pH of 1 (98.08%),
could recover up to 16.8 L (m2 h)1 following simple water 7 (98.31%) and 14 (98.14%), respectively. This further revealed
ushing, with a ux recovery ratio of 91.0% and a ux decline the excellent anti-oil-fouling properties of the RC membranes.
ratio of 7.5%. For the RC-LU membrane, the initial ux was The above results indicate that the RC membranes have
11.6 L (m2 h)1, which declined to 10.0 L (m2 h)1 aer oil-in- potential application in the treatment of oily wastewater.
water emulsion ltration, while the ux could recover to
10.5 L (m2 h)1 following simple water ushing with a ux 4. Conclusions
recovery ratio of 90.3% and a ux decline ratio of 13.6%, which
Novel regenerated cellulose oil–water separation membranes
were developed using immersion precipitation phase inversion
strategies, with ve kinds of mixed solvent (7 wt% NaOH/12
wt% urea, 9.5 wt% NaOH/4.5 wt% thiourea, 5 wt% LiOH/12
wt% urea, 4.6 wt% LiOH/10 wt% urea/4.5 wt% thiourea, 8
wt% NaOH/8 wt% urea/6.5 wt% thiourea). Their similar struc-
tures and morphologies were conrmed using SEM and X-ray
diffraction. Pore diameters ranging from 12.77 to 17.09 nm,
measured using the water-ow-rate method, indicated that the
RC membranes belong to a typical ultraltration membrane
class. Measurements with an edible oil/water emulsion indi-
cated that the prepared membranes possessed excellent oil
retention ratios of up to 98% and a ux of up to 18.5 L h1 m2
under a pressure of 0.1 MPa, with water ux recovery ratios of
around 90% following simple water ushing. Meanwhile, our
membranes also showed excellent stability under acidic and
alkaline conditions with no obvious change in the pure water
ux (J), dimensions and weight aer immersion at different pH
values (pH ¼ 1 or 14) for a week.

Acknowledgements
The research was supported by the National High Technology
Research and Development Program 863 (No. 2015AA030502),
and the Natural Science Foundation of Zhejiang Province (No.
LY16B060013).

Notes and references


1 M. Padaki, R. S. Murali, M. S. Abdullah, N. Misdan,
A. Moslehyani, M. A. Kassim, N. Hilal and A. F. Ismail,
Fig. 9 Permeate flux, oil-in-water rejection and flux recovery ratio for
Desalination, 2015, 357, 197–207.
the RC membranes; (a) with different RC membranes, and (b) for the 2 M. Gryta, K. Karakulski and A. W. Morawski, Water Res.,
RC-LTU membrane at different pHs. 2001, 35, 3665–3669.

114756 | RSC Adv., 2016, 6, 114750–114757 This journal is © The Royal Society of Chemistry 2016
View Article Online

Paper RSC Advances

3 J. Cui, X. Zhang, H. Liu, S. Liu and K. L. Yeung, J. Membr. Sci., 27 Z. Jiang, Y. Fang, J. Xiang, Y. Ma, A. Lu, H. Kang, Y. Huang,
2008, 325, 420–426. H. Guo, R. Liu and L. Zhang, J. Phys. Chem. B, 2014, 118,
4 T. Mohmmadi, M. Kazemimoghadam and M. Saadabadi, 10250–10257.
Desalination, 2003, 157, 369–375. 28 A. Lu, Y. Liu, L. Zhang and A. Potthast, J. Phys. Chem. B, 2011,
5 S. R. H. Abadi, M. R. Sebzari, M. Hemati, F. Rekabdar and 115, 12801–12808.
T. Mohammadi, Desalination, 2011, 265, 222–228. 29 S. Liang, L. Zhang and J. Xu, J. Membr. Sci., 2007, 287, 19–28.
6 X. Zhu, W. Tu, K. H. Wee and R. Bai, J. Membr. Sci., 2014, 466, 30 D. Ruan, L. Zhang, Y. Mao, M. Zeng and X. Li, J. Membr. Sci.,
36–44. 2004, 241, 265–274.
7 A. Salahi, A. Gheshlaghi, T. Mohammadi and S. S. Madaeni, 31 S. Liang, L. Zhang, Y. Li and J. Xu, Macromol. Chem. Phys.,
Published on 30 November 2016. Downloaded by University of Newcastle on 17/12/2016 03:43:06.

Desalination, 2010, 262, 235–242. 2007, 208, 594–602.


8 A. Salahi, M. Abbasi and T. Mohammadi, Desalination, 2010, 32 A. Lue, L. Zhang and D. Ruan, Macromol. Chem. Phys., 2007,
251, 153–160. 208, 2359–2366.
9 A. Mansourizadeh and A. Javadi Azad, J. Polym. Res., 2014, 21, 33 D. Ruan, A. Lue and L. Zhang, Polymer, 2008, 49, 1027–1036.
1–9. 34 S. Liu, J. Zeng, D. Tao and L. Zhang, Cellulose, 2010, 17,
10 J. D. Ramos, C. Milano, V. Romero, S. Escalera, M. C. Alba, 1159–1169.
M. I. Vázquez and J. Benavente, J. Membr. Sci., 2010, 352, 35 S. Liu and L. Zhang, Cellulose, 2008, 16, 189–198.
153–159. 36 M. C. V. Nagel, A. Koschella, K. Voiges, P. Mischnick and
11 X. Chen, J. Chen, T. You, K. Wang and F. Xu, Carbohydr. T. Heinze, Eur. Polym. J., 2010, 46, 1726–1735.
Polym., 2015, 125, 85–91. 37 N. Isobe, U. J. Kim, S. Kimura, M. Wada and S. Kuga, J.
12 Y. Ahn, Y. Song, S. Y. Kwak and H. Kim, Carbohydr. Polym., Colloid Interface Sci., 2011, 359, 194–201.
2016, 137, 321–327. 38 J. Cai, Y. Liu and L. Zhang, J. Polym. Sci., Part B: Polym. Phys.,
13 J. Chen, Y. Guan, K. Wang, X. Zhang, F. Xu and R. Sun, 2006, 44, 3093–3101.
Carbohydr. Polym., 2015, 128, 147–153. 39 J. Cai and L. Zhang, Macromol. Biosci., 2005, 5, 539–548.
14 H. Geng, Z. Yuan, Q. Fan, X. Dai, Y. Zhao, Z. Wang and 40 A. Lue, Y. Liu, L. Zhang and A. Potthas, Polymer, 2011, 52,
M. Qin, Carbohydr. Polym., 2014, 102, 438–444. 3857–3864.
15 T. Rosenau, A. Potthast, H. Sixta and P. Kosma, Prog. Polym. 41 S. Zhang, F. Li, J. Yu and Y. Hsieh, Carbohydr. Polym., 2010,
Sci., 2003, 26, 1763–1837. 81, 668–674.
16 T. Eiji and K. Tetsuo, J. Polym. Sci., Part B: Polym. Phys., 1999, 42 H. Jin, C. Zha and L. Gu, Carbohydr. Res., 2007, 342, 851–858.
37, 451–459. 43 S. Zhang, C.-F. Fu, F. Li, J. Yu and L. Gu, Iran. Polym. J., 2009,
17 M. B. Turner, S. K. Spear, J. D. Holbrey and R. D. Rogers, 18, 767–776.
Biomacromolecules, 2004, 5, 1379–1384. 44 S. Zhang, F. Li, J. Yu and L. Gu, Fibers Polym., 2009, 10, 34–
18 H. Li, Y. Cao, J. Qin, X. Jie, T. Wang, J. Liu and Q. Yuan, J. 39.
Membr. Sci., 2006, 279, 328–335. 45 S. Zhang, F. Li and J. Yu, Chem. Fibers Int., 2009, 59, 169–171.
19 Z. Yuan, J. Zhang, A. Jiang, W. Lv, Y. Wang, H. Geng, J. Wang 46 S. Zhang, F. Li and J. Yu, Chem. Fibers Int., 2012, 62, 186–188.
and M. Qin, Carbohydr. Polym., 2015, 117, 414–421. 47 J. Cai, L. Zhang, S. Liu, Y. Liu, X. Xu, X. Chen, B. Chu, X. Guo,
20 R. Li, L. Zhang and M. Xu, Carbohydr. Polym., 2012, 87, 95– J. Xu, H. Cheng, C. Han and S. Kuga, Macromolecules, 2008,
100. 41, 9345–9351.
21 J. Cai, L. Wang and L. Zhang, Cellulose, 2007, 14, 205–215. 48 B. Zhang, J. I. Azuma and H. Uyama, RSC Adv., 2015, 5, 2900–
22 R. Li, S. Wang, A. Lu and L. Zhang, Cellulose, 2015, 22, 339– 2907.
349. 49 Y. Mao, J. Zhou, J. Cai and L. Zhang, J. Membr. Sci., 2006, 279,
23 C. Liu, G. Zhong, H. Huang and Z. Li, Cellulose, 2014, 21, 246–255.
383–394. 50 T. Yuan, J. Meng, X. Gong, Y. Zhang and M. Xu, Desalination,
24 H. Qi, C. Chang and L. Zhang, Cellulose, 2008, 15, 779–787. 2013, 328, 58–66.
25 B. Xiong, P. Zhao, K. Hu, L. Zhang and G. Cheng, Cellulose, 51 M. Ichwan and T. Son, J. Appl. Polym. Sci., 2012, 124, 1409–
2014, 21, 1183–1192. 1418.
26 H. Qi, C. Chang and L. Zhang, Green Chem., 2009, 11, 177– 52 Z. Yi, L. Zhu, Y. Xu, Y. Zhao, X. Ma and B. Zhu, J. Membr. Sci.,
184. 2010, 365, 25–33.

This journal is © The Royal Society of Chemistry 2016 RSC Adv., 2016, 6, 114750–114757 | 114757

You might also like