Вы находитесь на странице: 1из 28

Desalination 245 (2009) 321348

Polymeric nanofiltration membranes for textile dye wastewater


treatment: Preparation, performance evaluation, transport
modelling, and fouling control a review
Woei-Jye Lau, A.F. Ismail*
Advanced Membrane Technology Research Centre, Faculty of Chemical and Natural Resources Engineering,
Universiti Teknologi Malaysia, 81310 UTM, Skudai Johor, Malaysia
Tel. +60 (7) 553-5592; Fax: +60 (7) 558-1463; email: afauzi@utm.my
Received 24 October 2007; Accepted 30 December 2007

Abstract

This paper reviews the application of polymeric nanofiltration membranes (NF) in the specific waste stream in
the textile industry, which typically generates large volumes of wastewater containing complex contaminants from
its daily operation. It is necessary that as much of this waste as possible is recycled instead of being disposed of in
landfill sites. Most of the conventional technologies seem unable to provide sufficient treatment for the effluents.
Therefore, it is generally accepted that NF membranes offer solutions for the problem. Of these NF membranes, the
thin-film composite nanofiltration (TFCNF) membrane is the most widely used by researchers in their studies. The
effects of the manufacturing conditions of TFCNF are discussed to provide valuable information for those who are
going to choose NF membranes in textile wastewater treatment. The preliminary performances of commercial NF
membranes have been examined in terms of dye rejection, salt rejection, permeate flux and COD rejection. Some
of the commercial membranes achieved maximum separation of dye and salts while some achieved higher flux. This
is because of the large variability of the parameters of textile wastewater and the NF membranes chosen. Due to the
scarcity of published papers covering the transport model that is specifically suitable for textile coloured wastewater,
a brief review of transport models of NF membranes with the presence of the dyes and salts is given. Membrane
fouling mechanisms and methods to control fouling are also reviewed. Future directions in NF membrane research
are also discussed to further expand research and development related to textile wastewater treatment.
Keywords:

Polymeric nanofiltration; Textile dyeing wastewater; Interfacial polymerization technique;


Transport model; Fouling control

*Corresponding author.

0011-9164/09/$ See front matter 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2007.12.058

322

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

1. Introduction
Textile industries traditionally use a huge
amount of water, which is normally discharged
after the wastewater treatment system to decrease
the pollution load in order to meet the legislative
requirement for the discharge. With increasing
regulatory pressures and demand for cost reduction of water and chemicals, such treatment
systems have been enhanced to address these
challenges. Textile manufacturers have therefore
converted the traditional money-wasting process of pollution control to a profitable operation
through recycling the waste effluent [1]. This
operation allows for the recovery of the valuable
chemical components and water from a number
of different textile process streams. Due to inefficiency of conventional treatment systems, nanofiltration (NF) frequently becomes the chosen
treatment process.
NF has been recognized having the properties
in between ultrafiltration (UF) and reverse
osmosis (RO) and thus offers significant advantages, e.g. lower osmotic pressure difference,
higher permeate flux, higher retention of multivalent salts and molecular weight compounds
(>300), relatively low investment and low operation and maintenance costs [2]. Many researchers
have evaluated the performance of NF membranes in terms dye retention, salt rejection, permeate flux and COD retention. The effects of
different operating conditions of wastewater and
membrane properties have been systematically
studied. The results have proven that NF membranes are the suitable separation process to be
employed for the treatment of textile wastewater
and generally showed an acceptable rejection [3].
However, to maintain the efficiency of NF membranes at a reasonable operating cost, it is necessary to use a suitable pre-treatment in order to
prevent fouling and severe module damage [4].
In view of integrating the various aspects of
NF membranes for the treatment of textile
wastewater, the aim of this paper is to review and

critically discuss performance evaluation of


various technologies for the treatment of textile
effluents, at either the laboratory-scale or pilotplant-scale level, specifically discharged from the
dyeing process, in comparison with NF membranes.
The preparation of one of the well-known
thin-film composite NF membranes (TFCNF) as
well as the performance evaluation of commercially available NF membranes in terms of dye
rejection, salt rejection, flux and COD retention
under various operating conditions is the main
focus. To gain further understanding on the transport properties of dyes and salts in NF, a
discussion of the currently available transport
models is also carried out. Furthermore, a fundamental knowledge of fouling mechanisms and
suggested methods for fouling control are also
discussed in order to provide the most efficient
solutions to minimize fouling. In addition, the
future direction of NF in textile industries is also
provided in view of developing a more competitive NF membrane, particularly for textile
wastewater treatment.
2. Nature of textile effluent
In textile refining processes, substantial
amounts of water, mineral salts and reactive dyes
are used on average for every kilogram of cotton
processed. As a consequence, they generate a
large amount of wastewater which contains complex contaminants from its daily operation. The
textile effluents typically contain many types of
dyes, detergents, solvents and salts depending on
the particular textile process such as scouring,
bleaching, dyeing, printing, finishing, etc. [5].
Table 1 shows the typical characteristics of
wastewater from the effluents of the dyeing and
finishing processes that contain a variety of components of varying concentrations [6]. For each of
the parameters involved, a range of values is
given, confirming the large variability of the

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


Table 1
Typical characteristics of wastewater from a textile
dyeing process [6]
Aspect/component

Value

pH
Temperature, EC
COD, mg/L
BOD, mg/L
TSS, mg/L
Organic nitrogen, mg/L
Total phosphorus, mg/L
Total chromium, mg/L
Color, mg/L

210
3080
505000
200300
50500
1839
0.315
0.20.5
>300

wastewater from the dyeing process [7]. Besides


these components, surfactants are also used to
reduce surface tension of water during processing; however, it only contributes to a small
amount of the wastewater. Non-ionic surfactants
(alkyl phenol ethoxylates) in wastewater should
be properly treated as it can be biodegraded to
alkyl phenols which are much more toxic than the
ethoxylated [8].
There are many classes of dyes used during
the dyeing process. The method of dye application and estimated degree of fixation for different
dye-fibre combinations are described in Table 2.
Further details regarding this information are
available elsewhere [3,9,10]. Nowadays, reactive
dyes are the most widely used dyes due to the
rapid growth in the use of this kind of dyes for
cellulosic fiber and the technical and economic
limitation of the other dyes [11]. Generally
speaking, reactive dyes which have the reactive
groups enable them to react chemically with the
fiber substrate to form a covalent bond [12]. In
comparison to other classes of dyes, the degree of
fixation of reactive dyes on the fabric is still very
low where about 550% of dyes remain in textile
wastewater due to their incomplete exhaustion
and dye hydrolysis in the alkaline dye bath during
the dyeing processes (Table 2). The hydrolyzed
dyes are derived when the reactive dyes react

323

with water instead of reacting with the functional


group of textile fabrics. The loss of dyes to the
effluent, however, is dependent on the degree of
fixation of the combination of different dye and
fiber [13]. Generally, all of the dye classes present the same problem in terms of not being
environmentally friendly. Hence, it is important
to decolorize the effluents properly before discharging into the environment in order to minimize the water pollution.
On the other hand, in the dyeing process, the
inorganic salt is added in order to enhance the
dye uptake by the fabric. Monovalent salt-sodium
chloride (NaCl) is the most common inorganic
salt that has been widely used in dyeing process.
Besides NaCl, divalent salts, e.g. sodium sulphate
(Na2SO4), are the alternative salts being used
during the process. A higher concentration of salt
in the waste stream may be a main environmental
problem in some areas due to the salination of the
soil. Therefore, it must be emphasised that a
wastewater treatment system is not just a process
to cope with the environmental problem but also
a step to recover valuable rinsed water as well as
to minimize the waste volume discharged.
3. Performance evaluation of various technologies for the treatment of textile effluents
Many researchers had investigated the performance of textile effluent treatment using various technologies [14,15]. Textile wastewater
typically consists of different types of dyes, detergents, grease and oil, heavy metal, inorganic salts
and fibers in amounts depending on the processing regime [16]. The effluents are generally
characterized using parameters such as biological
oxygen demand (BOD), chemical oxygen
demand (COD), total organic carbon (TOC), pH,
color and suspended solids (SS). Nowadays,
many of the worlds textile manufacturers are
equipped with their own wastewater treatment
plant, which usually combines an aerobic

324

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

Table 2
Method of dye application and estimated degree of fixation for different dye fiber combinations [3,9,10]
Class

Characteristics

Substrates

Method of application

Fiber

Degre of fixation
(loss to effluent), %

Acid

Anionic, water
soluble

Nylon, wool,
silk

Usually from neutral to acidic


dyebaths

Polyamide 8995 (520)

Basic

Cationic, water
soluble

PAN, modified
nylon, inks,
polyester

Applied from acidic dyebaths

Acrylic

Direct

Anionic, water
soluble

Cotton, rayona,
leather, nylon

Applied from neutral or slightly


alkaline baths containing
additional electrolytes

Cellulose 7095 (530)

Disperse

Very low water


solubility

Polyester, polyamide, acetate,


plastic, acrylic

Fine aqueous dispersions often


applied by high temperature
pressure or lower temperature
carrier methods

Polyester

Reactive

Anionic, water
soluble

Cotton, wool,
silk, nylon

Reactive site on dye reacts with


Cellulose 5090 (1050)
functional group on fiber to bind
eye covalently under influence of
heat and pH (alkaline)

Sulfur

Colloidal,
insoluble

Cotton, rayona

Aromatic substrate vatted with


Cellulose 6090 (1040)
sodium sulfide and re-oxidized to
insoluble sulfur-containing
products on fiber

Vat

Colloidal,
insoluble

Cotton, rayona

Water-insoluble dyes solubilized


by reducing with sodium
hydrosulfite, then exhausted on
fiber and re-oxidized

95100 (05)

90100 (010)

Cellulose 8095 (520)

Also known as viscose.

biological process and a physicochemical process. However, most of these traditional methods
were found inadequate due to the large variability
of composition of textile wastewater.
Table 3 illustrates the efficiencies of various
treatment systems on decolorization and COD
removal which have been employed on textile
reactive dyeing effluent. According to Marmagne
and Coste [15], the coagulation and flocculation
process is not an excellent one for reactive dye
removal. The poor quality of floc resulted in
uneven settlement even after introduction of a
flocculant. This treatment method, however, was

suitable to be used in sulphur and dispensed dye


removal due to the good quality of floc formation. They also revealed their studies on color
removal of different types of dyes using an
activated carbon treatment. Results indicated that
high removal rates (>90%) could only be
achieved for acid and cationic dyes. For reactive
dyes, a moderate removal (>50%) is considered
good. As shown in Table 3, the ozonation process
shows a higher reactive dye removal compared to
the other treatments, regardless of types of reactive dyes used. It is very effective towards oxidation of dyes and removing color, which is the

325

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


Table 3
Efficiencies (in %) of various treatment systems [15]
Dye

Reactive Blue 204


Reactive Blue 209
Reactive Red 184
Reactive Blue 41
Reactive Blue 49

Coagulationflocculation

Activated
carbon

Ozonationa

UF

MF

NF

RCOD

RColor

RCOD

RColor

RCOD

RColor

RCOD RColor RCOD RColor

RCOD RColor

28.5
31.0
23.4
60.0
19.0

53.0
88.8
22.6
38.3
35.4

70.6
89.8
69.4
74.6
19.2

69.0
78.5
77.6
57.4
94.6

67.3
45.8
85.2
44.8
85.9

99.7
99.0
99.7
99.5
99.4

2.3
0.9
0.6
0.4
0.5

89.6
94.2
89.1
93.3
93.8

5.2
4.2
6.7
7.4
5.0

80.0
76.5
80.9
76.1
81.3

95.5
94.0
96.2
94.0
92.3

97.9
97.0
98.3
97.2
96.9

Ozone used in the ozonation process: Reactive Blue 204, 83.1 mg/L; Reactive Blue 209, 80.3 mg/L; Reactive Red 184,
82.2 mg/L; Reactive Blue 41, 88.7 mg/L; Reactive Blue 49, 81.4 mg/L.

main disturbing factor for water recycling in the


textile industry. This observation was similar to
the work carried out by Selcuk [17] where almost
complete removal of color absorbances (>98%)
was achieved using ozonation in short ozone
contact time. Furthermore, to increase the efficiency of the ozonation treatment system, Tzitzi
et al. [18] reported that much better results were
obtained using a combined coagulationprecipitation/ozonation treatment instead of using a
single-stage treatment system. However, frequently, this combined treatment exhibited low
rejection in the removal of COD. To increase
both color and COD removal, sometimes a higher
ozone dosage of ozone and coagulant is required
to achieve the objective. However, it may increase the cost of operation or even form toxic
by-products from biodegradable substances during the ozonation process [19].
Other literature has revealed that the pH level
of the raw wastewater was also responsible for
the efficiency of ozone for complete decolorization. The complete decoloration time required,
however, varies significantly with varying pH
levels, irrespective of the content of salt in the
effluent [20]. As compared to the ozonation process, it was found that only approximately 50
60% color and 60% COD removal were achieved
using 1000 mgL!1 and 1500 mgL!1 of ferrous

sulphate (FeSO4.7H2O) and aluminum sulphate


(Al2(SO4)3.18H2O) in a single coagulation treatment, respectively [17]. Besides the low rejection
of color and COD, single coagulation treatment
was also found not to be a suitable treatment on
textile wastewater due to high chemical sludge
production.
Traditionally, integrated treatment processes
such as activated sludge and chemical coagulation had been widely used to deal with textile
wastewater [8,21]. The integrated treatment processes were intended to treat textile wastewater to
a level that meets the stringent regulations
required by governments. At present, numerous
laboratory-scale experiments have documented
the feasibility of various integrated treatment
technologies on recovery of rinsed water and
chemicals from textile effluent. The performance
of these integrated treatment processes is summarized in Table 4 and they may offer considerable practical information to the textile industry.
Mostly, the integrated treatment technologies are
promising, but they all still suffer limitations
which require further research and development.
Membranemembrane hybrid (MMH) treatments were also proposed and investigated
(Table 4). Compared to the studies on conventional treatment processes, the amount of information in the literature on MMH processes is

326

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

Table 4
Summary of applications of combined treatment systems on textile effluent
Treatment processes

First
stage

Second
stage

Remarks

Physical/membrane
treatment (2007)

Coagulation

UF

Achieved substantial colloidal particle removal (>97%) of


turbidity removal) regardless of type and dosage of coagulants
used, but degree of membrane fouling was highly dependent on
type of coagulants used. Study has proven that inorganic
coagulants were more efficient to reduce fouling compared to
polymeric coagulants [22]

Membrane
treatment (2006)

UF

NF

Authors claimed that UF was an appropriate pre-treatment of a


NR/RO process for textile wastewater reuse. To deal with the
wastewater with high variability values of COD and conductivity,
they observed flux decline was significant at the lowest cross
flow velocity studied due to the solid deposition onto the
membrane surface [5]

Physical/membrane
treatment (2005)

Coagulation/
flocculation

NF

Study reported that the quality of permeate after coagulation/


flocculation did not match the requirement of reuse on the site.
However, this method could act as pretreatment of NF to limit
membrane fouling. By using this integrated approach, highquality permeate could be obtained [7].

Chemical/membrane
treatment (2005)

Electrochemical
oxidation

Membr.

Study indicated the feasibility of combined processes for


treatment of textile wastewater. Membrane prior to
electrochemical oxidation process showed promising results in
terms of COD, turbidity and color removal (RCOD = 89.2%,
Rturbidity = 98.3%; Rcolor = 91.1%;) compared to electrochemical
oxidation prior to membrane process (RCOD = 86.2%, Rturbidity =
95.1%, Rcolor = 85.2%). This is due to lower color concentration
remaining in wastewater after the electrochemical oxidation
process [23]

Chemical/biological
treatment (2003)

Ozonation

Aerobic

Use of ozonation as pretreatment was able to increase the


bioavailability of the dye before it was treated with the aerobic
process. To achieve higher color (99.8%) and DOC (85%)
removal, higher doses of ozone were required. This would make it
less economically favorable [24]

Physical/membrane
treatment

Sand
filtration
and MF

NF

Sand filtration and MF in a pilot plant were fundamental in


reduction of suspended solids (100%) and turbidity (78%). To
completely remove COD, conductivity and color, NF was
responsible for removal [4]

Physical/chemical
treatment (1997)

Coagulation
and electrochemical
oxidation

Ion
exchange

Water produced from this integrated treatment was reported good


to reduce color, turbidity and COD; however, efficiency of
treatments was significantly different with varying reaction times
of H2O2 and current of electrochemical treatment [25]

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

327

Treatment processes

First
stage

Second
stage

Remarks

Physical/chemical/
biological treatment
(1996)

Coagulation
and electrochemical
oxidation

Activated
sludge

Continuous treatment showed that promising quality of permeate


water could be achieved as well as 24% cost savings over
conventional methods. However, oeprating variables (e.g.,
wastewater flow rate, applied current, aeration time in activated
sludge, egc.) should be taken into account in order to optimize
treatment performance [26]

Physical/chemical
treatment (1994)

Coagulation

Ozonation

Ozonation of wastewater after coagulation treatment exhibited


more efficient color and COD reduction compared to coagulation
of wastewater after ozonation treatment under the same
conditions of wastewater. It was due to further 90% and 2025%
of reduction of color and COD, respectively, could be achieved
by using ozonation after the coagulation process [18]

very limited, mainly because research on MMH


processes has only recently started. However, as
the amount of research and industrial applications
increases, MMH is expected to attract more attention for treating the textile wastewater in the near
future. Processes that combine membrane with
conventional treatment process/membrane process have been widely applied to achieve lower
capital cost and higher productivity. However,
fouling is often the main problem of membrane
system for complex textile manufacturing wastewater. Dyes are the components which mainly
contribute to colloidal fouling layer onto the
membrane surface. To investigate the fouling
mechanism and fouling control techniques on NF,
Section 7 of this review discusses the state-of-theart on the topic in details.
For economic and environmental reasons, it is
necessary that as much of this waste as possible
is recycled instead of being disposed of in landfill
sites. Due to the recent technological innovations
in membrane technology, the cost of membrane
systems has decreased and has led to an increase
in the use of membrane systems for wastewater
treatment processes. Though cost analysis is a
paramount exercise to undertake for textile industry process, estimating the cost of installing a
treatment system is very difficult. This is because
of variation in the raw water characteristics; the

efficiency of the process; the technological


innovations; the systems capacity; the permeate
characteristics and etc., which make the system
cost evaluation vary significantly [27]. Apart
from this, the price of purchasing various components as well as capital and operational costs
are different in each country. Further, the reports
on full-scale application of membrane systems in
the textile industry are apparently lacking. There
are only a limited number of cost analysis studies
looking into the reuse and recovery of textile
effluent using integrated membrane systems on
pilot-plant scale. This, however, could be an indication of the economical feasibility of the implementation on full industrial scale in the following
days.
In a study by Ciardelli et al. [28], it was
reported that about US $1.27/m3 treated wastewater would be a reasonable cost for the implementation of membrane techniques for treatment
of dyehouse effluents for reuse in Italy. But it is
expected to increase in the future. In India, the
total expenses incurred for water treatment and
recovery using RO/NF membrane are about
US $1.80/m3 of the effluent [29]. The cost of
recovery may be too high in some countries.
They, however, reported that the cost was still
lower than the cost of water purchased due to
non-availability of good quality water for dyeing

328

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

processes in Tirupur, India. Babursah and coworkers [27] reported that the cost of recovering
wastewater within the textile industry using a
membrane recovery system was US $0.55/m3
based on the current market conditions in Istanbul, Turkey. Details of economic analysis based
on the membrane technologies can be found
elsewhere [30,31]. With this limited information,
there is a need for comprehensive studies to
assess the economic feasibility of using NF
membrane technology for producing purified
water from wastewater, particularly in the textile
industry.

4. Preparation and characterization of polymeric nanofiltration composite membranes


There are a number of commercially available
NF membranes in the current market, which are
mostly monopolized by the niche international
companies. Among the most widely used is the
TFCNF membrane. Its excellent permeability
and selectivity over asymmetric NF membranes
offered competitive improvement of this kind of
membrane [32]. The currently available TFCNF
membranes are mainly prepared by forming a
very thin polyamide (PA) active layer on the
porous support layer which is mainly prepared
from polysulfone (PSf) or polyethersulfone (PES)
[3335]. The substrate membrane is commonly
prepared through a drywet phase inversion technique while the top active layer is formed via the
interfacial polymerization (IP) technique.
The combined techniques offer significant
advantages as either thin skin layer or porous
substrate layer can be optimized independently.
The support layer membrane can be optimized to
enhance strength and compression resistance
while the top skin layer can be optimized to
enhance desired solvent flux and solute rejection.
Through such optimization process, TFCNF
membranes generally exhibit higher salt rejection
over asymmetric NF membrane due to the ultra-

thin skin layer (300400 nm) formed onto the


porous support membrane [33]. Although TFC
NF is manufactured based on the IP method, its
performance in terms of rejection and flux profile
is different. The performance very much depends
on the support membrane employed, concentration of reactant, reaction time, organic solution
used and others, which are not fully explored.
At present, a number of studies on the preparation and characterization of TFCNF membranes via the IP technique have been reported.
Interfacial polymerization reaction occurs between two very reactive monomer (or one prepolymer) e.g., amine-type and acid chloride, at
the interface of two immiscible solvents [36]. In
this section, studies on the effects of manufacturing conditions on the TFCNF membrane have
been reviewed. In addition, it is very important to
provide valuable information for those who are
going to choose NF membranes in textile wastewater treatment.
Song et al. [32] introduced a thin active layer
of PA on polysulfone (PSf)/sulfonated polysulfone (SPSf) alloy substrates via IP using three
different types of PA: p-phenylenediamine
(PPD), m-phenylenediamine (MPD) and piperazine (PIP). The PSf/SPSf substrates prepared by
the drywet phase inversion method were
immersed into an aqueous solution before dipping
into an organic solution. Polyamide, surfactant
and phase transfer catalysts were dissolved into
distilled water to form the first solution while a
certain amount of trimesoyl chloride (TMC) was
dissolved into hexane to form the organic solution. The results indicated interactions of ionic
bonds of an interpenetrating layer between the
PAs active layer and the substrates using Fourier
transfer infrared (FTIR) and attenuated total
reflection infrared (ATIR). The thermal properties of different membranes were also investigated using differential scanning calorimetry
(DSC) and thermalgravimetry (TGA) under
nitrogen flow at a heating ramp of 10 K/min. The
thermal analysis results confirmed that an inter-

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

penetrating layer was formed between the active


layer and the support membrane based on the
investigation of the chemical composition and the
thermal property of these membranes. The strong
interaction will not cause the active layer to be
detached from support layer under harsh conditions of wastewater from the dyeing process. It
is because the chemicals in the wastewater are
able to swell the support layer and then cause the
TFCNF membrane in an undesirable condition.
In this case, they proposed the use of SPSf into
PSf porous substrates where SPSf could also act
to further improve the hydrophilicity of the
membrane and provide higher permeability instead of using PSf as the substrate.
Oh et al. [37] used the interfacial polymerization of PIP with TMC on the surfaces of
microporous polyacrylonitrile (PAN) supports to
form a strong interaction between the active and
support layers. Interestingly, they observed that
the functional groups of CN in PAN could be
modified to be COOH groups through a simple
treatment with NaOH solution at ambient temperature. The ionic bond between these two
layers is shown in Fig. 1. In addition, they also
studied the influence of modified PAN concentration on the membrane surface roughness using
atomic force microscopy (AFM). The surface
roughness increased significantly with increasing
the modified PAN concentration from 10 wt% to
20 wt%. Therefore, to remove dyes effectively
from water solutions, membrane surface roughness is an important factor to be considered.
Membrane fouling by dyes would reduce the
water flux, thus resulting in the membrane being
less economical.
In a study on the influence of monomer
compositions and organic solutions, Jegal et al.
[38] found that PA composite membranes prepared by IP of piperazine/m-phenylene diamine
(8/2 w/w) and TMC on the microporous PSf
membrane, with hexane as organic solution, were
able to increase the flux to 2.5 m3/ (m2.day) at
200 psi as the composition ratio was changed

329

Fig. 1. Ionic bond formation between the PIP of PA


active layer and COOH on the PAN support [37].

from 7/3 w/w. In addition, with changing hexane


solution into benzene/hexane mixture solutions,
interesting results were achieved. They attributed
this to benzene that was a good cosolvent to control the permeation properties of the membrane.
With the addition of 40 vol% of benzene into
hexane, a PA composite membrane could provide
promising results both in flux and solute
rejection.
Mohammad et al. [39] used Bisphenol A
(BPA) as the top active layer material on the
porous substrate made from a mixture of PSf and
polyvinylpyrrolidone (PVP) using the IP technique. They observed that an increase in the
reaction or concentration of BPA resulted in
decreasing of water permeability. This condition
was attributed to the increase of active layer
thickness. Though the thickness was increased
with increasing the reaction time, it was found
that the membrane pore size was independent of
the reaction time. However, based on AFM
results, the pore size differs considerably with
increasing the monomer concentration.
Detailed studies on the characteristics of top
active layer have been conducted by Ooi [40] and
Ahmad and Ooi [41]. Ahmad and Ooi [41]
observed that by immersing the PSf support
membrane in two different solutions containing
(a) PIP, 3,5-diaminebenzoic acid and distilled
water and (b) TMC and n-hexane, respectively, a
strong interaction of these materials would occur
where the structure of organic chemistry for the
reaction is shown in Fig. 2. They also indicated

330

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

Fig. 2. Reaction scheme of trimesoyl chloride with piperazine and 3,5 diaminobenzoic acid [41].

that strong interactions of these materials would


improve the efficiency of separation process. In
addition to the effects of top layer materials on
the TFCNF performance, different kinds of PAs
used as the top skin layer membrane would also
have the influence on separation performance.
Verisimmo et al. [42] used PAs such as PIP, N,Ndiaminopiperazine (DAP), 1,4-bis(3-aminopropyl)-piperazine (DAPP) and N-(2-aminoethyl)-piperazine (EAP) to react with TMC separately during composite membrane manufacture.
Among these membranes, it was reported that
PIPTMC exhibited higher water permeability
and rejection of monovalent and divalent salts
than those of other membranes. This may be the
reason for an increase in the use of PIP for
commercial products.
To further improve the separation performance
of TFCNF membranes via the IP technique,
Chen et al. [43] proposed the use of dimethyl
formamide (DMF) as a swelling agent in the
aqueous solution. It was found that 20 vol% of
DMF in solution was the optimum composition
since higher concentrations of swelling agents
added would not improve the thin layer formation. They observed that permeation rate of
3 L/(m2.h) and NaCl rejection rate of 94% were
achieved at optimum composition when a salt
solution of 2000 ppm was fed at 13.6 bar. This
represents an interesting finding as the rejection
of monovalent salt could be achieved as high as
divalent salt.

To date, membrane processes such as MF, UF


and RO have often been used as the wastewater
treatment. Due to certain technical reasons, NF
membranes have grown significantly as the membrane separation process during the last decade.
However, if NF can be applied successfully in the
textile industry, the TFCNF membrane is the
way to meet the requirements for the applications
with achieving both high permeability and high
rejection of inorganic salt, e.g., NaCl. The performance of TFCNF membranes has been widely
studied on textile wastewater removal; however,
in most cases, it was just carried out at laboratory
or pilot-plant scale. Therefore, to be more practical, extensive efforts are still needed to further
enhance TFCNF membrane performance.
5. Performance evaluation of NF membranes
for specific textile effluent: Effects of process
conditions
In the literature, there are a number of studies
reported on the effects of different operating conditions of textile effluents on NF performance.
The laboratory or pilot studies indicated the high
potential of using NF for reuse of water and
chemicals from textile effluents. The following is
a summary of studies based on the performance
of commercial NF under different operating conditions. Table 5 summarizes the application of the
commercially available polymeric NF membranes
used for textile effluent treatment. It was found

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

that some of the commercial membranes achieved


maximum separation of dye and salts while others
achieved higher flux.
5.1. Dye retention of nanofiltration
In a study of color and COD retention by NF
at pilot-plant scale, Lopes et al. [44] reported that
NF membranes such as NF 45 and DK 1073
exhibited good performance in terms of dye
retention. The maximum dye rejection was up to
99.2% and 99.8% respectively, with an initial dye
concentration of 450500 mg/L. Meanwhile, the
performance of MPS 31 was also investigated and
gave results of dye retention which varied from
90.197.3%. However, on average, the percentage of color rejection of MPS 31 was slightly
higher than NF 45 and DK 1073 (Table 5). This
may be due to its smaller molecular weight cut
off (MWCO). On the other hand, Sungpet and coworkers [52] attributed the dye rejection to the
secondary layer formed by retained dye on the
membrane surface. It was because the MPF 36
(MWCO 1000), having larger MWCO than MPF
34 (MWCO 200), showed higher dye removal in
the presence of a reactive dye and sodium
chloride. Thus, they found that secondary layers
formed by dye and the Donnan effect may be
responsible for dye removal instead of membrane
MWCO. Fouling layer occurred resulting from
the absorption of dye onto the membrane, resulting in an increase of dye rejection.
Tang and Chen [46] studied dye retention
using the TFC-SR2 membrane. They found that
with increasing dye concentration of Reactive
Black 5 gradually from 92 ppm to 1583 ppm, the
dye rejection remained constant at a feed pressure
of 5 bar. This indicates that dye rejection is independent of dye concentration. The results were
also supported by Akbari et al. [3] and Van der
Bruggen et al. [48]. Akbari et al. [3] reported that
dye rejection only slightly decreased with an
increasing concentration from 2000 ppm to
6000 ppm at feed pressure of 10 bar. In this case,

331

they concluded that dye molecules could perform


a good mass transfer throughout the membrane
and avoid build-up of dye concentration polarization on the membrane surface. Nevertheless, dye
molecules were able to induce color on the
membrane surface, which resulted in a fouling
problem [49].
Apart from the effect of concentration of dye
itself on dye removal, Koyuncu [53] conducted a
study to investigate the effect of salt concentration on dye rejection using the DS5 DK membrane. They reported that lower color removal
was observed with increasing NaCl concentration. Similar results were also reported elsewhere
[46]. By increasing salt concentration, the Donnan effect becomes less effective on the negatively charged membrane. This would promote
the penetration of dye molecules through the
membrane and further decrease dye retention.
However, in the work of Jiraratananon et al. [54],
the unchanged dye rejection in three different NF
membranes (ES20, NTR-729HF and LES90) in
the presence of NaCl salt indicated that retention
of a reactive red dye (Benefix) was mainly dominated by the steric effect rather than the Donnan
effect. This is reasonable as the pore radius of
these NF membranes is typically smaller than the
effective hydrodynamic radius of the dye.
The efficiency of color removal was also
dependent of cross flow velocity. The DS5 DK
membrane was further tested with different cross
flow velocities (1.11 m/s, 0.41 m/s and 0.11 m/s)
to investigate its influence on the efficiency of
color removal [53]. Color removal for cross flow
velocities of 1.11 m/s and 0.41 m/s was better
than that of 0.11 m/s due to the decrease in concentration polarization on the membrane surface.
However, no significant color rejection was
observed when increasing salt concentration to
the feed sample. Cross flow velocity, therefore,
has not played an important role in color removal
due to the presence of dye agglomeration at high
NaCl concentrations.

332

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

Table 5
Summary of the applications of commercially available polymeric NF membranes on textile effluent
Membrane
(manufacturer)

Configuration
(polymer material)

MWCO
(Da)

Process conditions

Evaluation

MPS 31
(Weizmann)

Spiral wound
(NA)a

NAa

Experiments conducted at dye


Pave = 66.25 L/(m2.h)
concentrations varying between 400
Rdye, ave = 94.9%
500 mg/L at 60EC and operating pressure of
25 bars. NaCl (10 g/L), CaCl2 (10 g/L) and
Na2SO4 (15 g/L) were added to the solution
[44]

NF45
Spiral wound
(Dow/Film Tec) (PA)b

200

Same as above.

Pave = 39.2 L/(m2.h)


Rdye, ave = 92%

DK 1073
(Osmonics)

Spiral wound
(PA)b

300

Same as above.

Pave = 60.25 L/(m2.h)


Rdye, ave = 94.5%

ATF 50
(Adv. Membr.
Tech.)

Spiral wound
(TFCc of PIPd on
PSf e)

340

Two kinds of industrial wastewater were


analyzed with (a) COD = 14,200 mg/L at
pH 10.2; (b) COD = 5430 mg/L at pH 5.5.
Experiments were carried out at
transmembrane pressure of 0.21.1 MPa
and temperatures of 2540EC [45]

RCOD = 95% for


pH 10.2
wastewater
RCOD = 80.9% for
pH 5.5 wastewater

TFCSR2
(Fluid System)

Flat sheet
(TFCc of PSf e)

200
400

921583 ppm of Reactive Black 5 and


1080 g/L of NaCl were used to synthesize
wastewater. Solutions were filtered under
cross flow velocity of 35 L/min and
operating pressures of 100500 KPa [46]

Pave = 45.05 L/(m2.h)


Rdye, ave = 97.71%

DK 2540
(Osmonics)

Spiral wound
(NA)a

NAa

Industrial wastewater with COD 1576 g/L, Pave = ~60 L/(m2.h)


color >500 Hz and conductivity of
Rsalt = 6080%
3.5 S/cm were used during experiments at
transmembrane pressure of 20 bars and cross
flow velocity of 1.66 m/s [47]

NTR 7450
(NittoDenko)

Flat sheet
(sPES)f

600800

Synthetic dyebath solution containing


Reactive Orange 16 (RO16) or Reactive
Blue 2 (RB2) (15 g/l), Na2SO4 (56 g/L),
surfactant-EDTA (0.2 g/L), Na2SO3 (1 g/L)
and NaOH (2.5 g/L) were used in experiments at operating pressures of 060 bar
and cross flow velocity of 00.75 m/s [48]

UTC 20
(Toray Ind.)

Flat sheet
(PA)b

180

Dye retentions for RO16 and RB2 were


Rdye, RO16 = ~99%
studied with cumulative addition component Rdye, RB2 = >99.3%
of 1 g/L dye, followed by 2.5 g/L of NaOH,
1 g/L of Na2SO3, 0.2 g/L of EDTA, 11 g/L
of Na2SO4 and 19 g/L of Na2SO4 [48]

P = 64 L/(m2.h)
Rdye = 92.1%
Rsalt = 87.3% at
operating pressure
of 20 bar

333

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


Membrane
(manufacturer)

MWCO
(Da)

Process conditions

NF 70
Flat sheet
(Dow/Film Tec) (PA)b

250

Experiments were carried out using two


PSA = 33 L/(m2.h)
types of exhausted dye baths from the wool PMC = 32 L/(m2.h)
dyeing process: (1) acid dye bath (SA), (2)
metal complete dye bath (MC).
Transmembrane pressure in experiments was
10 bar [49]

DL 4040F
(Osmonics)

Spiral wound
(NA)a

150300

Wastewater from a dyeing process and


finishing plant was treated using a threestage treatment system: sand filtration, UF
and NF. Mean value of the parameters was
pH 7.8, COD 142 mg/L, TSS 12 mg/L and
conductivity 3950 S/cm. UF and NF
modules worked at 0.4 and 9 bar,
respectively [50]

Desal 5 DK
(Osmonics)

Flat sheet
(TFCcPAb)

150300

Synthetic dye solution was prepared at a dye Pave = 41.1 L/(m2.h)


concentration of 1 g/L without adding
Rdye, ave = 100% for
auxiliary compounds. Experiments were
Direct Red 80 at
carried out at pressure of 10 bar, temperature Reynolds no. of
25EC and pH 6 with and without stirring
4100
[51]

a
f

Configuration
(polymer material)

b
Not available.
Polyamide.
Sulfonated polyethersulfone.

d
Thin-film composite.
Piperazineamide.
Pave = average permeability.
R = rejection.

Different observations were reported by


Chakraborty et al. [55] on dye retention using an
organic NF membrane with MWCO 400. They
attributed the decrease in dye retention after a
certain period of study to the build-up of concentration polarization of solute particles over the
membrane surface, thus enhancing the solute permeation by convection through the membrane.
Besides these effects, more studies on color
removal have been conducted using NF [51,56
58]. The research reports that the efficiency of
color removal also depended on a number of
other factors such as wastewater characteristics,
molecular weights of the dyes used, hydraulic
conditions, volume reduction factor, temperature,
pH, pressure, etc.
5.2. Salt rejection of nanofiltration
Practically, sodium chloride and sodium sul-

Evaluation

RCOD = >93%
RTSS = >60%
Rconductivity = 40.5%

Polysulfone.

phate are used as the exhausting and retarding


agents during the dyeing processes. The amount
of salts required depends on production requirements. To determine the salt transport across a
membrane, Eq. (1) is normally used.

Q s K s ( C ) A /

(1)

where Qs is salt flow through membrane, Ks the


membrane permeability coefficient for salt, C
the salt concentration difference across the membrane, A the membrane surface area and the
membrane thickness. Salt transport through the
membrane is proportional to the salt concentration difference but independent of the applied
pressure [59]. Referring to Eq. (1), increase in
applied pressure will not affect the value of the
salt flow.
Tang and Chen [46] reported that a decrease
of salt rejection occurred with increasing salt

334

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

concentration. Inorganic salt (NaCl) can be ionized completely into Na+ and Cl! in acid, alkali
or pure water. When the salt concentration
increases, so does the concentration of Na+. Based
on the principle of the Donnan equilibrium,
repulsive force from the negatively charged
membrane decreases with increasing electrolyte
concentration. Lower repulsive force means that
more Cl! anions are allowed to pass through the
membrane and thus salt rejection is reduced.
Moreover, higher salt concentration could lead to
a build-up of concentration polarization on the
membrane surface resulting in lower flux and
separation. Other researchers attributed the
decrease in NaCl rejection with increasing NaCl
concentration to the shield effect [60]. The undesirable effect of concentration polarization,
however, can be minimized by maintaining a high
flow rate of the liquid phase along the membrane
surface and by applying turbulence promoters
(spacers) between the membranes [61].
Previous studies demonstrated that adoption of
a NF hollow fiber membrane (HA 3110, Toyobo,
Japan) in a submerged MBR was feasible because
it could provide extra-clean permeate for reuse
[62]. The rejection rates of monovalent and
divalent ion by this NF membrane varying from
40% to 60% and from 70% to 90%, respectively,
during the initial 80 days of filtration process.
The lower rejection rates of monovalent salt compared to divalent salts were also reported previously [51,60]. Choi and co-workers [62] found
that salt rejections tended to decrease gradually
after 80 days, which was probably due to the
increase in pore size and decrease in the surface
charge of membranes, which deteriorated the
membrane properties.
A neutral surface membrane typically shows
a lower salt rejection as compared to a charged
membrane for a given pore size. The mechanism
of salt rejection is primarily based on the steric
effect in neutral surface membrane. The Donnan
exclusion, however, plays an important role in
retaining salt in negatively charged membranes.

The Donnan effect becomes less effective with


increasing salt concentration in the feed due to
the lower repulsive force. Jiraratananon et al. [54]
elucidated that a higher concentration of Cl! ions
would contribute to an increase in the Donnan
equilibrium of Cl! ions in the membrane,
resulting in higher ionic flux through the membrane. Consequently, a lower salt rejection is
obtained.
Vrijenhoek and Waypa [63] used the NF-45
membrane to investigate the performance of NF
membranes under single and multiple salt
solutions. In the case of uncharged membranes,
the observed order of single salt rejection is CaCl2
> NaSO4 > NaCl, which generally is dominated
by steric exclusion. Rejection of divalent ions
increased with increasing feed concentration and
flux, while rejection of monovalent ions
decreased with increasing feed concentration.
Based on molecular weight of the ions, though
the size of Ca2+ is smaller than SO42!, the rejection
of Ca2+ is slightly higher than SO42!. This is due to
cations that are able to accumulate more water
molecules around the ions, thus resulting in a
larger hydrated radius than anions. For the separation of multiple salt solutions (NaCl/Na2SO4)
by NF-45, Vrijenhoek and Waypa [63] found that
size exclusion is the dominant mechanism for ion
retention where the observed order of rejection is
SO42! > Na+ > Cl!, which is in reverse order of the
ionic diffusion coefficients (Table 6). Theoretically, the diffusion of ion in the liquid can be
determined based on Eq. (2) [64]:

Di

kT
n a

4#n#6

(2)

where k is the Boltzmanns constant, a is the


radius of solute, is the solution viscosity and n
is the StokesEinstein coefficient. However, for
a NaCl/CaCl2 mixture, they reported that ionic
diffusivity is responsible for the ion retention
instead of the size exclusion effect based on the
observed rejection of Ca2+ > Cl! > Na+ [63].

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


Table 6
Ions, ion diffusivities, ion atomic or molecular weights
and hydrated radii for salt solutions [63]
Order of rejection
(highest to lowest)

Dbulk, I
(10!9 m2/s)

AW or
MW

rH2O,
nm

NaCl/Na2SO4 salt
mixture:
SO42!
Na+
Cl!

(lowest to
highest)
1.06
1.66
2.03

96.06
22.99
35.45

NA
0.36
0.33

NaCl/CaCl2 salt
mixture:
Ca2+
Cl!
Na+

(highest to
lowest)
0.92
2.03
1.33

40.08
35.45
22.00

0.41
0.33
0.36

On the contrary, the results obtained by


Peeters et al. [65] illustrated that salt rejection in
neutral NF membrane were affected both by size
exclu-sion and Donnan exclusion effects. The
order of rejection for various salts in neutral NF
is NaSO4 > CaCl2 > NaCl. Anionic NF
membranes which have positive groups attached
to a polymer back-bone are able to repel cations,
particularly divalent cations such as Ca2+, and
attract anions, particularly divalent anions such as
SO42!. The result is an order in salt rejection such
as CaCl2 > NaCl > Na2SO4. The cationic NF
membrane with fixed negative charges preferentially rejects SO42! but permeates Ca2+ and
results in an order of salt rejection NaSO4 > NaCl
> CaCl2. Du and Zhao [34] also observed that salt
rejection in a positively charged NF membrane is
dependent not only on pore size of the membrane
but also on the static electric action between the
ion in solution and membrane surface. This
observation of salt rejection is similar to those by
Peeter et al. [65], i.e. the order of the rejection
was MgCl2 > MgSO4 > NaCl > Na2SO4.
5.3. Permeate flux of nanofiltration
Water reclamation is a key subject in the
textile industry. When the level of solute

335

retention is met, the permeate flux becomes a


fundamental factor in the process optimization. A
study by Akbari et al. [66] showed that
wastewater pH variation from 6 to 10.3 did not
affect dye retention significantly. However, the
permeate flux was affected by the type of
membrane used. Sungpet et al. [58] proved that
lower acid concentration in the industrial effluent
could lead to higher flux. To increase permeate
flux, Bowen and Mohammad [67] claimed that
the characteristics of a membrane play a main
role for the improvement. A significant improvement in flux could be obtained using looser NF
membranes with higher effective charge density
compared to the tight membranes with typical
charge density, Xd.
Another aspect that significantly influences
the permeate flux is operating temperature. High
temperature water requires lower operating pressure to achieve a desired flux when compared to
operating at low temperature water. Lower solute
rejection occurs at higher temperatures due to a
greater solute penetration through the membrane.
For a NF membrane, an approximation of permeate flux can be made by:

Q p Q p (25o C) 1.03(T 25)

(3)

where Qp is the permeate flow at temperature T,


Qp(25EC) the permeate flow at 25oC and T is the
water temperature, oC. To obtain a desired quantity of permeate flux, controlling the temperature
of the feed solution is one of the critical parameters which should be considered. All polymeric membranes have their own maximum
operating temperature. In general, these membranes are considered sustainable in most of the
separation processes which require not very high
operating temperature [6,68]. However, good
control of feed temperature is still required in
order to minimize the change in the physical and
chemical properties of the membrane.
Studies by Chen et al. [45] clearly concluded

336

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

that permeate flux increases proportionally with


an increase in the transmembrane pressure drop
and with an increase in the operating temperature
due to the decrease in water viscosity. Viscosity
usually decreases significantly when temperature
is increased whereas viscosity increases with
increasing pressure. However, the effect is generally insignificant at a pressure less than 4 MPa
(~39.5 bar) [69]. Though the influence of feed
pressure on viscosity is insignificant, it has great
influence on permeate flux. Bes-Pia et al. [70]
studied the relationship between the permeate
flux and feed pressure of NF-90 and DK-5 and
found that by increasing the feed pressure from
10 to 20 bar, the permeate flux changed
significantly. Nevertheless, no influence of feed
flow rate (in the studied range of 200400 L/h)
on permeate flux was noticed.
On the other hand, with permeate recycled to
the system, the NF 70 membrane could provide a
stable flux even after 610 h of filtration [49].
However, when no permeate was recycled, no
stable value was achieved. This is due to the
osmotic pressure and adsorption of organic compounds on the membrane material. Thus, in order
to maintain stable permeate flux over the period
of study, a pre-treatment system was applied.
Marcucci et al. [4] proposed the use of MF membranes as the pre-treatment. The DL 4040F
membrane resulted in a very constant permeate
flux even after 530 h of operation. In addition,
Van der Bruggen et al. [71] proposed an alternative option by using biological treatment as a
pre-treatment before NF to lower the flux decline
as well as to provide higher permeate quality in
long-term operation. Without the pre-treatment,
significant flux decline was reported for a singlestage treatment of either the NF 70 or UTC-20.
Flux decline is suspected to be stronger over long
operation hours than that indicated in the study.
However, more intensive studies are needed to
confirm the applicability of the combined treatment processes as NF of the biological effluent
on a real scale might suffer from biofouling.

As mentioned before, the permeate flux may


be influenced by the feed solution velocity. In
such a case, the Reynolds number is widely used
to express solution turbulence:

Re

Dv

(4)

where D, v, and are the diameter of tube,


average velocity of liquid, density of liquid and
viscosity of liquid, respectively. The transition of
the fluid from laminar to turbulent flow was
found to be dependent on the Reynolds number
[72]. Tang and Chen [46] observed that in the NF
experiments, an increasing flow rate at the
applied pressure of 200 kPa did not increase the
flux since the Reynolds numbers for the cross
flow velocity of 3 L/min and 5 L/min were 3000
and 6000, respectively in the turbulent region.
However, for the Reynolds number which falls in
either the transition region or laminar region, it
may have an effect on the permeate flux since the
solid deposition happens onto the membrane
surface at low flow velocity [59]. To further
promote the turbulence flow, Auddy et al. [73]
employed thin wires as turbulent promoters in the
cross flow NF system. The flux was enhanced in
the range of 40% to 100% due to the decrease in
concentration polarization, resulting in less solute
deposition over the membrane surface.
The influence of inorganic salts on permeate
flux was also studied by several researchers. Van
der Bruggen et al. [49] have evaluated the influences of monovalent salts (NaCl) and bivalent
salts (Na2SO4, Na2CO3) on the permeate flux of
NF. It was observed that the presence of
inorganic salt in the feed solution increased the
osmotic pressure due to the increase of ionic
concentration in the solution. Thus, a higher feed
operating pressure is required to achieve the same
permeate flux. To optimize the separation performance, there are two models which are widely
employed to estimate osmotic pressure before the

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

filtration process. The osmotic pressure based on


the Vant Hoff model is defined as:

vc j

RT
M

(5)

where v is the number of ions, cj is concentration


of component j (kg/m3), R is 8.314 (J/mol.K), T is
absolute temperature (K), and M is molecular
weight (g/mol). This model, however, resulted in
an overestimation of osmotic pressure at high
concentrations of ions. Therefore, for better prediction of osmotic pressure for higher concentration solutions, the Pitzer model is recommended. Van der Bruggen et al. [71] employed
this model to evaluate the flux decline in three
different membranes at 10 bar where the osmotic
pressure can be expressed as follows:

RTM s
vm
1000vs

(6)

where Ms is the molecular weight of the solvent


(g/mol), vs the molar volume of solvent (m3/mol),
vm the number of positive ions and N is the osmotic pressure coefficient. By using the Pitzer
model, they estimated the osmotic pressure was
about 2.7 bar when the salt (Na2SO4) concentration was 7.8 g/L. Therefore, they reported that
flux decline was 27% at an operating pressure of
10 bar.
Apart from salt osmotic pressure, a study was
carried out by Gomes et al. [74] to investigate the
influence of dye osmotic pressure on permeate
flux in the NF 45. For a very low dye concentration, the osmotic pressure is increased proportionally with the concentration. However, dye
osmotic pressure was found to be concentration
independent, especially for concentrations higher
than 3 g/L. In this case, dye osmotic pressure was
not the dominant factor to affect the permeate
flux as dye would aggregate with increasing dye
concentration.

337

5.4. COD retention of nanofiltration


The COD test is normally used to measure the
oxygen equivalent of the organic material in
wastewater that can be oxidized chemically using
dichromate in an acid solution. Several
researchers reported NF membranes had merit to
minimize the COD values from the textile
effluent. Bes-Pia et al. [70] observed that the
measured COD values in permeate were lower
than 50 mg/L (initial values varying between
200400 mg/L) in all experiments using both NF90 and DK-5 membranes. With such percentage
of removal (7683%), they considered that the
reduction in textile industry wastewater was
satisfactory. According to Sojka-Ledakowicz et
al. [75], a very high reduction of COD (up to
99%) could only be achieved using RO membranes. However, with the combined treatment
system of NF and physical/chemical treatment, it
was reported that nearly 100% of COD removal
was achieved [76].
On the contrary, combined treatment with the
NF DL4040F as the final membrane process in a
pilot-plant scale testing showed that the quality of
the effluent did not match the requirement of
water to be reused since the percentage removal
of COD, conductivity and total hardness were
93%, 40% and 75%, respectively [50]. It was
recommended that the integrated approach (sand
filtration, UF and NF) should be used for minor
processes in the textile industry such as washing
and dyeing, since they required a relatively low
quality of water.
Studies by Chen et al. [45] on desizing wastewater for the bleaching and dyeing industry in
Hong Kong showed that only a minor increase in
COD retention was achieved for an increase in
transmembrane pressure drop as well as operating
temperature. The minor increase, however, is
expected from the analysis of COD transport
model as follows:

1
RCOD

1
RsCOD

BsCOD 1
.
RsCOD J v

(7)

338

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

where RCOD, RsCOD, BsCOD and Jv are the overall


retention of COD, overall retention parameter for
COD, overall mass transfer parameter for COD
and permeate flux, respectively. Referring to
Eq. (7), with an increase in transmembrane pressure, permeate flux would be increased since
1/RCOD decreases and RCOD increases. It was a
very interesting case where higher rejection of
COD was achieved with increasing permeate
flux. However, Chakraborty et al. [55] observed
that the COD removal decreased with increasing
operating pressure since more solutes were able
to permeate through the NF membrane at a higher
pressure. Therefore, more studies are needed to
verify the relationship between the COD removal
and operating pressure.
The influence of pH on COD removal was
also investigated by Chen et al. [45]. They
reported that a higher retention of COD was
achieved at a higher pH value. It might be the
acidic environment for lower pH value that made
the hydrolyzation of starch more significant.
Thus, the COD retention (95%) at pH 10.2 wastewater was higher than at pH 5.5 wastewater
(8085%) using the ATF 50 membrane. Furthermore, NF membranes, e.g., MPS 31, NF 45
and DK 1073, have been proven to reduce the
COD values ranging from 73% to 87% with an
initial COD value of 700 mg/L [44]. The remaining COD in filtrate was probably produced by the
solutes and other oxidizable materials.
6. Transport modelling in nanofiltration
membranes
Nowadays membrane filtration processes are
widely used for industrial separations. In this
circumstance, there is an increasing need for a
model-based tool to design new membrane systems for a variety of product separations or to
optimize existing membrane installations. Although the applicability of NF membranes in the
textile industry was increased significantly, their
transport mechanism is not yet well understood

due to their unknown structure and the complexity in composition of textile effluent. To date,
there are numerous studies reported in the literature regarding various salt removals using NF.
Generally, two main approaches have been used
to model the transport of inorganic salts through
NF. The first was based on the extended Nernst
Plank model (ENP) [41,77,78] and the second
was based on the Spiegler and Kedem model (SP)
[7981]. Other than these models, the Teorem
MeyerSiever (TMS) and the DonnanSteric
pore model (DSPM) also were used to determine
salt rejection [79,80,8284].
To model dye mixture separation, an
unsteady-state mass transfer model was
developed and successfully tested for prediction
of permeate flux and permeate concentration
polarization of dye components without considering the presence of salt in an unstirred NF
batch [85]. The research was continuously conducted to investigate the effects of different
operating parameters on per-eate flux and permeate concentration of dye in an unstirred batch
and a cross-flow cell [86]. Nevertheless, in the
literature there are no studies covering the
transport model that specifically is versatile for
textile colored wastewater. The mechanism of salt
transport in NF is complicated in the presence of
organic dye and many other components, thus
making the analysis of transport much more
difficult.
Generally, the fouling layer or gel polarization
layer occurs from the absorption of dye onto the
membrane. The situation could be worsened in
the presence of other components, e.g., wax,
fibers, oil, etc. in the textile effluent. Therefore,
the currently available transport models are
insufficient to predict the performance of NF in
textile wastewater. More intensive studies are
therefore desired. However, a brief review of
transport models of NF membranes relevant to
textile dyeing effluent is made below to provide
the basic knowledge for those who are interested
in further improving transport models.

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

According to Koyuncu and Topacik [87],


apart from the two layers which were considered
in the previous studiesa concentration
boundary layer on the high pressure side of the
membrane and membrane itselfthey also took
into consideration the effects of gel polarization
or fouling on salt and dye removal. Therefore, the
average mass transport of salt, kavg, was defined
as:

1
kavg

1
1

k s k sd

(8)

where ks is the mass transport coefficient of salt in


front of the gel layer of organic ion (dye) and ksd
the mass transport coefficient of salt inside the
gel layer of organic ion. They observed that
accumulation of dye molecules on the membrane
surface resulted in a decrease of permeate flux
due to the friction loss of the gel layer. Based on
the SpieglerKedem and PerryLinder models
and film theory equations, they described that dye
concentration had a significant effect on flux
values for a fixed NaCl concentration of
26 mol/m3, 340 mol/m3, 600 mol/m3 and
1145 mol/m3. However, interesting results were
obtained with increasing salt concentration higher
than 340 mol/m3. This is because high salt concentration increases the degree of aggregation and
subsequently has a positive effect on membrane
fouling, which in turn results in lower flux.
Comparison of the model with the experiments
showed that the model is able to predict rejection
as well as flux reduction behavior for systems
containing NaCl, dye and water.
Al-Bastaki [88] investigated the efficiency of
the membrane process in removing color and salt
from a synthetic colored wastewater using a
theoretical model which is based on the solution
diffusion (SD) mass transport theory, where both
salt and dye concentration polarization effects
were included as well as the possibility of
dynamic membrane formation. Similar to

339

Koyuncu and Topacik [87], they found that


dynamic membranes formed in the presence of
dye could reduce water permeability due to an
increase in membrane resistance. In this case,
water permeability, Kw, was proposed in terms of
the combined resistance of membrane and
dynamic membrane:

Kw

RM

1
R DM

(9)

where RM is the resistance of membrane and RDM


is the resistance of a dynamic membrane in the
presence of dye. Al-Bastaki [88] described the
transport of ions with different feed solutions
containing different concentrations of dye and
salt at different operating pressures based on the
SD theory. The theoretical results generally
showed good agreement with experimental results
in terms of salt rejection, color removal as well as
permeate flux, and made it possible to determine
the permeability of dyeing wastewater.
Because the Donnan effect leads to a difference in the rejection and permeability of NF, it
has been widely utilized in transport models to
improve the dyesalt separation [89,90].
Although salt passage in NF is expected to occur
both by diffusion and convection, Levenstein et
al. [90] proposed a simple two-parameter model:
the diffusivity parameter, B0, and the power
exponent in salt permeability equations, n, by
neglecting the convection term. This model
enables characterization of salt rejection in NF of
multi-component aqueous solutions, e.g., NaCl
H2O, NaClH2Odye and NaClH2OdyePNa.
With increasing dye concentrations in solutions,
lower permeate fluxes as well as lower chloride
rejections were obtained since the Donnan effect
is expected to occur, which could enhance
chloride passage through membranes. Generally,
the model proposed in this work was found to be
applicable for predicting salt removal due to very
good agreement with the experimental results.

340

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

7. Fouling control in polymeric nanofiltration


membranes
Fouling is often a weakness of NF for
complex textile manufacturing. Dyes can produce
a colloidal fouling layer, which further introduces
an undesirable flux decline in the operation [91].
Heavy membrane fouling is expected since dyes
can be accumulated on the active layer of polyamide NF membranes by chemical bonds either
in the ionic or covalent bond, depending on pH or
the class of dye [51].
Generally, the phenomenon of membrane
fouling is inevitable, but it is reversible by using
feed pre-treatment, modifying the membrane
surface or by controlling membrane cleaning
procedures. To evaluate the fouling potential in
an early stage, a membrane fouling simulator
(MFS) was developed for evaluation of fouling
control by using different chemicals in order to
make NF membrane systems less susceptible to
fouling [92]. The evaluation was also done in
combination with liquid chromatographyorganic
carbon detection (LCOCD) and other analytical
methods to characterize fouling material in
membranes and identify species responsible for
fouling [93].
The conventional prevention of fouling by
applying a pre-treatment can still be coped with
by using appropriate cleaning procedures. The
cleaning procedures are typically conducted using
physical and chemical methods [94]. Physical
methods can be intermittent back-washing,
application of critical flux, critical TMP, intermittent suction operation, low TMP, high cross
flow velocity and hydrodynamic shear stress
scouring. On the other hand, chemical cleaning
agents can be acids (strong or weak), alkalis
(NaOH), detergents, enzymes, complexing agents
(EDTA) and disinfectants. Furthermore, chemical
cleaning agents are recommended by membrane
manufacturers since they have the ability to
recover completely the initial membrane permeability and require less energy consumption

compared to physical methods [95]. Nevertheless,


chemical treatments are relatively expensive and
may cause severe membrane damage and produce
toxic by-product waste [96].
Some researchers reported that each class of
dyes could cause membrane fouling. The difference between these dyes was the thickness or
hardness of dye cake layer accumulated on the
membrane surface [56]. Fouling makes the NF
membrane separation process less economically
favorable. Thus, they suggested that a pretreatment with chemical coagulant-alum was required
to decrease the extent of membrane fouling. In
comparison, the degree of flux decline of pretreated wastewater over the operation time
became smaller than when original wastewater
was used (Fig. 3). The results showed that
approximately 20% of improvement of flux was
observed using the pre-treatment system prior to
NF. Furthermore, the use of an ozonation process,
MF as well as UF as pre-treatment for the NF
membrane in textile effluent has also been investigated in order to minimize membrane fouling
and deterioration to meet the objective of prolonging the NF membrane life span [4,5,70].
In terms of membrane modification, Mulder
[36] reported that negatively charged membranes

Fig. 3. Effect of pre-treatment with alum on the flux


decline along the operation time when artificial dyeing
wastewater was treated with the NF PA composite
membrane [97].

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

have the potential of reducing fouling, especially


in the presence of negatively charged colloids in
the feed. On the contrary, Kim and Lee [97]
modified the NF membrane surface by coating
with a neutral polyvinyl alcohol (PVA) to reduce
surface charge and membrane surface roughness
so that the ionic bridge of the cations between the
membrane and dye could be reduced. They experienced fouling could easily have resulted from the
divalent cations in the effluent using a typically
higher surface charge of NF. The results confirmed that the PVA-coated NF membranes were
successful in increasing fouling resistance and
subsequently reducing membrane fouling.
Furthermore, the antifouling NF membrane
was introduced in the work of Asatekin et al. [98]
by coating the PVDF support membrane with
an amphiphilic graft copolymer, poly(vinylidene
fluoride)-graft-poly(oxyethylene) methacrylate,
PVDF-g-POEM. The higher fouling resistance
and higher water produced can be attributed to
both the nanoscale dimensions of the hydrophilic
channels through the coating and to the neutral
charge of POE which creates the barrier to the
adsorption. Recently, the amphiphilic comb copolymer additive, polyacrylonitrile-graft-polyethylene oxide (PAN-g-PEO) was also found to
have excellent antifouling characteristics by coating it on a PAN UF membrane [99]. The authors
attributed this to the surface segregation and local
orientation of PAN-g-PEO molecules at the membrane surface and pore walls, forming a dense
brush layer as the barrier to the adsorption.
Instead of membrane surface coating, hydrophilicity or hydrophobicity of membrane is the
factor that should be taken into account in reducing the extent of fouling. Less fouling is observed
for aqueous solutions or suspensions when the
membranes are strongly hydrophilic due to the
preferential wetting of such material by water. In
the work conducted by Boussu et al. [100], they
reported that NF 270 (27E) was less fouled compared to the large contact angle membrane,
namely NF 90 (54E) and BW30XLE (51E) since

341

a small contact angle which is corresponding to


the hydrophilic surface could reduce the tendency
of membrane fouling.
Sungpet et al. [52] elucidated that the use of
NaCl in the dyeing process could enhance the
penetration of reactive dye into membranes,
which results in NF membranes heavily colored
after the experiments. This, however, subsequently led to a flux decline in the membrane
process. Interestingly, the use of chemical cleaning with 0.2 wt% HNO3 followed by 0.5 wt%
NaOH could recover 80100% of the flux where
the chemical cleaning procedures were periodically carried out. On the other hand, Marcucci et
al. [4] used alkaline detergent (l2%) for removal
of organic material and acid detergent (12%) for
removal of inorganic material from a textile
effluent using a combined membrane treatment
system. This method can be used immediately if
the hydraulic performance is worsened. The
results showed that after the chemical washing,
the initial permeate flux of MF and NF membrane
was re-established even after 300 h of operation.
However, a delay in chemical cleaning of NF
membranes led to irreversible changes in membrane structure and eventually deteriorated the
membrane performance [101].
In the work conducted by Shu et al. [102], it
was concluded that membrane fouling caused by
dye absorption was reversible, but was highly
dependent on membrane cleaning. Similar observations were also obtained by Lopes et al. [44]
with applying chemical cleaning on commercial
NF membranes. However, no details of the
chemical solution used as well as the frequency
of chemical cleaning procedure were given.
On the other hand, Van der Bruggen et al. [49]
claimed that membrane fouling caused by the
adsorption or pore blocking of organic compounds on the membranes had a large influence
on the permeate flux due to the high concentration of organic compounds used in textile
dyeing. In the application of NF membranes for
the treatment of exhausted dye baths, 2646% of

342

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

the irreversible flux declines were reported as the


relative difference between the initial pure water
flux and pure water flux measured after the
experiment. Although absorbed organic compounds might be desorbed by rinsing water, it
was still difficult for those dyes which were
supposed to attach strongly to the membranes.
The absorption of dyes onto the membrane surface is due to the reaction between the polyamide
fibers with dye either in ionic bonding, covalent
bond or Van der Waals.
Moreover, the fouling phenomenon is also
found to be linked with the hydrodynamic conditions of the filtration system. The hydrodynamic
of cross-flow velocities (CFV) plays an important
role in influencing membrane fouling to control
the build-up of solute in NF membrane surfaces.
A study was conducted by Petrinic et al. [103] to
evaluate the membrane fouling caused by the dye
bath wastewater at variable CFV. In comparison,
the higher CFV of 0.8m/s was sufficient to keep
the concentration polarization layer small enough
compared to the CFV of 0.4 m/s and 0.6 m/s and
provided promising results in terms of the permeate flux and color removal. This is supported
by Koyuncu [53] where he experienced that flux
was increased with increasing CFV from 0.11 m/s
to 1.11 m/s, regardless of the concentration of salt
in wastewater. However, the effects of CFV were
not significant for the high NaCl concentrations
(80 g/L) due to the aggregation of dye molecules
at high NaCl concentrations. Instead of controlling operating parameters, an alternative approach
for modeling flux decline during membrane
separation processes is based on the filtration
theory [104, 105]. Elimelech and Bhattacharje
[105] developed a theoretical model which is
based on the principles of thermodynamics and
hydrodynamics for prediction of permeate flux
during steady-state cross flow membrane filtration. The results showed that the predictions of
permeate flux compare remarkably well with a
detailed numerical solution of the convective
diffusion equation coupled with the osmotic

pressure model. Moreover, the model is also


capable of predicting the point where cake
formation is initiated. The prediction is useful
since cake formation on membranes is an inevitable phenomenon in the textile industry.

8. Future direction of research and development of nanofiltration membranes for textile


wastewater treatment
NF membranes have been proven to be one of
the most important separation processes for
textile dyeing effluents treatment based on
numerous research carried out so far. However,
improvements on these membranes are still
needed in order to further enhance its performance before NF becomes a dominant commercialized wastewater treatment system in a largescale industrial plant. At present, many published
papers are still at laboratory- or pilot-scale level
and, consequently, further works would be
required in the near future.
Continuous operation of pilot plants should be
optimized and intensive investigations on the
long-term performance of NF membranes should
be carried out to provide a good indication of
how a specific component in textile wastewater
leads to membrane fouling. This could assist in
the development of suitable pretreatment systems
prior to NF in order to achieve higher permeability and minimize membrane fouling on longterm operation.
With respect to the manufacture of TFCNF
membranes using the IP technique, development
of NF membranes should be further expanded
both in the range of chemical compatibilities and
physical operating conditions (including pressure,
CFV, temperature and pH) of membrane systems
due to the high variability of textile effluent
values in order to offer a greater potential over
other membranes. In terms of energy consumption, extensive development effort would help to
find the hidden energy sources in membrane

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

processes since the textile wastewater having


operating temperatures in a range between 30
80EC can be reused in their daily operation. A
detailed economic evaluation of this NF technology on textile wastewater treat-ment plants,
including maintenance and operation costs,
should also be conducted in order to gain
confidence in NF in the competitive market. With
all of these improvements, a major breakthrough
in NF membrane technologies research will
definitely overcome the limitations and weaknesses of current technologies and contribute
greatly to the textile industry worldwide in the
near future.
9. Conclusions
It is difficult to draw a general conclusion on
the feasibility and the efficiency of NF for dyeing
effluent treatment in the textile industry due to
the large variability of textile wastewater
parameters and the quality of the permeate
required. However, based on the numerous
studies conducted so far, NF membranes have
proved applicable in dealing with textile
wastewater which is highly colored as well as
highly loaded with monovalent and/or divalent
salts.
Generally, it can be concluded that NF offers
many more advantages compared to conventional
treatment methods and the other categories of
membrane technologies. To commission the full
scale of NF membrane treatment plants in the
textile industry, a long-term performance of the
system should be carried out along with the
installation of a pretreatment system prior to NF
for the purpose of minimizing membrane fouling
and prolonging the membrane life span as well as
increasing the efficiency of the overall treatment
system. Besides, with the understanding of the
transport mechanisms in NF membrane, it will
lead toward better prediction and optimization of
separation processes in the textile industry.
Finally, research and development in this field are

343

a must in order to gain confidence in NF for


textile wastewater treatment system.
10. Symbols
a
A
cj
C
D
Di
Jv
k
kavg
ks
ksd
Ks
Kw
M
Ms
n
R
RM
RDM
Re
Qp
Qs
T
v
v
vm
vs

StokesEinstein radius, m
Membrane surface area, m2
Concentration of component j,
kg/m3
Salt concentration difference across
the membrane, kg/m3
Diameter of tube, m
Diffusivity, m2/s
Permeate flux, m3/m2.h
Boltzmanns constant
Average mass transport coefficient
of salt
Mass transport coefficient of salt
Mass transport coefficient of salt
inside the gel layer of organic ion
Membrane permeability coefficient
Water permeability of combined dynamic membrane, m2.s/kg
Molecular weight, g/mol
Molecular weight of the solvent,
g/mol
StokesEinstein coefficient
Gas constant, J/mol.K
Resistance of membrane, kg/m2.s
Resistance of dynamic membrane in
the presence of dye, kg/m2.s
Reynolds number
Permeate flow, kg/m2
Salt flow through membranes, kg/m2
Water temperature, EC
Average velocity of liquid, m/s
Number of ions
Number of positive ions
Molar volume of solvent, m3/mol

Greek

Solution viscosity, N.s/m2


Viscosity of liquid, N.s/m2

344

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

Osmotic pressure, bar


Density of liquid, kg/m3
Membrane thickness, m
Osmotic pressure coefficient

References
[1] L. Cecille and J.C. Toussaint, Future Industrial
Prospects of Membrane Processes, Elsevier Applied
Science, London and New York, 1988.
[2] N. Hilal, H. Al-Zoubi, N.A. Darwish, A.W.
Mohammad and M.A. Arabi, A comprehensive
review of nanofiltration membranes: Treatment,
pretreatment, modelling, and atomic force microscopy, Desalination, 170 (2004) 281308.
[3] A. Akbari, J.C. Remigy and P. Aptel, Treatment of
textile dye effluent using a polyamide-based nanofiltration membrane, Chem. Eng. Proc., 41 (2002)
601609.
[4] M. Marcucci, G. Ciardelli, A. Matteucci, L. Ranieri
and M. Russo, Experimental campaigns on textile
wastewater for reuse by means of different membrane processes, Desalination, 149 (2002) 137143.
[5] S. Barredo-Damas, M.I. Alcaina-Miranda, M.I.
Iborra-Clar, A. Bes-Pia, J.A. Mendoza-Roca and
A. Iborra-Clar, Study of the UF process as pretreatment of NF membranes for textile wastewater reuse,
Desalination, 200 (2006) 745747.
[6] M. Marcucci, I. Ciabatti, A. Matteucci and G.
Vernaglione, Membrane technologies applied to
textile wastewater treatment, Ann. NY Acad. Sci.,
984 (2003) 5364.
[7] C. Suksaroj, M. Heran, C. Allegre and F. Persin,
Treatment of textile plant effluent by nanofiltration
and/or reverse osmosis for water reuse, Desalination,
178 (2005) 333341.
[8] P.C. Vandevivere, R. Bianchi and W. Verstraete,
Treatment and reuse of wastewater from the textile
wet-processing industry: review of emerging technologies, J. Chem. Tech. Biotech., 72 (1998) 289
302.
[9] Kirk-Othmer, Encylopaedia of Chemical Technology, Vol. 8, Wiley, New York, 1993.
[10] J.R. Easton, The dye makers view, in: Colour in
Dyehouse Effluent, P. Cooper, ed., Society of Dyers
and Colourists, Alden Press, Oxford, 1995, pp. 9
21.

[11] B.M. Gatewood and J. Hall, Evaluation of aftertreatments for reusing reactive dyes, Textile Chemist
Colorist, 28 (1996) 3842.
[12] S.R. Srivastva, Recent Processes of Textile
Bleaching, Dyeing and Finishing, Small Business,
Delhi, 1979.
[13] C. ONeill, F.R. Hawkes, D.L. Hawkes, N.D.
Lourenco, H.M. Pinheiro and W. Delee, Review:
colour in textile effluents-sources, measurement,
discharge consents and simulation, J. Chem. Tech.
Biotech., 71 (1999) 10091018.
[14] M. Senthikumar and M. Muthukumar, Studies on the
possibility of recycling reactive dye bath effluent
after decolouration using ozone, Dyes Pigments, 72
(2007) 251255.
[15] O. Marmagne and C. Coste, Color removal from
textile plant effluents, American Dyestuff Reporter,
April 1996.
[16] A. Lopez, G. Ricco, R. Cinnarella, A.C.D. Pinto and
R. Passino, Textile wastewater reuse: ozonation of
membrane concentrated secondary effluent, Water
Sci. Tech., 40(45) (1999) 99105.
[17] H. Selcuk, Decolorization and detoxification of
textile wastewater by ozonation and coagulation
processes, Dyes Pigments, 64 (2005) 217222.
[18] M. Tzitzi, D.V. Vayenas and G. Lyberatos, Pretreatment of textile wastewater with ozone, Water Sci.
Technol., 29 (1994) 151160.
[19] F. Zhang, A. Yediler, X. Liang and A. Kettrup,
Effects of dye additives on ozonation process and
oxidation by-products: a comparative study using
hydrolyzed C.I. Reactive Red, Dyes Pigments, 60
(2004) 17.
[20] M. Muthukumar and N. Selvakumar, Studies on the
effect of inorganic salts on decolouration of acid dye
effluents by ozonation, Dyes Pigments, 62 (2004)
221228.
[21] C.F. Gurnham, Industrial Waste Control, Academic
Press, New York, 1965.
[22] K.-H. Choo, S.-J. Choia and E.-D. Hwang, Effect of
coagulant types on textile wastewater reclamation in
a combined coagulation/ultrafiltration system,
Desalination, 202 (2007) 262270.
[23] X. Chen, Z. Shen, X. Zhu, Y. Fan and W. Wang,
Advanced treatment of textile wastewater for use
using electrochemical oxidation and membrane
filtration, Water SA, 31 (2005) 127132.
[24] J.A. Libra and F. Sosath, Combination of biological

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

[25]

[26]

[27]

[28]

[29]

[30]

[31]

[32]

[33]

[34]

[35]

[36]

and chemical processes for the treatment of textile


wastewater containing reactive dyes, J. Chem.
Technol. Biotechnol., 78 (2003) 11491156.
S.H. Lin and M.L. Chen, Treatment of textile wastewater by chemical methods for reuse, Water Res., 31
(1997) 868876.
S.H. Lin and C.F. Peng, Continuous treatment of
textile wastewater by combined coagulation, electrochemical oxidation and activated sludge, Water Res.,
30 (1996) 587592.
S. Babursah, M. akmakci and C. Kinaci, Analysis
and monitoring: costing textile effluent and reuse,
Filtration Sep., June 2006.
G. Ciardelli, L. Corsi and M. Marcucci, Membrane
separation for wastewater reuse in the textile
industry, Resources, Conservation Recycling, 31
(2000) 189197.
K. Ranganathan, K. Karunagaran and D.C. Sharma,
Recycling of wastewaters of textile dyeing industries
using advanced treatment technology and cost
analysisCase studies, Resources, Conservation
Recycling, 50(3) (2007) 306318.
M. Marcucci and L. Tognotti, Reuse of wastewater
for industrial needs: the Pontedera case, Resources,
Conservation Recycling, 34 (2002) 249259.
K. Madwar and H. Tarazi, Desalination techniques
for industrial wastewater reuse, Desalination, 152
(2002) 325332.
Y.J. Song, F. Liu and B.H. Sun, Preparation, characterization and application of thin film composite
nanofiltration membranes, J. Appl. Polym. Sci., 95
(2005) 12511261.
Y.F. Zhang, C.F. Xiao, E.H. Liu, Q.Y. Du, X. Wang
and H.L. Yu, Investigations on the structures and
performances of a polypiperazine amide/polysulfone
composite membrane, Desalination, 191 (2006)
291295.
R.H. Du and J.S. Zhao, Properties of poly (N,Ndimethylaminoethyl methacrylate)/polysulfone positively charged composite nanofiltration membrane,
J. Membr. Sci., 239 (2004) 183188.
Y.J. Song, P. Sun, L.L. Henry and B.H. Sun,
Mechanisms of structure and performance controlled
thin film composite membrane formation via
interfacial polymerization process, J. Membr. Tech.,
251 (2005) 6779.
M. Mulder, Basic Princiles of Membrane Technology, Kluwer Academic, London, 1996.

345

[37] N.W. Oh, J. Jegal and K.H. Lee, Preparation and


characterization of nanofiltration composite membranes using polyacrylonitrile (PAN). II. Preparation
and characterization of polyamide composite
membranes, J. Appl. Polym. Sci., 80 (2001) 2729
2736.
[38] J. Jegal, S.G. Min and K.-H. Lee, Factors affecting
the interfacial polymerization of polyamide active
layers for the formation of polyamide composite
membranes, J. Appl. Polym. Sci., 86 (2002) 2781
2787.
[39] A.W. Mohammad, N. Hilal and M.N.A. Seman,
Interfacially polymerized nanofiltration membranes:
atomic force microscopy and salts rejection studies,
J. Appl. Polym. Sci., 96 (2005) 605612.
[40] B.S. Ooi, Development of polymeric composite
nanofiltration membrane: synthesis, characterization
and its evaluation on copper sulfate rejection, Ph.D.
Thesis, Univesiti Sains Malaysia, 2005.
[41] A.L. Ahmad and B.S. Ooi, Characterization of
composite nanofiltration membrane using twoparameters model of Extended NernstPlanck
equation, Sep. Purif. Technol., 50 (2006) 300309.
[42] S. Verissimo, K.-V. Peinemann and J. Bordado,
Influence of the diamine structure on the nanofiltration performance, surface morphology and surface charge of the composite polyamide membranes,
J. Membr. Sci., 279 (2006) 266275.
[43] S.H. Chen, D.J. Chang, R.M. Liou, C.S. Hsu and
S.S. Lin, Preparation and separation properties of
polyamide nanofiltration membrane, J. Appl. Polym.
Sci., 83 (2002) 11121118.
[44] C.N. Lopes, J.C.C. Petrus and H.G. Riella, Color and
COD retention by nanofiltration membranes,
Desalination, 172 (2005) 7783.
[45] G. Chen, X. Chai, P.L. Yue and Y. Mi, Treatment of
textile desizing wastewater by pilot scale nanofiltration membrane separation, J. Membr. Sci., 127
(1997) 9399.
[46] C. Tang and V. Chen, Nanofiltration of textile
wastewater for water reuse, Desalination, 143 (2002)
1120.
[47] A. Bes-Pia, M.I. Iborra-Clar, A. Iborra-Clar, J.A.
Mendoza-Roca, B. Cuartas-Uribe and M.I. AlcainaMiranda, Nanofiltration of textile industry wastewater using a physicochemical process as a pretreatment, Desalination, 178 (2005) 343349.
[48] B. Van der Bruggen, B. Daems, D. Wilms and

346

[49]

[50]

[51]

[52]

[53]

[54]

[55]

[56]

[57]

[58]

[59]

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


C. Vandecasteele, Mechanisms of retention and flux
decline for the nanofiltration of dye baths from the
textile industry, Sep. Purif. Technol., 2223 (2001)
519528.
B. Van der Bruggen, G. Cornelis, C. Vandecasteele
and I. Devreese, Fouling of nanofiltration and
ultrafiltration membranes applied for wastewater
regeneration in textile industry, Desalination, 175
(2005) 111119.
M. Marcucci, G. Nosenzo, G. Capannelli, I. Ciabatti,
D. Corrieri and G. Ciardelli, Treatment and reuse of
textile effluents based on new ultrafiltration and
other membrane technologies, Desalination, 138
(2001) 7582.
A. Akbari, J.C. Remigy and P. Aptel, Treatment of
textile dye effluent using a polyamide-based nanofiltration membrane, Chem. Eng. Proc., 41 (2002)
601609.
A. Sungpet, R. Jiraratananon and P. Luangsowan,
Treatment of effluents from textile-rinsing operations
by thermally stable nanofiltration membranes,
Desalination, 160 (2004) 7581.
I. Koyuncu, Reactive dye removal in dye/salt mixtures by nanofiltration membranes containing
vinylsulphone dyes: Effects of feed concentration
and cross flow velocity, Desalination, 143 (2002)
243253.
R. Jiraratananon, A. Sungpet and P. Luangsowan,
Performance evaluation of nanofiltration membranes
for treatment of effluents containing reactive dye and
salt, Desalination, 130 (2000) 177183.
S. Chakraborty, M.K. Purkait, S. DasGupta, S. De
and J.K. Basu, Nanofiltration of textile plant effluent
for color removal and reduction in COD, Sep. Purif.
Technol., 31 (2003) 141151.
J.H. Mo, Y.H. Lee, J. Kim, J.Y Jeong and J. Jegal,
Treatment of dye aqueous solutions using nanofiltration polyamide composite membranes for the
dye wastewater reuse, Dyes Pigments, 76 (2008)
429434.
C. Fersi, L. Gzara and M. Dhahbi, Treatment of
textile effluent by membrane technologies, Desalination, 185 (2005) 399409.
J.-J. Qin, M.H. Oo and K.A. Kekre, Nanofiltration
for recovering wastewater from a specific dyeing
facility, Sep. Purif. Technol., 56(2) (2007) 199203.
Z. Amjad, Reverse Osmosis: Membrane Technology, Water Chemistry, and Industrial Applica-

tions, Van Nostrand Reinhold, New York, 1992.


[60] L. Shu, T.D. Waite, P.J. Bliss, A. Fane and V.
Jegatheesan, Nanofiltration for the possible reuse of
water and recovery of sodium chloride salt from
textile effluent, Desalination, 172 (2005) 235243.
[61] J.G.A. Bitter, Transport Mechanisms in Membrane
Separation Processes, Plenum Press, New York,
1991.
[62] J.H. Choi, K. Fukushi and K. Yamamoto, A submerged nanofiltration membrane bioreactor for
domestic wastewater treatment: the performance of
cellulose acetate nanofiltration membranes for longterm operation, Sep. Purif. Technol., 52 (2007) 470
477.
[63] E.M. Vrijenhoek and J.J. Waypa, Arsenic removal
from drinking water by a loose nanofiltration
membrane, Desalination, 130 (2000) 265277.
[64] R.W. Baker, Membrane Technology and Applications, Wiley, West Sussex, 2004.
[65] J.M.M. Peeters, J.P. Boom, M.H.V. Mulder and H.
Strathmann, Retention measurements of nanofiltration membranes with electrolyte solutions, J.
Membr. Sci., 145 (1998) 199209.
[66] A. Akbari, S. Desclaux, J. C. Remigy and P. Aptel,
Treatment of textile dye effluents using a new
photografted nanofiltration membrane, Desalination,
149 (2002) 101107.
[67] W.R. Bowen and A.W. Mohammad, A theoretical
basis for specifying nanofiltration membranes
Dye/salt/water streams, Desalination, 117 (1998)
257264.
[68] C.J.M. van Rijin, Nano and Microengineered
Membrane Technology, Elsevier, Amsterdam, 2004.
[69] W.L. McCabe, J.C. Smith and P. Harriott, Unit
Oprations of Chemical Engineering, McGraw-Hill,
New York, 2001.
[70] A. Bes-Pia, J.A. Mendoza-Roca, L. Roig-Alcover, A.
Iborra-Clar, M.I. Iborra-Clar and M.I. AlcainaMiranda, Comparison between nanofiltration and
ozonation of biologically treated textile wastewater
for its reuse in the industry, Desalination, 157 (2003)
8186.
[71] B. Van der Bruggen, I. De Vreese and C. Vandecasteele, Water reclamation in the textile industry:
nanofiltration of dye baths for wool dyeing, Ind. Eng.
Chem. Res., 40 (2001) 39733978.
[72] W.J. Thomson, Introduction to Transport Phenomena, Prentice Hall PTR, Englewood, NJ, 2000.

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348


[73] K. Auddy, S. De and S. DasGupta, Flux enhancement in nanofiltration of dye solution using turbulent
promoters, Sep. Purif. Technol., 40 (2004) 3139.
[74] A.C. Gomes, I.C. Goncalves and M.N.D. Pinho, The
role of adsorption on nanofiltration of azo dyes,
J. Membr. Sci., 255 (2005) 157165.
[75] J. Sojka-Ledakowicz, T. Koprowski, W. Machnowski and H.H. Knudsen, Membrane filtration of
textile dyehouse wastewater for technological water
reuse, Desalination, 119 (1998) 110.
[76] A. Bes-Pia, J.A. Mendoza-Roca, M.I. AlcainaMiranda, A. Iborra-Clar and M.I. Iborra-Clar, Reuse
of wastewater of textile industry after its treatment
with a combination of physico-chemical treatment
and membrane technologies, Desalination, 149
(2002) 169174.
[77] M.D. Afonso and M.N. de Pinho, Transport of
MgSO4, MgCl2, and Na2SO4 across an amphoteric
nanofiltration membrane, J. Membr. Sci., 179 (2000)
137154.
[78] W.R. Bowen and H. Mukhtar, Characterisation and
prediction of separation performance of nanofiltration membranes, J. Membr. Sci., 112 (1996)
263274.
[79] A.F. Ismail and A.R. Hassan, The deduction of fine
structural details of asymmetric nanofiltration
membranes using theoretical models, J. Membr. Sci.,
231 (2004) 2536.
[80] A.F. Ismail and A.R. Hassan, Formation and
characterization of asymmetric nanofiltration membrane: Effect of shear rate and polymer concentration, J. Membr. Sci., 270 (2006) 5772.
[81] Z.V.P. Murthy, Estimation of mass transfer coefficient using a combined nonlinear membrane
transport and film theory model, Desalination, 109
(1997) 3949.
[82] X.-L. Wang ,T. Tsuru, M. Togoh, S.-I. Nakao and
S. Kimura, Evaluation of pore structure and electrical properties of nanofiltration membranes,
J. Chem. Eng., Japan, 28 (1995) 186192.
[83] X.-L. Wang, T. Tsuru, S.-I. Nakao and S. Kimura,
Electrolyte transport through nanofiltration
membranes by the space-charge model and the
comparison with Teorell-Meyer-Sievers model,
J. Membr. Sci., 103 (1995) 117133.
[84] C. Labbez, P. Fievet, A. Szymczyk, A. Vidonne,
A. Foissy and J. Pagetti, Retention of mineral salts
by a polyamide nanofiltration membrane, Sep. Purif.

347

Technol., 30 (2003) 4755.


[85] K.M. Pastagia, S. Chakraborty, S. DasGupta, J.K.
Basu and S. De, Prediction of permeate flux and
concentration of two-component dye mixture in
batch nanofiltration, J. Membr. Sci., 218 (2003)
195210.
[86] S. Chakraborty, B.C. Bag, S. DasGupta, J.K. Basu
and S. De, Prediction of permeate flux and permeate
concentration in nanofiltration of dye solution, Sep.
Purif. Technol., 35 (2004) 141152.
[87] I. Koyuncu and D. Topacik, Effect of organic ion on
the separation of salts by nanofiltration membranes,
J. Membr. Sci., 195 (2002) 247263.
[88] N. Al-Bastaki, Removal of methyl orange dye and
Na2SO4 salt from synthetic wastewater using reverse
osmosis, Chem. Eng. Proc., 43 (2004) 15611567.
[89] P. Schirg and F. Widmer, Characterisation of nanofiltration membranes for the separation of aqueous
dye-salt solutions, Desalination, 89 (1992) 89107.
[90] R. Levenstein, D. Hasson and R. Semiat, Utilization
of the Donnan effect for improving electrolyte
separation with nanofiltration membranes, J. Membr.
Sci., 116 (1996) 7792.
[91] I. Machenbach, Membrane technology for dyehouse
effluent treatment, Membrane Technology No. 96.
[92] J.S. Vrouwenvelder, J.A.M. van Paassen, L.P.
Wessels, A.F. van Dama and S.M. Bakker, The
membrane fouling simulator: A practical tool for
fouling prediction and control, J. Membr. Sci., 281
(2006) 316324.
[93] V. Jacquemet, G. Gaval, S. Rosenberger, B. Lesjean
and J.-C. Schrotter, Towards a better characterisation and understanding of membrane fouling in
water treatment, Desalination, 178 (2005) 1320.
[94] A.W. Zularisam, A.F. Ismail and R. Salim, Behaviours of natural organic matter in membrane filtration
for surface water treatmenta review, Desalination,
194 (2006) 211231.
[95] K. Scott, Handbook of Industrial Membranes,
Elsevier Advanced Technology, Oxford, 1995.
[96] I.C. Soung, L.P. Clech, J. Bruce and S. Judd,
Membrane fouling in membrane bioreactors for
wastewater, ASCE., 121(11) (2002) 1018.
[97] I.-C. Kim and K.-H. Lee, Dyeing process wastewater
treatment using fouling resistant nanofiltration and
reverse osmosis membranes, Desalination, 192
(2006) 246251.
[98] A. Asatekin, A. Menniti, S.Kang, M. Elimelech,

348

W.J. Lau, A.F. Ismail / Desalination 245 (2009) 321348

E. Morgenroth and A.M. Mayes, Antifouling


nanofiltration membranes for membrane bioreactors from self-assembling graft copolymers, J.
Membr. Sci., 285 (2006) 8189.
[99] S. Kanga, A. Asatekin, A.M. Mayesc and M.
Elimelech, Protein antifouling mechanisms of
PAN UF membranes incorporating PAN-g-PEO
additive, J. Membr. Sci., 296 (2007) 4250.
[100] K. Boussu, Y. Zhang, J. Cocquyt, P. Van der
Meeren, A. Volodin, C. Van Haesendonck, J.A.
Martens and B. Van der Bruggen, Characterization of polymeric nanofiltration membranes for
systematic analysis of membrane performance, J.
Membr. Sci., 278 (2006) 418427.
[101] K. Kosutic and B. Kunst, RO and NF membrane
fouling and cleaning and pore size distribution
variations, Desalination, 150 (2002) 113120.

[102] L. Shu, A.G. Fane, T.D. Waite, M.T. Pailthorpe


and P.J. Bliss, I: Proc. World Filtration Congress,
London, 2000, pp. 647650.
[103] I. Petrinic, N.P.R. Andersen, S. Sostor-Turk and
A.M.L. Marechal, The removal of reactive dye
printing compounds using nanofiltration, Dyes
Pigments, 74 (2007) 512518.
[104] M.R. Mackley and N.E. Sherman, Cross-flow cake
filtration mechanisms and kinetics, Chem. Eng.
Sci., 47 (1992) 3067.
[105] M. Elimelech and S. Bhattacharjee, A novel
approach for modeling concentration polarization
in cross flow membrane filtration based on the
equivalence of osmotic pressure model and
filtration theory, J. Membr. Sci., 145 (1998) 223
241.

Вам также может понравиться