Vous êtes sur la page 1sur 10

DESALINATION

Desalination 143 (2002) 1 l-20

www.elsevier.comhcate/desal

Nanofiltration of textile wastewater for water reuse

C. Tang, V. Chen*
UNESCO Centre for Membrane Science and Technology, School of Chemical Engineering and Industrial Chemistry,
University of New South Wales, Sydney, NSW 2052, Australia
Tel. +61 (2) 9385-4813; Fax 61 (2) 9385-5966; email: v.chen@unsw.edu.au

Receved 28 August 2001; accepted 6 November 2001

Abstract

The textile industry produces a large amount of wastewater that is highly coloured with high loading of inorganic
salt. Crossflow nanofiltration using thin film composite polysulfone membrane was used to recover the electrolyte
solution and reject the colour. Using a synthetic textile effluent of reactive dye and NaCl solution, the study focused
on the mechanism controlling flux and rejection by varying four main parameter; crossflow velocity, initial dye
concentration, feed pressure, and electrolyte concentration. Results show that flux was dominated by the osmotic
pressure created from the presence of NaCI, and that dye concentration did not significantly effect the flux or
rejection. Working at low pressures of up to 500 kPa, relatively high fluxes were obtained, with an average dye
rejection of 98% and NaCl rejections of less than 14%. Thus, a high quality of reuse water could be recovered. Even
after a number of cycles, the membrane did not foul irreversibly, with an overall mean water-flux recovery of 99%.

Keywords: Textile wastewater; Nanofiltration; Salt reuse; Osmotic pressure; Mass transfer

1. Introduction
agricultural activities. However, with an increasing
The world population is ever increasing putting population, there will be pressure for industries
a considerable amount of stress on the environ- to reclaim and reuse some of its wastewater, or
ment. Water will continue to become a major face the prospect of being shutdown. This is due
factor for the survival of humans and human to the combine pressures of increasing water and
activities. This is especially true in the industri- wastewater costs and increasing regulatory
alised areas of the world. At present, appro- requirements of discharged wastewater [ 11.
ximately 50% of the water is being used by The textile industry is one that demands large
households, and the other 50% for industrial and quantities of water, and produce large amounts
of wastewater. The wastewater is characterised
*Corresponding author. by strong colour and a high concentration of dis-

001 l-91 64/02/$- See front matter 0 2002 Elsevier Science B.V. All rights reserved
F'II:SOOll-9164(02)00216-3
12 C. Tang, V Chen /Desalination 143 (2002) 11-20

solved solids (organic and inorganic). Colour is the ocean. However, these have been proven to
not typically removed by conventional waste- be environmentally or economically unacceptable
water treatment, but mixed with rinse water to [5]. There are studies into treatment of highly
be diluted before it is discharged to sewage systems. concentrated textile effluent, including ozonation
Studies with unconventional treatment systems and Fentons oxidation [2,4,6]. However, these
to treat textile wastewater have been widespread. methods are expensive to operate and produce
These treatment techniques used include adsorption, undesirable side reactions and products [2,7].
ozonation, photocatalysis, and filtration. They all Filtration processes offer the reuse of water and
have varying degrees of success and all have their salt, while minimising the amount of effluent that
limitations. is discharged.
The most widely used treatment systems are Research has shown that there are three main
conventional activated sludge. These systems terms controlling the mass transfer for nano-
poorly removes the widely used reactive dyes [ 11, filtration. Due to its small pore size, the observed
and is clearly ineffective in decolourising textile mass transfer mechanism for NF is diffusion and
effluents, even when mixed and treated together convection. In addition, the active membrane layer
with sewage [2]. This leads to highly coloured normally consists of negatively charged chemical
water at the point of discharge, and complaints groups, thus mass transfer via migration of ions
from the public. Powered activated carbon (PAC) in an electric field must also be considered [8].
is the most commonly used and most successful The extended Nernst-Planck equation [9] has
adsorbent. However, PAC is expensive and the been used extensively in describing mass transfer
level of colour removal depends on the dye type. across a nanofilter, as it takes into account all the
Also, 100% colour removal is rarely achieved [3]. three terms. Even though all three forms of mass
Ozone has been shown to have the ability to transfer may exist during filtration, only one form
breakdown most dyes. However, even high doses is expected to dominate [lo]. The domination will
of ozone do not completely mineralise the organic depend on solute and membrane characteristics.
dye to carbon dioxide and water. This is due to Thus, nanofilters may be viewed as separation
the decolourisation rate decreasing with increasing due to molecular sieve effect and charge effect.
initial dye colour [4]. Ultrafiltration has been successfully applied
Filtration process is one of the popular treat- in many industries, but it has not been widely
ments used in textile industry. Although filtration accepted by the textile industry [ 111.One reason is
techniques require an initial high setup cost, it is the variability of the rejection coefficient, ranging
outweighed by the significant costs savings from 30 to 90%. This makes direct reuse impossible
achieved through the reuse of salts and permeate. and will require further filtration by either NF or
Costs are reduced by the use of pre-filters, regular RO. Groves and Buckley [ 121studied such a system
cleaning to eliminate membrane fouling problems, and found that reverse osmosis had problems with
and by choosing the most appropriate membrane fouling, which resulted in low fluxes and poor
system. Ultrafiltration (UQ, reverse osmosis (RO), separation.
and nanofiltration (NF) have been widely used There is growing interest for use of nano-
for the full-scale treatment and re-use of chemicals filtration for the minimisation and reuse of textile
and water in South Africa. One severe limitation effluents. Erswell [13], was one of the first to
of the application of filtration technologies is the investigate the prospect of a charged membrane
disposal of the retentate or concentrate stream. for the reuse of reactive dye liquor. Charged ultra-
At the moment, retentate stream is disposed by filtration (another name for nanofiltration) mem-
evaporation, incineration or discharging it into brane was used in a closed-loop recycling system.
C. Tang, L! Chen /Desalination 143 (2002) 11-20 13

Membrane performance was monitored in terms brine from textile effluent may be more feasible
of dye and salt rejection and permeate flux, under than that reported.
varying conditions. They concluded that charged This paper will investigate the mechanisms
ultrafiltration is a technically feasible process for controlling flux and rejection by varying the
treatment of the dye bath at high water recoveries. pressure, NaCl concentration and dye concen-
Also, the electrolyte recovered in the permeate tration in the simulated dye effluent. The maximum
is valued an order of magnitude greater than that pressure used is 500 kPa, which is significantly
of fresh water, making the process more econo- lower than those used in previous studies, whilst
mically favourable. However, low fluxes were still achieving a relatively high flux.
achieved, maximum flux of 30 L/m2h, even when
high pressure was used (up to 4 MPa). Thus, 2. Materials and methods
improvements can still be made to the process
parameters. 2.1. Materials
Xu [14] characterised a nanofiltration mem- A flat sheet polysulfone based thin film com-
brane by its pure water permeability, mass transfer posite (TFC-SR2) nanofilter, supplied by Fluid
coefficient of NaCl, and the mean radius of the Systems, California was selected to perform the
membrane pores. Dye filtration experiments were dye separation. Synthetic dyebaths consisting of CI
performed with both pure and industrial dye pulp reactive black 5 (Bayer, Sydney), salt (NaCl) and
solutions at a maximum pressure of 4 MPa. Milli-Q water were made up for each filtration.
Fluxes of up to 220 Wm2h and dye rejection of The reactive black 5 structure is shown in Fig. 1.
greater than 98.5% were achieved. The study Dye and salt were analytically weighed out to
concluded that the identification of transfer desired concentrations, depending on the set para-
phenomena in dye solution during nanofiltration meters. However, dye and salt concentrations were
is complex. No dominant transfer mechanism was very had to replicate exactly due to the presence
given. of water in the nanofiltration system, which could
Van der Bruggen [ 151examined the mechanisms not be completely drained out. Thus concentra-
of retention and flux decline of nanofiltration tions were check again after circulation of the
membranes using a synthetic dye bath consisting feed solution through the system. Sodium hydroxide
of reactive blue 2, reactive orange 16, Na,SO,, and hydrochloric acid was used to adjust the
Na,CO,, NaOH and a surfactant. The study found dyebath pH to 7. Ambient temperature was main-
that rejection of ions decreased with the ion con- tained for the duration of the experiment. Dye
centration, and the dye concentration or the ionic concentration was measured using the absorbance
strength did not influence the high rejection of at 596 nm via a Cat-yIE W-Vis spectrophotometer
dye. Due to sigificant rejection of the salts, the by Varian. The dye solution had to be diluted by
flux dropped rapidly with increasing salt concen- a factor of ten in order to measure its absorbance
tration, reducing the feasibility of nanofiltration within the limits of the spectrophotometer. Con-
of this effluent at the higher salt concentrations. centration of sodium chloride was measured with
However, this study was done using a divalent a conductivity meter, EcoScan Con 5 from
electrolyte Na$O,, and results of the more widely EUTECH Instruments (Singapore).
used NaCl salt was not given. In addition to
potential lower rejection, NaCl salt has a lower
2.2. Apparatus
ionic strength, creating a much lower osmotic
pressure resistance than Na,SO,. Thus, factors Crossflow nanofiltration (NF) experiments
suggest that the use of nanofiltration to recover were carried out in a single flat sheet stainless
14 C. Tang, V Chen /Desalination 143 (2002) II-20

Fig. 1. CI reactive black 5 structure, molecular weight = 991 Da.

steel membrane module with channel 1.4 mm the dye solution and allows for final pH adjust-
high, 25 mm wide, and 232 mm long. The effective ments to be made. Regular samples of permeate
membrane surface area was 58 cm2. Acentrifuge and concentrate were taken at different time
pump (Cat Pumps Model 357, with 3-phase intervals. The system is then washed out with tap
induction motor) was used to supply the operating water until the recycled water is clear. Finally,
pressure and feed circulation. The flowrate was Mini-Q water is again used to measure the final
monitored using a rotameter and was adjusted via water flux. This procedure was then repeated for
the pump speed. A throttle valve at the outlet of subsequent experiments with different parameters.
the module was used to control the feed pressure
into the modules, shown by the pressure gauge.
2.3. Membrane characterisation
Since the pressure used is so high, the pressure
drop was assumed to be equal to the module inlet The nanofiltration membrane was character-
pressure. The concentrate stream is recycled back ised using 2 different methods. Atomic Force
to the feed tank and the permeate flow was Microscopy (AFM) (Nanoscope, Digital Instru-
measured with an electronic balance and monitored ments) was used to scan the surface of the mem-
by a computer. The feed solution temperature was
keep constant using the cooling coils and thermo-
meter. The schematic of this process is shown in
Fig. 2. Pressure and crossflow were modified
depending on the experimental specifications, and
varied between 100-500 kPa and 3-5 L/min
respectively.
Each experiment consists of three flux filtra-
tions. Firstly, the initial water flux was measured
Feed hmk
using Mini-Q water, with the process taking 15-
30 min depending on the stability of the flux. The
water is then evacuated from the system,
removing as much water as possible. Secondly, a
5-L dye solution is fed to the feed tank and Duta Lagger

circulated for 15 min before filtration begins. This


insures all the residue water is well mixed with Fig. 2. Schematic diagram of NF experiment.
C. Tang, I! Chen /Desalination 143 (2002) II-20 15

brane for it morphology and determine whether


any pores could be seen, and if so, measure the
pore size. Fig. 3 shows the result of an AFM scan,
showing that the surface of the membrane is too
rough for any pores to be detected.
Electra-Kinetic Analyser, EKA (Anton Paar)
was used to determine the membrane isoelectric
point and surface charge characteristics. This was
done for several similar membranes to allow for
comparison. Fig. 4 shows the zeta potential of
four different TFC membranes. The isoelectric
point of the TFCSR2 membrane is 4.7. The zeta
potential curves for the three TFC-SR membrane
are similar, whereas the TFC-S membrane is Fig. 3. AFM image of nanofikration membrane.
always more negatively charged at all pH ranges.
This is due to different polymer coatings on the
membrane surface.

3. Results and discussion


3. I. Eflect of dye concentration
To investigate the effect of dye concentration,
experiments were performed using different dye
concentrations, at different pressure and flowrate. PH

The NaCl concentration used was kept constant Fig. 4. Zeta potential measure for TFC membranes.
at 20 g/L for these experiments. The expectation
is for flux to decrease with increasing dye concen-
tration, due to the increased polarization of the was observed to only increase slightly with in-
dyes on the membrane surface; thus, a higher creasing dye concentration. Hence, for these con-
probability of fouling to occur. The dye rejection ditions, initial dye concentration will not effect
is also expected to be higher at lower feed concen- the flux or dye rejection significantly, which is
tration and lower at higher feed concentration. consistent with finding from a previous study
The results are shown in Table 1.With increasing [ 151.This is due to the good mass transfer across
dye concentration (at P = 200 kPa, Q = 3 L/min), the membrane surface that does not allow for
the flux values was found to remain reasonably con- sufficient build-up of a dye concentration polariz-
stant, varying between 23.5 L/m*h to 25.07 L/m*h. ation layer for fouling to occur. Comparisons
The dye rejection values were also constant. This between flowrate at 3 L/min and 5 L/min at
is clearly shown in Fig. 5. Increasing the pressure pressure of 200 kPa clearly shows that increasing
and was used to study the build up of a foulant flowrate does not increase the flux. One reason
layer. Pressure was increased to 500 kPa and for this may be the Reynolds number, 3600 for
flowrate to 5 L/min. Once again, the results show the 3 L/min and 6000 for 5 L/min, achieved in
that with increasing dye concentration, the flux the module which are in the turbulent region.
and dye rejection remained stable. Dye rejection Thus, the flow in the module channel has already
16 C. Tang, V Chen /Desalination 143 (2002) 11-20

Table 1
Effect of initial dye concentration on flux and dye rejection at different pressure and flowrate

Dye concentration, ppm Pressure, kPa Flowrate, Wmin Flux, Wm*h Dye rejection, %
82 200 3 23.85 97.2
147 200 3 25.07 97.2
398 200 3 23.5 97.1
705 200 3 24.27 97.4
448 200 5 23.12 97.6
92 500 5 59.58 97.5
188 500 5 62.72 97.8
455 500 5 55.75 98.1
708 500 5 59.76 98.4
890.8 500 5 59.58 98.5
1583 500 3 78.4 98.1

reached turbulence at 3 L/min, and


maximum in this study from the water recovery of the used
increasing the flowrate will not increase the actual membranes. The used membranes, with a blue
mass transfer and flux. This holds true even when coloured surface, were taken for further analysis
the initial dye concentration was doubled (Table 1) using the Electrokinetic analyser to see if any
to 1583 ppm. The filtration was done using a dif- change in electrostatic interactions have occurred
ferent piece of membrane, showing the flux varia- due to the adsorption of dyes. The zeta potential
bility that can occur due to the manufacturing plot is shown in Fig. 6.
process. However, the filtration flux to pure water Used membranes number 1 and 3 were analyzed
flux ratio (J/Jiw) was within the same range as after filtration of alkaline solutions, whereas
the other experiments. number 2 was analyzed after filtration of a neutral
Van der Bruggen [ 151reported that the mem- pH solution. The plot shows little to no effect on
brane fouling was only limited to a colour change the membrane surface charge after filtration of
on the membrane surface, which was also seem dye and salt solutions. However, after filtration

l New TFC-SR2 batch 2 Usedl,pH= I1


A Used 2,200Oppm. pH = 7 l Used 3, pH = 10.7

-0- 023 Llmin, P=2OOkPa -O- 0-5 L!min, P-WOkPa

0 ?UU 4011 bUU suu IUUU

Dye concentration (ppm)

Fig. 5. Relationship between dye concentration and flux at Fig. 6. Comparison of Zeta potentials of new and used
different pressure and flowrate. membranes.
C. Tang, b! Chen /Desalination 143 (2002) 11-20 17

of an alkaline solution, the membrane surface concentration polarisation layer, placing little
becomes less negatively charged at higher pHs. importance on this phenomenon. However, with
Nevertheless, the coloured, used membranes did the 20 g/L salt in the dyebath, the flux does not
not exhibit any charge changes to the membrane increase as rapidly as does normal waterflux. As
surface, nor did it change the dye and salt rejection established earlier, dye concentration does not
characteristics. Therefore, the membrane must have effect flux, thus the waterflux line can be assume
tight uniform pores since near uniform dye rejec- to also be the same as that observed with dye
tion was maintained with increasing dye concen- filtration with no salt. Thus despite the high mass
tration. This membrane has the pore structure and transfer and low rejection of the salt, the salt must
surface charge, which is highly suitable for NaCl therefore exhibit some kind of resistance, most
filtration since it allows about 90% of NaCl to likely in the form of osmotic pressure in the pola-
permeate while retaining the dye. rization layer.

3.2. EfSect of pressure 3.3. Effects of salt and osmotic pressure d#erence
With an increase in pressure, flux is expected Osmotic pressure difference (Dp) of NaCl was
to increase accordingly (solution-diffusion model). calculated using the osmotic pressure model, with
The increase in feed pressure will increase the the wall concentration calculated using the fitted
driving force, overcoming membrane resistance. Sourirajan correlation [ 161.Fig. 8 shows the effect
Fig. 7 shows this phenomena where flux increases of Dp with increasing salt concentration, with an
linearly with increasing pressure and also shows applied feed pressure is 500 kPa for the five expe-
that the mass transfer coefficient relationship with riments. The osmotic pressure difference can be
pressure. seen to increase more rapidly at lower salt con-
The mass transfer coefficient remains constant centrations, and slowly taper off with increasing
between the pressures of 100 to 300 kPa, and salt concentrations, signaling near a limiting
increases rapidly after 300 kPa. Both these relation- osmotic pressure resistance. This effect can also
ships imply that mass transfer is not a limiting be seen in the percent reduction in pressure line.
factor with regards to the high operating pressure. With increasing osmotic pressure, the effective
Also, it signifies the under-development of the driving force of the system decreases accordingly.

200 300 0 10 20 30 40 50 60 70 80 90
Presrurc (kPa)
CsB t&L)
Fig. 7. Effect of pressure on flux and mass transfer (dye
concentration = 500 ppm, salt concentration = 20 g/L, Fig. 8. Osmotic pressure effects (dye concentration = 500 ppm,
Q = 5 L/min). feed pressure = 500 kPa, Q = 3 Umin).
18 C. Tang, V Chen /Desalination 143 (2002) II-20

Thus, with respects to Fig. 7, a feed solution of


20 g/L salt showed a reduction of driving force
caused by osmotic pressure resistance, from 500 kPa
to 350 kPa. At the highest salt concentration, the
feed pressure had been reduced by more than half.
Fig. 9 shows the effect of bulk salt concen-
tration (C,) on wall salt concentration (C,, and
mass transfer coefficient. CsBcould be seen to be 0 20 40 60 80

directly proportional to Csw.In fact, the calculated CS'WU

salt wall concentration was only marginally Fig. 9. The effect of bulk salt concentration on wall salt
higher than that of the bulk salt concentration. concentration and mass transfer coefficient (dye concentration
The ratio of Csw to CsBwas close to 1:l at an = 500 ppm, feed pressure = 500 kPa, Q = 3 Llmin).
operating pressure of 500 kPa. This finding was
supported by Xu et al. [ 141, and further substan-
tiated the fact that the phenomenon of concen- loo

r-*----=- toll
~ll-.x-~--.-.

tration polarisation is not significant. However, YO YO

,.,!
even the small differences between the bulk and
wall concentrations at these high concentrations
of salts was sufficient to generate substantial
osmotic pressure differences.
-$ \
The mass transfer coefficient was high at low 20 I 20
c
salt concentrations, but decreased sharply at NaCl IO
k
IO
0 L...
^"_,,.,,,,"._
.._^.~_"^
"..^
^ ._^^_
,.."....-~........._
.._
......'_.._
.__..-......
0
concentrations increased from 10 to 30 g/L, before
0 10 20 30 40 SO 60 70 80
rising slightly with NaCl concentrations greater
csnwL)
than 30 g/L. An average mass transfer coefficient
of 5.8x 1O-5m/s was reported for NaCl at pressures Fig. 10. Effect of salt concentration (dye concentration = 383-
above 1 MPa [ 141.A similar order of magnitude 455 ppm, Q = 3 Lhnin, pressure = 500 kPa).
was found at a feed pressure of 500 kPa, but only
with the higher salt concentrations (>30 g/L).
Fig. 10 shows how dye and salt rejection is other words, the phenomena known as dye
effected by salt concentration. With increasing aggregation [ 17,181 does not exist in this system.
bulk salt concentration, both dye and salt rejection This may be due to higher mass transfer regime
slowly decrease. Dye rejection varied from 98.2% used in this system compared to the stirred cell
and 96.1% from low to high salt concentration. configuration where the interactions between dye
Dye rejection was shown to remain constant with and salt rejections have been observed [ 181.
increasing salt concentration. Slightly lower dye Salt rejection varied from 13.7% and 2.9%
rejections were seen for higher dye concentra- with increasing salt concentration. Repulsive
tions. This was also seen from the results in force from the membrane (negatively charged)
Table 1. Thus, if the same dye concentrations were decreases with increasing salt concentration due
used, dye rejection remained almost constant to higher concentrations of Na+ cations on the
regardless of what salt concentration is used. It membrane surface. Overcoming the repulsive
could therefore be concluded that there is no force also allows more Cl- anions to pass through
significant interaction between dye and salt to the membrane which effectively means that more
cause an apparent change in the dye rejection. In NaCl will permeate, lowering the salt rejection.
C. Tang, c! Chen /Desalination 143 (2002) II-20 19

This is the principle of Donnan equilibrium. At


these high electrolyte concentrations, the repulsive
forces would not extend far into solution; however,
the change of rejection with increasing salt concen-
trations indicated that some charge interactions
may still contribute to the rejection mechanism
within tight pores. Salt rejection was found to
increase with increasing feed pressure, which was I 2 3 4 5 6 7 8 9 IO 11 12 13 14 1.5 IG
also discovered by Erswell [13] and Schirg and Cycle number
Widmer [ 191 at much higher pressures. This
Fig. 11. Waterflux recovery.
increase is due to the fact that salt rejection is a
function of both salt concentration and operating
pressure. salt concentrations. For the relatively low pressure
If the permeate contains high concentrations used (500 kPa), the flux obtained was much
of salt, less supplemental salt would have to be higher than that obtained in previous studies. Flux
added in the dyeing recipe, reducing cost. However, depended on mainly on osmotic pressure effects.
lower salt rejections are obtained at lower fluxes. Significant osmotic pressures differences were
The higher the flux required, less membrane is generated even with small differences in wall and
required, thus it may become a trade off between permeate electrolyte concentrations. A high dye
a high flux with higher salt rejection; or to use a rejection value was also achieved, giving a decolour-
lower pressure, obtaining lower flux, but also ised permeate, suitable for possible reuse or
lower salt rejection. further polishing. In addition to the low pressures
used in this process, the low salt rejections offers
3.4. Water-fluxrecovery promise for direct nanofiltration of the effluent.
From these results, the dominant mass transfer
Fig. 11 shows water flux recovery over 16 con-
mechanism for NF appear to be due to convection
secutive cycles using the same membrane. For
through the membrane pores (steric rejection) for
each cycle, three procedures were performed.
the dye; however, charge interactions within small
Firstly, the initial watefflux was measured with
pores may still play a role in the rejection of NaCl
Milli-Q water, then the dye solution was filtrated,
even at very high salt concentrations.
and thirdly the final water-flux was taken. The
mean initial waterflux is 86.55 L/m*h, and the
mean final waterflux is 86 L/m*h. Generally Acknowledgement
waterflux was found to be recoverable, with the
The authors acknowledge financial support
lowest recovery value of 92%, and an overall
from the CRC for Waste Management and
mean recovery of 99%. This meant that there was
Pollution Control Ltd. The authors also thank
no permanent fouling on or within the membrane.
Bayer for supply of the dyes and Fluid Systems
The concentration polarization layer formed by
for the nanofiltration membranes.
the dye and salt mixture can be easily removed
by crossflow or relaxation operations.
References
4. Conclusion U. Pagga and K. Taeger, Development of a method
[l]
The nanofiltration system has been proven to for adsorption of dyestuff on activated sludge, Water
be well suited for the treatment of high dye and Research, 28 (1994) 1051-1057.
20 C. Tang, V Chen /Desalination 143 (2002) 11-20

PI PC. Vandevivere, R. Bianchi and W. Verstraete, The University of New South Wales, Sydney, 1997,
Treatment and reuse of wastewater from the textile 124 p.
wet-processing industry: Review of emerging [ 1I] J.C. Watters, E. Biagtan and 0. Senler, Ultrafiltration
technologies, J. Chem. Tech. Biotech., 72(4) (1998) of a textile plant effluent, Separation Sci.& Tech.,
289-302. 26(1&11) (1991) 1295-1313.
131 A. Rozzi, F. Malpei, L. Bonomo and R. Bianchi, [ 121 GR. Groves and C.A. Buckley, Treatment of textile
Textile wastewater reuse in northern Italy effluents by membrane separation processes, Proc. 7th
(Como),Water Sci. Tech., 39(5) (1999) 121-128. Int Symp. Fresh Water form the Sea, Amsterdam, 2
[41 J.N. Wu, M.A. Eiteman and S.E. Law, Evaluation of (1980) 249-257.
membrane filtration and ozonation processes for treat- [13] A. Erswell, C.J. Brouchaert and C.A. Buckley, The
ment of reactive dye wastewater, J. Environ. Eng.- reuse of reactive dye liquors using charged ultra-
ASCE, 124(3) ( 1998) 272-277. filtration membrane technology, Desalination, 70
PI L.R. Gravelet-Blondin, C.M. Carliell, S.J. Barclay and (1988) 157-167.
C.A. Buckley, Management of water resources in [ 141 Y. Xu, R. Lebrun, P.J. Gal10 and P. Blond, Treatment
South Africa with respect to the textile industry,Proc. of textile dye plant effluent by nanofiltration mem-
2nd IAWQ Specialised Conference on Pretreatment brane, Separation Sci. &Tech., 34(13) (1999) 2501-
of Industrial Wastewaters, Athens, Greece, 16-18 2519.
October 1996. [ 151 B. Van der Bruggen et al., Mechanisms of retention
WI F. Gahr, E Hermanutz and W. Oppermann, Ozona- and flux decline for the nanofiltration of dye baths
tion - an important technique to comply with new from the textile industry, Separation & Purification
German laws for textile wastewater treatment, Water Tech., 22-23(1-3) (2001) 519-528.
Sci. Tech., 30 (1994) 255-263. [ 161 S. Sourirajan, Reverse Osmosis. New York, Academic
[71 W.H. Glaze et al. , Byproducts of oxidation processes Press, 1970,580 p.
in water and wastewater treatment. BookofAbstracts, [17] L. Shu, T.D. Waite, A.G Fane, M.T. Pailthrope and
210th ACS National Meeting, Chicago, IL, August P.J. Bliss, Membrane processing of dye wastewater:
20-24, 1995, P. 1, p. ENVR-073. effect of dye aggregates. In: Environmental Engi-
@I T. Tsuru, U. Masakatsu, S. Nakao and S. Kimura, neering Research Event, Castlemaine, Victoria: EERE
Negative rejection of anions in the loose reverse Organising team, 1999.
osmosis separation of mono- and divalent ion [ 181 L. Shu, Membrane processing of Dye Wastewater: Dye
mixtures, Desalination, 81 (1991) 219-231. Aggregation, Membrane Performance, and Mathe-
PI J.M.M. Peeters, J.P. Boom, M.H.V. Mulder and H. matical Modelling, Ph. D. Thesis, The University of
Strathmann, Retention measurements of nanofiltration New South Wales, 2001.
membranes with electrolyte solutions, J. Membr. Sci., [ 191 P Schirg and F. Widmer, Characterisation of nano-
145 (1998) 199-209. filtration membranes for the separation of aqueous
[lo] A. Lai, The use of membranes for the recovery of dye-salt solutions, Desalination, 89 (1992) 89-107.
reactive dyes. In: School of Chemical Engineering.

Vous aimerez peut-être aussi